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

Summability of formal solutions for some generalized moment partial differential equations

Alberto Lastra Departamento de Física y Matemáticas
University of Alcalá
Ap. de Correos 20, E-28871 Alcalá de Henares (Madrid), Spain
[email protected]
Sławomir Michalik Faculty of Mathematics and Natural Sciences, College of Science
Cardinal Stefan Wyszyński University
Wóycickiego 1/3, 01-938 Warszawa, Poland
[email protected] http://www.impan.pl/~slawek
 and  Maria Suwińska Faculty of Mathematics and Natural Sciences, College of Science
Cardinal Stefan Wyszyński University
Wóycickiego 1/3, 01-938 Warszawa, Poland
[email protected]
Abstract.

The concept of moment differentiation is extended to the class of moment summable functions, giving rise to moment differential properties. The main result leans on accurate upper estimates for the integral representation of the moment derivatives of functions under exponential-like growth at infinity, and appropriate deformation of the integration paths. The theory is applied to obtain summability results of certain family of generalized linear moment partial differential equations with variable coefficients.

Key words and phrases:
summability, formal solution, moment estimates, moment derivatives, moment partial differential equations
2010 Mathematics Subject Classification:
35C10, 35G10

1. Introduction

This work is devoted to the study of the summability properties of formal solutions of moment partial differential equations in the complex domain. The purpose of this work is twofold. On the one hand, a deeper knowledge on the moment derivative operator acting on certain functional spaces of analytic functions is put into light; and on the other hand, the previous knowledge serves as a tool to attain summability results of the formal solutions of concrete families of Cauchy problems.

The study of moment derivatives, generalizing classical ones, and the solution of moment partial differential equations is of increasing interest in the scientific community. The concept of moment derivative was put forward by W. Balser and M. Yoshino in 2010, in [2]. Given a sequence of positive real numbers (in practice a sequence of moments), say m:=(m(p))p0m:=(m(p))_{p\geq 0}, the operator of moment derivative m,z:[[z]][[z]]\partial_{m,z}:\mathbb{C}[[z]]\to\mathbb{C}[[z]] acts on the space of formal power series with complex coefficients into itself in the following way (see Definition 9):

m,z(p0apm(p)zp)=p0ap+1m(p)zp.\partial_{m,z}\left(\sum_{p\geq 0}\frac{a_{p}}{m(p)}z^{p}\right)=\sum_{p\geq 0}\frac{a_{p+1}}{m(p)}z^{p}.

This definition can be naturally extended to holomorphic functions defined on a neighborhood of the origin.

The choice m=(Γ(1+p))p0=(p!)p0m=(\Gamma(1+p))_{p\geq 0}=(p!)_{p\geq 0} determines the usual derivative operator, whereas m=(Γ(1+ps))p0m=\left(\Gamma\left(1+\frac{p}{s}\right)\right)_{p\geq 0} is linked to the Caputo 1/s1/s-fractional differential operator z1/s\partial_{z}^{1/s} (see [17], Remark 3). Given q(0,1)q\in(0,1) and m=([p]q!)p0m=([p]_{q}!)_{p\geq 0}, with [p]!q=[1]q[2]q[p]q[p]!_{q}=[1]_{q}[2]_{q}\cdots[p]_{q} and [h]q=j=0h1qj[h]_{q}=\sum_{j=0}^{h-1}q^{j}, the operator m,z\partial_{m,z} coincides with the qq-derivative Dq,zD_{q,z} defined by

Dq,zf(z)=f(qz)f(z)qzz.D_{q,z}f(z)=\frac{f(qz)-f(z)}{qz-z}.

Several recent studies of the previous functional equations have been made in the complex domain and in terms of summability of their formal solutions, such as [16] regarding summability of fractional linear partial differential equations; [7, 8] in the study of difference equations; or [5, 6, 12] in the study of qq-difference-differential equations.

In the more general framework of moment partial differential equations, the seminal work [2] was followed by other studies such as [17] where the second author solves certain families of Cauchy problems under the action of two moment derivatives. We also refer to [18, 19] and [25] (Section 7), where conditions on the convergence and summability of formal solutions to homogeneous and inhomogeneous linear moment partial differential equations in two complex variables with constant coefficients are stated. Further studies of moment partial differential equations with constant coefficients are described in [13], or in [21] when dealing with the Stokes phenomenon, closely related to the theory of summability. We also cite [14], where the moments govern the growth of the elements involved in the problem under study.

A first step towards the study of summability of the formal solution of a functional equation is that of determining the growth rate of its coefficients, which is described in the works mentioned above, and also more specifically in the recent works [15, 20, 26] when dealing with moment partial differential equations with constant and time-dependent coefficients. See also the references therein for a further knowledge on the field.

The present work takes a step forward into the theory of generalized summability of formal solutions of moment partial differential equations. The first main result (Theorem 3) determines the integral representation of the moment derivatives (mm-derivatives) of an analytic function defined on an infinite sector with the vertex at the origin together with a neighborhood of the origin, with prescribed exponential-like growth governed by a second sequence, say 𝕄~\tilde{\mathbb{M}}. In addition to this, accurate upper estimates of such derivatives are provided showing the same exponential-like growth at infinity, but also its dependence on the moment sequence mm. This result entails that the set of 𝕄~\tilde{\mathbb{M}}-summable formal power series along certain direction dd\in\mathbb{R}, {z}𝕄~,d\mathbb{C}\{z\}_{\tilde{\mathbb{M}},d} (see Definition 8 and Theorem 2), is closed under the action of the operator m,z\partial_{m,z}. As a consequence, it makes sense to extend the definition of m,z\partial_{m,z} to {z}𝕄~,d\mathbb{C}\{z\}_{\tilde{\mathbb{M}},d} (Definition 10) and also to provide analogous estimates as above for the mm-derivatives of the elements in {z}𝕄~,d\mathbb{C}\{z\}_{\tilde{\mathbb{M}},d} (Proposition 2).

We apply the previous theory to achieve summability results on moment partial differential equations of the form

(1) {(m1,tka(z)m2,zp)u(t,z)=f^(t,z)m1,tju(0,z)=φj(z),j=0,,k1,\left\{\begin{array}[]{lcc}\left(\partial_{m_{1},t}^{k}-a(z)\partial_{m_{2},z}^{p}\right)u(t,z)=\hat{f}(t,z)&\\ \partial_{m_{1},t}^{j}u(0,z)=\varphi_{j}(z),&\quad j=0,\ldots,k-1,\end{array}\right.

where 1k<p1\leq k<p are integer numbers and m1,m2m_{1},\,m_{2} are moment sequences under additional assumptions. The elements a(z),a(z)1,φj(z)a(z),\,a(z)^{-1},\,\varphi_{j}(z) for j=0,,k1j=0,\ldots,k-1 are assumed to be holomorphic functions in a neighborhood of the origin, and f^[[t,z]]\hat{f}\in\mathbb{C}[[t,z]]. The second main result of this research (Theorem 4) states that summability of the unique formal solution of (1) u^(t,z)\hat{u}(t,z) (with respect to zz variable) along direction dd\in\mathbb{R} is equivalent to summability of f^\hat{f} and m2,zju^(t,0)\partial_{m_{2},z}^{j}\hat{u}(t,0), for j=0,,p1j=0,\ldots,p-1, along dd. A result on the convergence of the formal solution is also provided (Corollary 2). It is worth mentioning that the results on the upper estimates of formal solutions obtained in [15] remain coherent with these results, and also with those in [23], in the Gevrey classical settings. The study of more general moment problems remains open and it is left for future research of the authors.

The paper is structured as follows: After a section describing the notation followed in the present study (Section 2), we recall the main concepts and results on the generalized moment differentiation of formal power series. Section 3.1 is devoted to recalling the main tools associated with strongly regular sequences and some of their related elements. In Section 3.2, based on the general moment summability methods, we state the first main result of the paper (Theorem 3) and its main consequences. The work is concluded in Section 4 with the application of the theory to the summability of formal solutions of certain family of Cauchy problems involving moment partial differential equations in the complex domain (Theorem 4).

2. Notation

Let \mathbb{N} denote the set of natural numbers {1,2,}\{1,2,\cdots\} and 0:={0}\mathbb{N}_{0}:=\mathbb{N}\cup\{0\}.

\mathcal{R} stands for the Riemann surface of the logarithm.

Let θ>0\theta>0 and dd\in\mathbb{R}. We write Sd(θ)S_{d}(\theta) for the open infinite sector contained in the Riemann surface of the logarithm with the vertex at the origin, bisecting direction dd\in\mathbb{R} and opening θ>0\theta>0, i.e.,

Sd(θ):={z:|arg(z)d|<θ2}.S_{d}(\theta):=\left\{z\in\mathcal{R}:|\hbox{arg}(z)-d|<\frac{\theta}{2}\right\}.

We write SdS_{d} in the case when the opening θ>0\theta>0 does not need to be specified. A sectorial region Gd(θ)G_{d}(\theta) is a subset of \mathcal{R} such that Gd(θ)Sd(θ)D(0,r)G_{d}(\theta)\subseteq S_{d}(\theta)\cap D(0,r) for some r>0r>0, and for all 0<θ<θ0<\theta^{\prime}<\theta there exists 0<r<r0<r^{\prime}<r such that (Sd(θ)D(0,r))Gd(θ)(S_{d}(\theta^{\prime})\cap D(0,r^{\prime}))\subseteq G_{d}(\theta). We denote by arg(S)\hbox{arg}(S) the set of arguments of SS, in particular arg(Sd(θ))=(dθ2,d+θ2)\hbox{arg}(S_{d}(\theta))=\left(d-\frac{\theta}{2},d+\frac{\theta}{2}\right).

We put S^d(θ;r):=Sd(θ)D(0,r)\hat{S}_{d}(\theta;r):=S_{d}(\theta)\cup D(0,r). Analogously, we write S^d(θ)\hat{S}_{d}(\theta) (resp. S^d\hat{S}_{d}) whenever the radius r>0r>0 (resp. the radius and the opening r,θ>0r,\,\theta>0) can be omitted. We write SSd(θ)S\prec S_{d}(\theta) whenever SS is an infinite sector with the vertex at the origin with S¯Sd(θ)\overline{S}\subseteq S_{d}(\theta). Analogously, we write S^S^d(θ;r)\hat{S}\prec\hat{S}_{d}(\theta;r) if S^=SD(0,r)\hat{S}=S\cup D(0,r^{\prime}), with SSd(θ)S\prec S_{d}(\theta) and 0<r<r0<r^{\prime}<r. Given two sectorial regions Gd(θ)G_{d}(\theta) and Gd(θ)G_{d^{\prime}}(\theta^{\prime}), we use notation Gd(θ)Gd(θ)G_{d}(\theta)\prec G_{d^{\prime}}(\theta^{\prime}) whenever this relation holds for the sectors involved in the definition of the corresponding sectorial regions.

Given a complex Banach space (𝔼,𝔼)(\mathbb{E},\left\|\cdot\right\|_{\mathbb{E}}), the set 𝒪(U,𝔼)\mathcal{O}(U,\mathbb{E}) stands for the set of holomorphic functions in a set UU\subseteq\mathbb{C}, with values in 𝔼\mathbb{E}. If 𝔼=\mathbb{E}=\mathbb{C}, then we simply write 𝒪(U)\mathcal{O}(U). We denote the formal power series with coefficients in 𝔼\mathbb{E} by 𝔼[[z]]\mathbb{E}[[z]].

Given f^,g^𝔼[[z]]\hat{f},\hat{g}\in\mathbb{E}[[z]], with f^(z)=p0fpzp\hat{f}(z)=\sum_{p\geq 0}f_{p}z^{p} and g^(z)=p0gpzp\hat{g}(z)=\sum_{p\geq 0}g_{p}z^{p}, such that gp0g_{p}\geq 0 for all p0p\geq 0, we write f^(z)g^(z)\hat{f}(z)\ll\hat{g}(z) if |fp|gp|f_{p}|\leq g_{p} for all p0p\geq 0.

3. On generalized summability and moment differentiation

The aim of this section is to recall the concept and main results on the so-called generalized moment differentiation of formal power series. Certain algebraic properties associated with the families of analytic functions which are related to this notion allow to go further by defining the moment differentiation associated with the sum of a given formal power series.

3.1. Strongly regular sequences and related elements

As a first step, we recall the main tools associated with strongly regular sequences and some of their related elements. The concept of a strongly regular sequence is put forward by V. Thilliez in [27].

Definition 1.

Let 𝕄:=(Mp)p0\mathbb{M}:=(M_{p})_{p\geq 0} be a sequence of positive real numbers with M0=1M_{0}=1.

  • (lc)(lc)

    The sequence 𝕄\mathbb{M} is logarithmically convex if

    Mp2Mp1Mp+1, for all p1.M_{p}^{2}\leq M_{p-1}M_{p+1},\hbox{ for all }p\geq 1.
  • (mg)(mg)

    The sequence 𝕄\mathbb{M} is of moderate growth if there exists A1>0A_{1}>0 such that

    Mp+qA1p+qMpMq, for all p,q0.M_{p+q}\leq A_{1}^{p+q}M_{p}M_{q},\hbox{ for all }p,q\geq 0.
  • (snq)(snq)

    The sequence 𝕄\mathbb{M} satisfies the strong non-quasianalyticity condition if there exists A2>0A_{2}>0 such that

    qpMq(q+1)Mq+1A2MpMp+1, for all p0.\sum_{q\geq p}\frac{M_{q}}{(q+1)M_{q+1}}\leq A_{2}\frac{M_{p}}{M_{p+1}},\hbox{ for all }p\geq 0.

Any sequence satisfying properties (lc)(lc), (mg)(mg) and (snq)(snq) is known as a strongly regular sequence.

It is worth recalling that given a (lc) sequence 𝕄=(Mp)p0\mathbb{M}=(M_{p})_{p\geq 0}, one has

(2) MpMqMp+q, for all p,q0M_{p}M_{q}\leq M_{p+q},\hbox{ for all }p,q\in\mathbb{N}_{0}

(see Proposition 2.6 (ii.2) [25]), which entails that given s0s\in\mathbb{N}_{0}, one has

(3) MpsMps, for all p0,M_{p}^{s}\leq M_{ps},\hbox{ for all }p\in\mathbb{N}_{0},

following an induction argument.

Examples of such sequences are of great importance in the study of formal and analytic solutions of differential equations. Gevrey sequences are predominant among them, appearing as upper bounds for growth of the coefficients of the formal solutions of such equations. Given α>0\alpha>0, the Gevrey sequence of order α\alpha is defined by 𝕄α:=(p!α)p0\mathbb{M}_{\alpha}:=(p!^{\alpha})_{p\geq 0}. A natural generalization of the previous are the sequences defined by 𝕄α,β:=(p!αm=0plogβ(e+m))p0\mathbb{M}_{\alpha,\beta}:=(p!^{\alpha}\prod_{m=0}^{p}\log^{\beta}(e+m))_{p\geq 0} for α>0\alpha>0 and β\beta\in\mathbb{R}. These sequences turn out to be strongly regular sequences, provided that the first terms are slightly modified in the case that β<0\beta<0, without affecting their asymptotic behavior. Formal solutions of difference equations are quite related to the 1+ level, associated with the sequence 𝕄1,1\mathbb{M}_{1,-1}, see [7, 8].

Given a strongly regular sequence 𝕄=(Mp)p0\mathbb{M}=(M_{p})_{p\geq 0}, one can define the function

(4) M(t):=supp0log(tpMp),t>0,M(0)=0,M(t):=\sup_{p\geq 0}\log\left(\frac{t^{p}}{M_{p}}\right),\quad t>0,\qquad M(0)=0,

which is non-decreasing, and continuous in [0,)[0,\infty) with limtM(t)=+\lim_{t\to\infty}M(t)=+\infty. J. Sanz ([24], Theorem 3.4) proves that the order of the function M(t)M(t), defined in [4] by

ρ(M):=lim suprmax{0,log(M(r))log(r)}\rho(M):=\limsup_{r\to\infty}\max\left\{0,\frac{\log(M(r))}{\log(r)}\right\}

is a positive real number. Moreover, its inverse ω(𝕄):=1/ρ(M)\omega(\mathbb{M}):=1/\rho(M) determines the limit opening for a sector in such a way that nontrivial flat function in ultraholomorphic classes of functions defined on such sectors exist, see Corollary 3.16, [11]. Indeed, ω(𝕄)\omega(\mathbb{M}) can be recovered directly from 𝕄\mathbb{M} under some admissibility conditions on the sequence 𝕄\mathbb{M} (Corollary 3.10, [10]):

limplog(Mp+1/Mp)log(p)=ω(𝕄).\lim_{p\to\infty}\frac{\log(M_{p+1}/M_{p})}{\log(p)}=\omega(\mathbb{M}).

Such conditions are satisfied by the sequences of general use in the asymptotic theory of solutions to functional equations.

The next results can be found in [3, 27] under more general assumptions.

Lemma 1.

Let 𝕄\mathbb{M} be a strongly regular sequence, and let s1s\geq 1. Then, there exists ρ(s)1\rho(s)\geq 1 which only depends on 𝕄\mathbb{M} and ss, such that

exp(M(t))exp(sM(t/ρ(s))),\exp\left(-M(t)\right)\leq\exp(-sM(t/\rho(s))),

for all t0t\geq 0.

In view of Lemma 1.3.4 [27], one has that given a strongly regular sequence, 𝕄=(Mp)p0\mathbb{M}=(M_{p})_{p\geq 0} and s>0s>0. Then, the sequence 𝕄s=(Mps)p0\mathbb{M}^{s}=(M_{p}^{s})_{p\geq 0} is strongly regular, and ω(𝕄s)=ω(s𝕄)\omega(\mathbb{M}^{s})=\omega(s\mathbb{M}).

Following [22], one has the next definition.

Definition 2.

Given two sequences of positive real numbers, 𝕄=(Mp)p0\mathbb{M}=(M_{p})_{p\geq 0} and 𝕄~:=(M~p)p0\tilde{\mathbb{M}}:=(\tilde{M}_{p})_{p\geq 0}, we say that 𝕄\mathbb{M} and 𝕄~\tilde{\mathbb{M}} are equivalent if there exist B1,B2>0B_{1},B_{2}>0 with

(5) B1pMpM~pB2pMp,B_{1}^{p}M_{p}\leq\tilde{M}_{p}\leq B_{2}^{p}M_{p},

for every p0p\geq 0.

The next result is a direct consequence of the definition of the function MM in (4).

Lemma 2.

Let 𝕄,𝕄~\mathbb{M},\,\tilde{\mathbb{M}} be two strongly regular sequences which are equivalent. Let M(t)M(t) (resp. M~(t)\tilde{M}(t)) be the function associated with 𝕄\mathbb{M} (resp. 𝕄~\tilde{\mathbb{M}}) through (4). Then, it holds that

M(tB2)M~(t)M(tB1) for all t0,M\left(\frac{t}{B_{2}}\right)\leq\tilde{M}(t)\leq M\left(\frac{t}{B_{1}}\right)\hbox{ for all }t\geq 0,

where B1,B2B_{1},\,B_{2} are the positive constants in (5).

Definition 3.

Let (Mp)p0(M_{p})_{p\geq 0} be a sequence of positive real numbers with M0=1M_{0}=1, and let ss\in\mathbb{R}. A sequence of positive real numbers (m(p))p0(m(p))_{p\geq 0} is said to be an (Mp)(M_{p})-sequence of order ss if there exist A3,A4>0A_{3},A_{4}>0 with

(6) A3p(Mp)sm(p)A4p(Mp)s,p0.A_{3}^{p}(M_{p})^{s}\leq m(p)\leq A_{4}^{p}(M_{p})^{s},\quad p\geq 0.

3.2. Function spaces and generalized summability

In the whole subsection (𝔼,𝔼)(\mathbb{E},\left\|\cdot\right\|_{\mathbb{E}}) stands for a complex Banach space.

Definition 4.

Let r,θ>0r,\,\theta>0 and dd\in\mathbb{R}. We also fix a sequence 𝕄\mathbb{M} of positive real numbers. The set 𝒪𝕄(S^d(θ;r),𝔼)\mathcal{O}^{\mathbb{M}}(\hat{S}_{d}(\theta;r),\mathbb{E}) consists of all functions f𝒪(S^d(θ;r),𝔼)f\in\mathcal{O}(\hat{S}_{d}(\theta;r),\mathbb{E}) such that for every 0<θ<θ0<\theta^{\prime}<\theta and 0<r<r0<r^{\prime}<r there exist c~,k~>0\tilde{c},\tilde{k}>0 with

(7) f(z)𝔼c~exp(M(|z|k~)),zS^d(θ,r).\left\|f(z)\right\|_{\mathbb{E}}\leq\tilde{c}\exp\left(M\left(\frac{|z|}{\tilde{k}}\right)\right),\quad z\in\hat{S}_{d}(\theta^{\prime},r^{\prime}).

Analogously, the set 𝒪𝕄(Sd(θ),𝔼)\mathcal{O}^{\mathbb{M}}(S_{d}(\theta),\mathbb{E}) consists of all f𝒪(Sd(θ),𝔼)f\in\mathcal{O}(S_{d}(\theta),\mathbb{E}) such that for all 0<θ<θ0<\theta^{\prime}<\theta, there exist c~,k~>0\tilde{c},\tilde{k}>0 such that (7) holds for all zSd(θ)z\in S_{d}(\theta^{\prime}).

The aforementioned definition generalizes that of functions of exponential growth at infinity of some positive order. Indeed, if 𝕄=𝕄α\mathbb{M}=\mathbb{M}_{\alpha} for some α>0\alpha>0, then the property (7) determines that ff is of exponential growth at most 1/α1/\alpha.

The general moment summability methods developed by W. Balser (see Section 5.5, [1]) were adapted by J. Sanz to the framework of strongly regular sequences (see Section 5, [24]; or Definition 6.2., [25]).

Definition 5.

Let 𝕄\mathbb{M} be a strongly regular sequence with ω(𝕄)<2\omega(\mathbb{M})<2. Let MM be the function associated with 𝕄\mathbb{M}, defined by (4). The complex functions e,Ee,\,E define kernel functions for 𝕄\mathbb{M}-summability if the following properties hold:

  1. (1)

    e𝒪(S0(ω(𝕄)π))e\in\mathcal{O}(S_{0}(\omega(\mathbb{M})\pi)). The function e(z)/ze(z)/z is locally uniformly integrable at the origin, i.e., there exists t0>0t_{0}>0, and for all z0S0(ω(𝕄)π)z_{0}\in S_{0}(\omega(\mathbb{M})\pi) there exists a neighborhood UU of z0z_{0}, with US0(ω(𝕄)π)U\subseteq S_{0}(\omega(\mathbb{M})\pi), such that

    (8) 0t0supzU|e(tz)|t𝑑t<.\int_{0}^{t_{0}}\frac{\sup_{z\in U}\left|e\left(\frac{t}{z}\right)\right|}{t}dt<\infty.

    Moreover, for all ϵ>0\epsilon>0 there exist c,k>0c,k>0 such that

    (9) |e(z)|cexp(M(|z|k)) for all zS0(ω(𝕄)πϵ).|e(z)|\leq c\exp\left(-M\left(\frac{|z|}{k}\right)\right)\quad\hbox{ for all }z\in S_{0}(\omega(\mathbb{M})\pi-\epsilon).

    We also assume that e(x)e(x)\in\mathbb{R} for all x>0x>0.

  2. (2)

    E𝒪()E\in\mathcal{O}(\mathbb{C}) and satisfies that

    (10) |E(z)|c~exp(M(|z|k~)),z,|E(z)|\leq\tilde{c}\exp\left(M\left(\frac{|z|}{\tilde{k}}\right)\right),\quad z\in\mathbb{C},

    for some c~,k~>0\tilde{c},\,\tilde{k}>0. There exists β>0\beta>0 such that for all 0<θ~<2πω(𝕄)π0<\tilde{\theta}<2\pi-\omega(\mathbb{M})\pi and ME>0M_{E}>0, there exist c~2>0\tilde{c}_{2}>0 with

    (11) |E(z)|c~2|z|β,zSπ(θ~)D(0,ME).|E(z)|\leq\frac{\tilde{c}_{2}}{|z|^{\beta}},\quad z\in S_{\pi}(\tilde{\theta})\setminus D(0,M_{E}).
  3. (3)

    Both kernel functions are related via the Mellin transform of ee. More precisely, the moment function associated with ee, defined by

    (12) me(z):=0tz1e(t)𝑑tm_{e}(z):=\int_{0}^{\infty}t^{z-1}e(t)dt

    is a complex continuous function in {z:Re(z)0}\{z\in\mathbb{C}:\hbox{Re}(z)\geq 0\} and holomorphic in {z:Re(z)>0}\{z\in\mathbb{C}:\hbox{Re}(z)>0\}. The kernel function EE has the power series expansion at the origin given by

    (13) E(z)=p0zpme(p),z.E(z)=\sum_{p\geq 0}\frac{z^{p}}{m_{e}(p)},\quad z\in\mathbb{C}.

Remark: In the remainder of the work we will only mention the kernel function ee, rather than the pair e,Ee,\,E, as EE is determined by the knowledge of ee in terms of its Taylor expansion at the origin.

Remark: The growth condition of the kernel function E(z)E(z) at infinity (11) is usually substituted in the literature ([1, 19, 24, 25]) by the less restrictive condition:

“The function E(1/z)/zE(1/z)/z is locally uniformly integrable at the origin in the sector Sπ((2ω(𝕄))π)S_{\pi}((2-\omega(\mathbb{M}))\pi). Namely, there exists t0>0t_{0}>0, and for all z0Sπ((2ω(𝕄))π)z_{0}\in S_{\pi}((2-\omega(\mathbb{M}))\pi) there exists a neighborhood UU of z0z_{0}, USπ((2ω(𝕄))π)U\subseteq S_{\pi}((2-\omega(\mathbb{M}))\pi), such that

0t0supzU|E(zt)|t𝑑t<.\int_{0}^{t_{0}}\frac{\sup_{z\in U}\left|E\left(\frac{z}{t}\right)\right|}{t}dt<\infty.\hbox{''}

Condition (11) has already been used and justified in [9] (see Lemma 4.10, Remark 4.11 and Remark 4.12 in [9]), in order to obtain convolution kernels for multisummability.

Definition 6.

Let 𝕄\mathbb{M} be a strongly regular sequence and let e,Ee,\,E be a pair of kernel functions for 𝕄\mathbb{M}-summability. Let mem_{e} be the moment function given by (12). The sequence (me(p))p0(m_{e}(p))_{p\geq 0} is the so-called sequence of moments associated with ee.

Remark: The previous definition can be adapted to the case ω(𝕄)2\omega(\mathbb{M})\geq 2 by means of a ramification of the kernels (see [25], Remark 6.3 (iii)). For practical purposes, we will focus on the case that ω(𝕄)<2\omega(\mathbb{M})<2, taking into consideration that all the results can be adapted to the general case.

Remark: Given a strongly regular sequence 𝕄\mathbb{M}, the existence of pairs of kernel functions for 𝕄\mathbb{M}-summability is guaranteed, provided that 𝕄\mathbb{M} admits a nonzero proximate order (see [10, 13]).

Example 1.

Let α>0\alpha>0. We consider a Gevrey sequence 𝕄α\mathbb{M}_{\alpha}. Then, the functions eα(z):=1αz1αexp(z1α)e_{\alpha}(z):=\frac{1}{\alpha}z^{\frac{1}{\alpha}}\exp\left(-z^{\frac{1}{\alpha}}\right) and Eα(z):=p0zpΓ(1+αp)E_{\alpha}(z):=\sum_{p\geq 0}\frac{z^{p}}{\Gamma(1+\alpha p)} are kernel functions for 𝕄α\mathbb{M}_{\alpha}-summability. Indeed, the moment function is given by mα(z):=Γ(1+αz)m_{\alpha}(z):=\Gamma(1+\alpha z).

The definition of moment differentiation, moment (formal) Borel and moment Laplace transformation generalize the classical concepts of differentiation, formal Borel and Laplace transformations, respectively. In the classical setting of the Gevrey sequence of order α>0\alpha>0, the moment sequence is (Γ(1+αp))p0(\Gamma(1+\alpha p))_{p\geq 0} seen in Example 1. Classical differentiation corresponds to α=1\alpha=1.

At this point, given a strongly regular sequence 𝕄\mathbb{M}, one has that a sequence of moments can be constructed, provided a pair of kernel functions for 𝕄\mathbb{M}-summability, say ee and EE. The associated sequence of moments me:=(me(p))p0m_{e}:=(m_{e}(p))_{p\geq 0} is a strongly regular sequence (see [25], Remark 6.6), which is equivalent to 𝕄\mathbb{M} (see [25], Proposition 6.5). Therefore, ω(𝕄)=ω(me)\omega(\mathbb{M})=\omega(m_{e}). The definition of generalized derivatives is done in terms of a sequence of moments, rather than the initial sequence itself, and we will work directly with this sequence, obviating the initial strongly regular sequence and the pair of kernel functions. Hereinafter, when referring to a sequence of moments, we will assume without mentioning that such sequence is indeed the sequence of moments associated with some pair of kernel functions, and therefore with a strongly regular sequence (in conditions that admit such pair of kernels, e.g., if the strongly regular sequence admits a nonzero proximate order).

Departing from a sequence of moments mem_{e}, one can consider the formal mem_{e}-moment Borel transform. This definition can be extended for any sequence of positive numbers, and not only to a sequence of moments. We present it in this way for the sake of clarity.

Definition 7.

Let me=(me(p))p0m_{e}=(m_{e}(p))_{p\geq 0} be a sequence of moments. The formal mem_{e}-moment Borel transform ^me,z:𝔼[[z]]𝔼[[z]]\hat{\mathcal{B}}_{m_{e},z}:\mathbb{E}[[z]]\to\mathbb{E}[[z]] is defined by

^me,z(p0apzp)=p0apme(p)zp.\hat{\mathcal{B}}_{m_{e},z}\left(\sum_{p\geq 0}a_{p}z^{p}\right)=\sum_{p\geq 0}\frac{a_{p}}{m_{e}(p)}z^{p}.

There are several different equivalent approaches to the general moment summability of formal power series. We refer to [1], Section 6.5, under Gevrey-like settings, and [25], Section 6, in the framework of strongly regular sequences.

Definition 8.

Let 𝕄\mathbb{M} be a strongly regular sequence admitting a nonzero proximate order. The formal power series u^𝔼[[z]]\hat{u}\in\mathbb{E}[[z]] is 𝕄\mathbb{M}-summable in direction dd\in\mathbb{R} if the formal power series ^𝕄,z(u^(z))\hat{\mathcal{B}}_{\mathbb{M},z}(\hat{u}(z)) is convergent in a neighborhood of the origin and can be extended to an infinite sector of bisecting direction dd, say S^d\hat{S}_{d}, such that the extension belongs to 𝒪𝕄(S^d,𝔼)\mathcal{O}^{\mathbb{M}}(\hat{S}_{d},\mathbb{E}). We write 𝔼{z}𝕄,d\mathbb{E}\{z\}_{\mathbb{M},d} for the set of mem_{e}-summable formal power series in 𝔼[[z]]\mathbb{E}[[z]]. Here we have assumed that ee is a kernel function for 𝕄\mathbb{M}-summability and mem_{e} is its associated sequence of moments.

Remark: We recall that, given a sequence of moments associated with a strongly regular sequence 𝕄\mathbb{M} via a kernel function ee, say mem_{e}, 𝕄\mathbb{M} and mem_{e} are equivalent sequences. Regarding Lemma 2 and the definition of the formal Borel transformation, it is easy to check that the set 𝔼{z}me,d\mathbb{E}\{z\}_{m_{e},d} does not depend on the choice of the kernel function ee, and therefore one can write 𝔼{z}𝕄,d:=𝔼{z}me,d\mathbb{E}\{z\}_{\mathbb{M},d}:=\mathbb{E}\{z\}_{m_{e},d} for any choice of kernel function ee. Moreover, we observe that the formal power series ^me,z(u^)\hat{\mathcal{B}}_{m_{e},z}(\hat{u}) has a positive radius of convergence with independence of the kernel function considered, associated with 𝕄\mathbb{M}.

The statements in the next proposition can be found in detail in [25], Section 6.

Proposition 1.

Let dd\in\mathbb{R} and let e,Ee,\,E be a pair of kernel functions for 𝕄\mathbb{M}-summability. Let θ>0\theta>0. For every f𝒪𝕄(Sd(θ),𝔼)f\in\mathcal{O}^{\mathbb{M}}(S_{d}(\theta),\mathbb{E}), the ee-Laplace transform of ff along a direction τarg(Sd(θ))\tau\in\hbox{arg}(S_{d}(\theta)) is defined by

(Te,τf)(z)=0(τ)e(u/z)f(u)duu,(T_{e,\tau}f)(z)=\int_{0}^{\infty(\tau)}e(u/z)f(u)\frac{du}{u},

for |arg(z)τ|<ω(𝕄)π/2|\hbox{arg}(z)-\tau|<\omega(\mathbb{M})\pi/2, and |z||z| small enough. The variation of τarg(Sd)\tau\in\hbox{arg}(S_{d}) defines a function denoted by Te,dfT_{e,d}f in a sectorial region Gd(θ+ω(𝕄)π)G_{d}(\theta+\omega(\mathbb{M})\pi).

Under the assumption that ω(𝕄)<2\omega(\mathbb{M})<2 let G=Gd(θ)G=G_{d}(\theta) be a sectorial region with θ>ω(𝕄)π\theta>\omega(\mathbb{M})\pi. Given f𝒪(G,𝔼)f\in\mathcal{O}(G,\mathbb{E}) and continuous at 0, and τ\tau\in\mathbb{R} with |τd|<(θω(𝕄)π)/2|\tau-d|<(\theta-\omega(\mathbb{M})\pi)/2, the operator Te,τT^{-}_{e,\tau}, known as the ee-Borel transform along direction τ\tau is defined by

(Te,τf)(u):=12πiδω(𝕄)(τ)E(u/z)f(z)dzz,uSτ,(T^{-}_{e,\tau}f)(u):=\frac{-1}{2\pi i}\int_{\delta_{\omega(\mathbb{M})}(\tau)}E(u/z)f(z)\frac{dz}{z},\quad u\in S_{\tau},

where SτS_{\tau} is an infinite sector of bisecting direction τ\tau and small enough opening, and δω(𝕄)(τ)\delta_{\omega(\mathbb{M})}(\tau) is the Borel-like path consisting of the concatenation of a segment from the origin to a point z0z_{0} with arg(z0)=τ+ω(𝕄)(π+ϵ)/2\hbox{arg}(z_{0})=\tau+\omega(\mathbb{M})(\pi+\epsilon)/2, for some small enough ϵ(0,π)\epsilon\in(0,\pi), followed with the arc of circle centered at 0, joining z0z_{0} and the point z1z_{1}, with arg(z1)=τω(𝕄)(π+ϵ)/2\hbox{arg}(z_{1})=\tau-\omega(\mathbb{M})(\pi+\epsilon)/2, clockwise, and concluding with the segment of endpoints z1z_{1} and the origin.

Let Gd(θ)G_{d}(\theta) and ff be as above. The family {Te,τ}τ\{T^{-}_{e,\tau}\}_{\tau}, with τ\tau varying among the real numbers with |τd|<(θω(𝕄)π)/2|\tau-d|<(\theta-\omega(\mathbb{M})\pi)/2 defines a holomorphic function denoted by Te,dfT^{-}_{e,d}f in the sector Sd(θω(𝕄)π)S_{d}(\theta-\omega(\mathbb{M})\pi) and Te,df𝒪𝕄(Sd(θω(𝕄)π),𝔼)T^{-}_{e,d}f\in\mathcal{O}^{\mathbb{M}}(S_{d}(\theta-\omega(\mathbb{M})\pi),\mathbb{E}).

Remark: We recall that if λ\lambda\in\mathbb{C}, with Re(λ)0\hbox{Re}(\lambda)\geq 0, then Te,d(uuλ)(z)=zλme(λ)T^{-}_{e,d}(u\mapsto u^{\lambda})(z)=\frac{z^{\lambda}}{m_{e}(\lambda)}, which relates Te,dT^{-}_{e,d} with the formal Borel operator ^me,u\hat{\mathcal{B}}_{m_{e},u}.

Theorem 30 [1] can be adapted to the strongly regular sequence framework, under minor modifications.

Theorem 1.

Let S=Sd(θ)S=S_{d}(\theta) for some θ>0\theta>0. Let 𝕄\mathbb{M} be a strongly regular sequence with ω(𝕄)<2\omega(\mathbb{M})<2 admitting a nonzero proximate order, and choose a kernel function for 𝕄\mathbb{M}-summability ee. Let f𝒪𝕄(S,𝔼)f\in\mathcal{O}^{\mathbb{M}}(S,\mathbb{E}) and define g(z)=(Te,df)(z)g(z)=(T_{e,d}f)(z) for zz in a sectorial region Gd(θ+ω(𝕄)π))G_{d}(\theta+\omega(\mathbb{M})\pi)). Then it holds that fTe,dgf\equiv T^{-}_{e,d}g.

The following is an equivalent of 𝕄\mathbb{M}-summable formal power series (see Theorem 6.18, [25]).

Theorem 2.

Let 𝕄=(Mp)p0\mathbb{M}=(M_{p})_{p\geq 0} be a strongly regular sequence admitting a nonzero proximate order. Let u^=p0upzp𝔼[[z]]\hat{u}=\sum_{p\geq 0}u_{p}z^{p}\in\mathbb{E}[[z]] and dd\in\mathbb{R}. The following statements are equivalent:

  • (a)

    u^(z)\hat{u}(z) is 𝕄\mathbb{M}-summable in direction dd.

  • (b)

    There exists a sectorial region Gd(θ)G_{d}(\theta) with θ>ω(𝕄)π\theta>\omega(\mathbb{M})\pi and u𝒪(Gd(θ),𝔼)u\in\mathcal{O}(G_{d}(\theta),\mathbb{E}) such that for all 0<θ<θ0<\theta^{\prime}<\theta, Sd(θ;r)Gd(θ)S_{d}(\theta^{\prime};r)\subseteq G_{d}(\theta) and all integers N1N\geq 1, there exist C,A>0C,\,A>0 with

    u(z)p=0N1upzp𝔼CANMN|z|N,zSd(θ;r).\left\|u(z)-\sum_{p=0}^{N-1}u_{p}z^{p}\right\|_{\mathbb{E}}\leq CA^{N}M_{N}|z|^{N},\qquad z\in S_{d}(\theta^{\prime};r).

If one of the previous equivalent statements holds, the function uu in Definition 8 can be constructed as the ee-Laplace transform of ^𝕄,z(u^(z))\hat{\mathcal{B}}_{\mathbb{M},z}(\hat{u}(z)) along direction τarg(Sd)\tau\in\hbox{arg}(S_{d}).

The previous construction can be done for, and it is independent of, any choice of the kernel for 𝕄\mathbb{M}-summability ee.

The function uu satisfying the previous equivalent properties is unique (see Corollary 4.30 [25]), and it is known as the 𝕄\mathbb{M}-sum of u^\hat{u} along direction dd. We write 𝒮𝕄,d(u^)\mathcal{S}_{\mathbb{M},d}(\hat{u}) for the 𝕄\mathbb{M}-sum of u^\hat{u} along direction dd.

The concept of a moment differential operator was put forward by W. Balser and M. Yoshino in [2], and extended to mem_{e}-moment differential operators in [15], which leans on moment sequences of some positive order.

Definition 9.

Let (𝔼,𝔼)(\mathbb{E},\left\|\cdot\right\|_{\mathbb{E}}) be a complex Banach space. Given a sequence of moments (me(p))p0(m_{e}(p))_{p\geq 0}, the mem_{e}-moment differentiation me,z\partial_{m_{e},z} is the linear operator me,z:𝔼[[z]]𝔼[[z]]\partial_{m_{e},z}:\mathbb{E}[[z]]\to\mathbb{E}[[z]] defined by

me,z(p0upme(p)zp):=p0up+1me(p)zp.\partial_{m_{e},z}\left(\sum_{p\geq 0}\frac{u_{p}}{m_{e}(p)}z^{p}\right):=\sum_{p\geq 0}\frac{u_{p+1}}{m_{e}(p)}z^{p}.

This definition can be naturally extended to f𝒪(D,𝔼)f\in\mathcal{O}(D,\mathbb{E}), for some complex Banach space (𝔼,𝔼)(\mathbb{E},\left\|\cdot\right\|_{\mathbb{E}}), and any neighborhood of the origin DD, by applying the definition of me,z\partial_{m_{e},z} to the Taylor expansion of ff at the origin. Moreover, one defines the linear operator me,z1:𝔼[[z]]𝔼[[z]]\partial_{m_{e},z}^{-1}:\mathbb{E}[[z]]\to\mathbb{E}[[z]] as the inverse operator of me,z\partial_{m_{e},z}, i.e. me,z1(zj)=me(j)me(j+1)zj+1\partial_{m_{e},z}^{-1}(z^{j})=\frac{m_{e}(j)}{m_{e}(j+1)}z^{j+1} for every j0j\geq 0.

Lemma 3.

Let m1=(m1(p))p0,m2=(m2(p))p0m_{1}=(m_{1}(p))_{p\geq 0},\,m_{2}=(m_{2}(p))_{p\geq 0} be two sequences of moments. The following statements hold:

  • The sequence of products m1m2:=(m1(p)m2(p))p0m_{1}m_{2}:=(m_{1}(p)m_{2}(p))_{p\geq 0} is a sequence of moments.

  • ^m1,zm2,zm1m2,z^m1,z\hat{\mathcal{B}}_{m_{1},z}\circ\partial_{m_{2},z}\equiv\partial_{m_{1}m_{2},z}\circ\hat{\mathcal{B}}_{m_{1},z} as operators defined in 𝔼[[z]]\mathbb{E}[[z]].

Proof.

The first part is a direct consequence of Proposition 4.15, [9]. The second part is a direct consequence of the definition of the formal Borel transform (Definition 7) and the moment differentiation (Definition 9). ∎

The first statement of the next result extends Proposition 6 [19] to the framework of strongly regular sequences. Its proof rests heavily on that of Proposition 3, [17]. The second statement will be crucial in the sequel at the time of giving a coherent meaning to the moment derivatives of the sum of a formal power series.

Theorem 3.

Let me=(me(p))p0m_{e}=(m_{e}(p))_{p\geq 0} be a sequence of moments, and let 𝕄~\tilde{\mathbb{M}} be a strongly regular sequence. We also fix d,θ,rd,\,\theta,\,r\in\mathbb{R}, with θ,r>0\theta,\,r>0, and φ𝒪𝕄~(S^d(θ;r),𝔼)\varphi\in\mathcal{O}^{\tilde{\mathbb{M}}}(\hat{S}_{d}(\theta;r),\mathbb{E}). Then there exists 0<r~<r0<\tilde{r}<r such that for all 0<θ1<θ0<\theta_{1}<\theta, all zS^d(θ1;r~)z\in\hat{S}_{d}(\theta_{1};\tilde{r}) and n0n\in\mathbb{N}_{0}, the following statements hold:

  1. (1)
    (14) me,znφ(z)=12πiΓzφ(w)0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ𝑑w,\partial_{m_{e},z}^{n}\varphi(z)=\frac{1}{2\pi i}\oint_{\Gamma_{z}}\varphi(w)\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi dw,

    with τ=τ(ω)(arg(w)ω(me)π2,arg(w)+ω(me)π2)\tau=\tau(\omega)\in(-\arg(w)-\frac{\omega(m_{e})\pi}{2},-\arg(w)+\frac{\omega(m_{e})\pi}{2}), which depends on ww. The integration path Γz\Gamma_{z} is a deformation of the circle {|w|=r1}\{|w|=r_{1}\}, for any choice of 0<r1<r0<r_{1}<r, which depends on zz.

  2. (2)

    There exist constants C1,C2,C3>0C_{1},C_{2},C_{3}>0 such that

    (15) me,znφ(z)𝔼C1C2nme(n)exp(M~(C3|z|)),\left\|\partial_{m_{e},z}^{n}\varphi(z)\right\|_{\mathbb{E}}\leq C_{1}C_{2}^{n}m_{e}(n)\exp\left(\tilde{M}(C_{3}|z|)\right),

    for all n0n\in\mathbb{N}_{0} and zS^d(θ1;r~)z\in\hat{S}_{d}(\theta_{1};\tilde{r}).

Proof.

We first give a proof for the first statement. From Taylor expansion of φ\varphi at the origin and the definition of mem_{e}-moment derivatives one has

φ(z)=me(0)p0me,zpφ(0)me(p)zp,\varphi(z)=m_{e}(0)\sum_{p\geq 0}\frac{\partial_{m_{e},z}^{p}\varphi(0)}{m_{e}(p)}z^{p},

for all zD(0,r)z\in D(0,r). The application of the Cauchy integral formula for the derivatives yields

me,zpφ(0)=me(p)p!me(0)φ(p)(0)=me(p)2πime(0)|w|=r1φ(w)wp+1𝑑w,\partial_{m_{e},z}^{p}\varphi(0)=\frac{m_{e}(p)}{p!m_{e}(0)}\varphi^{(p)}(0)=\frac{m_{e}(p)}{2\pi im_{e}(0)}\oint_{|w|=r_{1}}\frac{\varphi(w)}{w^{p+1}}dw,

for any 0<r1<r0<r_{1}<r. Let ww\in\mathbb{C} with |w|=r1|w|=r_{1}. Following (12) and the change of variables x=ξwx=\xi w we derive

me(p)wp+1=0xp1e(x)wp+1𝑑x=0(τ)ξpe(ξw)ξw𝑑ξ,\frac{m_{e}(p)}{w^{p+1}}=\int_{0}^{\infty}x^{p-1}\frac{e(x)}{w^{p+1}}dx=\int_{0}^{\infty(\tau)}\xi^{p}\frac{e(\xi w)}{\xi w}d\xi,

where τ=arg(w)\tau=-\arg(w). We observe from (9) that the previous equality can be extended to any direction of integration τ(arg(w)ω(me)π2,arg(w)+ω(me)π2)\tau\in\left(-\arg(w)-\frac{\omega(m_{e})\pi}{2},-\arg(w)+\frac{\omega(m_{e})\pi}{2}\right). Therefore, regarding the definition of the kernel function E(z)E(z) in (13), one has

(16) φ(z)=me(0)p0me,zpφ(0)me(p)zp=12πi|w|=r1φ(w)0(τ)e(ξw)ξwp0ξpzpme(p)dξdw=12πi|w|=r1φ(w)0(τ)E(ξz)e(ξw)ξw𝑑ξ𝑑w.\varphi(z)=m_{e}(0)\sum_{p\geq 0}\frac{\partial_{m_{e},z}^{p}\varphi(0)}{m_{e}(p)}z^{p}=\frac{1}{2\pi i}\oint_{|w|=r_{1}}\varphi(w)\int_{0}^{\infty(\tau)}\frac{e(\xi w)}{\xi w}\sum_{p\geq 0}\frac{\xi^{p}z^{p}}{m_{e}(p)}d\xi dw\\ =\frac{1}{2\pi i}\oint_{|w|=r_{1}}\varphi(w)\int_{0}^{\infty(\tau)}E(\xi z)\frac{e(\xi w)}{\xi w}d\xi dw.

We conclude the first part of the proof, at least from the formal point of view, by observing that

me,znE(ξz)=me,zn((ξz)pme(p))=p0ξp+nzpme(p)=ξnE(ξz),\partial_{m_{e},z}^{n}E(\xi z)=\partial_{m_{e},z}^{n}\left(\frac{(\xi z)^{p}}{m_{e}(p)}\right)=\sum_{p\geq 0}\frac{\xi^{p+n}z^{p}}{m_{e}(p)}=\xi^{n}E(\xi z),

for every ξ,z\xi,z\in\mathbb{C}. It only rests to guarantee that the formal interchange of sum and integrals in (16) can also be made with analytic meaning. We give details about this issue in the second part of the proof.

We proceed to give proof for the second statement of the result. Let us first consider the integral

0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ,\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi,

for zz belonging to some neighborhood of the origin, ww\in\mathbb{C} with |w|=r1|w|=r_{1} as above. We choose τ(arg(w)ω(me)π2,arg(w)+ω(me)π2)\tau\in\left(-\arg(w)-\frac{\omega(m_{e})\pi}{2},-\arg(w)+\frac{\omega(m_{e})\pi}{2}\right). We first prove that

(17) |0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ|A0B0nme(n),\left|\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi\right|\leq A_{0}B_{0}^{n}m_{e}(n),

for some A0,B0>0A_{0},\,B_{0}>0 and all n0n\geq 0. This can be done following analogous arguments as those in the proof of Lemma 7.2, [25]. Let us consider the parametrization [0,)sseiτ[0,\infty)\ni s\mapsto se^{i\tau}. In view of (9) and (10), we have

(18) |0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ|0sn|E(seiτz)||e(seiτw)|sr1𝑑scc~r10sn1exp(M(s|z|k~))exp(M(sr1k))𝑑s,\left|\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi\right|\leq\int_{0}^{\infty}s^{n}|E(se^{i\tau}z)|\frac{|e(se^{i\tau}w)|}{sr_{1}}ds\\ \leq\frac{c\tilde{c}}{r_{1}}\int_{0}^{\infty}s^{n-1}\exp\left(M\left(\frac{s|z|}{\tilde{k}}\right)\right)\exp\left(-M\left(\frac{sr_{1}}{k}\right)\right)ds,

for some c,c~,k,k~>0c,\,\tilde{c},\,k,\,\tilde{k}>0. We apply Lemma 1 to arrive at

(19) exp(M(s|z|k~))exp(M(sr1k))exp(M(s|z|k~)2M(sr1ρ(2)k)).\exp\left(M\left(\frac{s|z|}{\tilde{k}}\right)\right)\exp\left(-M\left(\frac{sr_{1}}{k}\right)\right)\leq\exp\left(M\left(\frac{s|z|}{\tilde{k}}\right)-2M\left(\frac{sr_{1}}{\rho(2)k}\right)\right).

We recall that the function MM is a monotone increasing function. Therefore, if |z|r~:=r1k~kρ(2)|z|\leq\tilde{r}:=\frac{r_{1}\tilde{k}}{k\rho(2)}, then (19) is bounded from above by exp(M(sr1/(ρ(2)k)))\exp(-M(sr_{1}/(\rho(2)k))). Let us write

(20) 0sn1eM(sr1ρ(2)k)𝑑s=01sn1eM(sr1ρ(2)k)𝑑s+1sn1eM(sr1ρ(2)k)𝑑s=I1+I2.\int_{0}^{\infty}s^{n-1}e^{-M\left(\frac{sr_{1}}{\rho(2)k}\right)}ds=\int_{0}^{1}s^{n-1}e^{-M\left(\frac{sr_{1}}{\rho(2)k}\right)}ds+\int_{1}^{\infty}s^{n-1}e^{-M\left(\frac{sr_{1}}{\rho(2)k}\right)}ds=I_{1}+I_{2}.

The definition of MM guarantees upper bounds for |I1||I_{1}| which do not depend on nn. We proceed to study I2I_{2}. Bearing in mind the definition of MM, we arrive at

1sn1eM(sr1ρ(2)k)𝑑s(ρ(2)kr1)n+2me(n+2)11s3𝑑s.\int_{1}^{\infty}s^{n-1}e^{-M\left(\frac{sr_{1}}{\rho(2)k}\right)}ds\leq\left(\frac{\rho(2)k}{r_{1}}\right)^{n+2}m_{e}(n+2)\int_{1}^{\infty}\frac{1}{s^{3}}ds.

The application of (mg)(mg) condition on me(n+2)A1n+2me(2)me(n)m_{e}(n+2)\leq A_{1}^{n+2}m_{e}(2)m_{e}(n) allows to conclude (17) for zD(0,r~)z\in D(0,\tilde{r}). The estimate (15) is attained by applying (17) to (14). More precisely, we have

(21) me,znφ(z)𝔼(sup|w|=r1φ(w))A0B0nme(n),\left\|\partial_{m_{e},z}^{n}\varphi(z)\right\|_{\mathbb{E}}\leq\left(\sup_{|w|=r_{1}}\left\|\varphi(w)\right\|\right)A_{0}B_{0}^{n}m_{e}(n),

which entails (15) for zD(0,r~)z\in D(0,\tilde{r}).

Let 0<θ1<θ0<\theta_{1}<\theta, and zSd(θ1)z\in S_{d}(\theta_{1}) with |z|r~|z|\geq\tilde{r}. We study (15) in such a domain. We deform the integration path {|w|=r1}\{|w|=r_{1}\} as follows. Let θ1<θ2<θ\theta_{1}<\theta_{2}<\theta and let R=R(z)=ρ(2)kk~|z|R=R(z)=\frac{\rho(2)k}{\tilde{k}}|z|. We write Γz=Γ(R)=Γ1+Γ2(R)+Γ3(R)+Γ4(R)\Gamma_{z}=\Gamma(R)=\Gamma_{1}+\Gamma_{2}(R)+\Gamma_{3}(R)+\Gamma_{4}(R), where Γ1\Gamma_{1} is the arc of the circle joining the points r1ei(d+θ2)r_{1}e^{i(d+\theta_{2})} and r1ei(dθ2)r_{1}e^{i(d-\theta_{2})} counterclockwise, Γ2(R)\Gamma_{2}(R) is the segment [r1,R]ei(dθ2)[r_{1},R]e^{i(d-\theta_{2})}, Γ3(R)\Gamma_{3}(R) is the arc of the circle joining the points Rei(dθ2/2)Re^{i(d-\theta_{2}/2)} and Rei(d+θ2/2)Re^{i(d+\theta_{2}/2)} counterclockwise and Γ4(R)\Gamma_{4}(R) is the segment [r1,R]ei(d+θ2/2)[r_{1},R]e^{i(d+\theta_{2}/2)}. Figure 1 illustrates this deformation path.

Refer to caption
Figure 1. Deformation path

We first study the case ωΓ1\omega\in\Gamma_{1}, i.e., |w|=r1|w|=r_{1}. Observe that for every wΓ1w\in\Gamma_{1} it is always possible to choose τ\tau such that

(22) τ(arg(w)ω(me)π2,arg(w)+ω(me)π2)(arg(z)+ω(me)π2,arg(z)+2πω(me)π2).\tau\in\left(-\hbox{arg}(w)-\frac{\omega(m_{e})\pi}{2},-\hbox{arg}(w)+\frac{\omega(m_{e})\pi}{2}\right)\cap\left(-\hbox{arg}(z)+\frac{\omega(m_{e})\pi}{2},-\hbox{arg}(z)+2\pi-\frac{\omega(m_{e})\pi}{2}\right).

For one of such directions τ\tau we have

0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ=0(seiτ)nE(zseiτ)e(wseiτ)wseiτeiτ𝑑s.\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi=\int_{0}^{\infty}(se^{i\tau})^{n}E(zse^{i\tau})\frac{e(wse^{i\tau})}{wse^{i\tau}}e^{i\tau}ds.

We split the previous integral into two parts. Let ME>0M_{E}>0. On the one hand, one has

(23) |0ME/|z|(seiτ)nE(zseiτ)e(wseiτ)wseiτeiτ𝑑s|0ME/|z|sn|E(zseiτ)||e(wseiτ)|r1s𝑑s(maxyD¯(0,ME)|E(y)|)1r10ME/|z|sn1|e(wseiτ)|𝑑s.\left|\int_{0}^{M_{E}/|z|}(se^{i\tau})^{n}E(zse^{i\tau})\frac{e(wse^{i\tau})}{wse^{i\tau}}e^{i\tau}ds\right|\leq\int_{0}^{M_{E}/|z|}s^{n}|E(zse^{i\tau})|\frac{|e(wse^{i\tau})|}{r_{1}s}ds\\ \leq\left(\max_{y\in\overline{D}(0,M_{E})}|E(y)|\right)\frac{1}{r_{1}}\int_{0}^{M_{E}/|z|}s^{n-1}|e(wse^{i\tau})|ds.

Bearing in mind (9), we have

0ME/|z|sn1|e(wseiτ)|𝑑sc0ME/|z|sn1exp(M(r1sk))𝑑s.\int_{0}^{M_{E}/|z|}s^{n-1}|e(wse^{i\tau})|ds\leq c\int_{0}^{M_{E}/|z|}s^{n-1}\exp\left(-M\left(\frac{r_{1}s}{k}\right)\right)ds.

Analogous estimates as in (20) allow us to arrive at

(24) |0ME/|z|(seiτ)nE(zseiτ)e(wseiτ)wseiτeiτ𝑑s|C1.1C2.1nme(n)(maxyD¯(0,ME)|E(y)|)1r1n+3,\left|\int_{0}^{M_{E}/|z|}(se^{i\tau})^{n}E(zse^{i\tau})\frac{e(wse^{i\tau})}{wse^{i\tau}}e^{i\tau}ds\right|\leq C_{1.1}C_{2.1}^{n}m_{e}(n)\left(\max_{y\in\overline{D}(0,M_{E})}|E(y)|\right)\frac{1}{r_{1}^{n+3}},

for some C1.1,C2.1>0C_{1.1},\,C_{2.1}>0. Analogously, we estimate the second integral by means of the upper bounds in (11). Indeed, one has

|ME/|z|(seiτ)nE(zseiτ)e(wseiτ)wseiτeiτ𝑑s|\displaystyle\left|\int_{M_{E}/|z|}^{\infty}(se^{i\tau})^{n}E(zse^{i\tau})\frac{e(wse^{i\tau})}{wse^{i\tau}}e^{i\tau}ds\right| ME/|z|snc2~(|z|s)β|e(wseiτ)|r1s𝑑s\displaystyle\leq\int_{M_{E}/|z|}^{\infty}s^{n}\frac{\tilde{c_{2}}}{(|z|s)^{\beta}}\frac{|e(wse^{i\tau})|}{r_{1}s}ds
c2~r1(ME)βME/|z|sn1|e(wseiτ)|𝑑s,\displaystyle\leq\frac{\tilde{c_{2}}}{r_{1}(M_{E})^{\beta}}\int_{M_{E}/|z|}^{\infty}s^{n-1}|e(wse^{i\tau})|ds,

for some c~2,β>0\tilde{c}_{2},\,\beta>0. The estimates in (20) can be applied again in order to arrive at

(25) |ME/|z|(seiτ)nE(zseiτ)e(wseiτ)wseiτeiτ𝑑s|C1.2C2.2nme(n)1r1n+3,\left|\int_{M_{E}/|z|}^{\infty}(se^{i\tau})^{n}E(zse^{i\tau})\frac{e(wse^{i\tau})}{wse^{i\tau}}e^{i\tau}ds\right|\leq C_{1.2}C_{2.2}^{n}m_{e}(n)\frac{1}{r_{1}^{n+3}},

for some C1.2,C2.2>0C_{1.2},\,C_{2.2}>0. From (24) and (25) one can conclude in the spirit of (21) that

(26) 12πiΓ1φ(w)0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ𝑑w𝔼D1D2nme(n)exp(M~(D3|z|)),\left\|\frac{1}{2\pi i}\int_{\Gamma_{1}}\varphi(w)\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi dw\right\|_{\mathbb{E}}\leq D_{1}D_{2}^{n}m_{e}(n)\exp\left(\tilde{M}(D_{3}|z|)\right),

for some D1,D2,D3>0D_{1},\,D_{2},\,D_{3}>0. We continue with the case wΓ2(R)w\in\Gamma_{2}(R). The same choice for τ\tau in (22) holds. We parametrize Γ2(R)\Gamma_{2}(R) by [r1,R]ρρei(dθ2/2)[r_{1},R]\ni\rho\mapsto\rho e^{i(d-\theta_{2}/2)} to arrive at

|0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ|1ρ0sn1|E(zseiτ)||e(ρsei(τ+dθ2/2))|𝑑s.\left|\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi\right|\leq\frac{1}{\rho}\int_{0}^{\infty}s^{n-1}\left|E(zse^{i\tau})\right|\left|e(\rho se^{i\left(\tau+d-\theta_{2}/2\right)})\right|ds.

The same splitting of the path into the segment [0,ME/|z|][0,M_{E}/|z|] and the ray [ME/|z|,)[M_{E}/|z|,\infty), and analogous arguments as in the first part of the proof yield

|0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ|C1.3C2.3nme(n)1ρn+3,\left|\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi\right|\leq C_{1.3}C_{2.3}^{n}m_{e}(n)\frac{1}{\rho^{n+3}},

for some C1.3,C2.3>0C_{1.3},\,C_{2.3}>0, and w=ρei(dθ2/2)w=\rho e^{i(d-\theta_{2}/2)} for some r1ρRr_{1}\leq\rho\leq R. We derive that

(27) Γ2(R)φ(w)0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ𝑑w𝔼C1.3C2.3nme(n)12πr1Rφ(ρei(dθ2/2))𝔼1ρn+3𝑑ρcφC1.32πr13(C2.3r1)nme(n)Rexp(M~(Rk~φ))\left\|\int_{\Gamma_{2}(R)}\varphi(w)\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi dw\right\|_{\mathbb{E}}\leq C_{1.3}C_{2.3}^{n}m_{e}(n)\frac{1}{2\pi}\int_{r_{1}}^{R}\left\|\varphi(\rho e^{i(d-\theta_{2}/2)})\right\|_{\mathbb{E}}\frac{1}{\rho^{n+3}}d\rho\\ \leq\frac{c_{\varphi}C_{1.3}}{2\pi r_{1}^{3}}\left(\frac{C_{2.3}}{r_{1}}\right)^{n}m_{e}(n)R\exp\left(\tilde{M}\left(\frac{R}{\tilde{k}_{\varphi}}\right)\right)

for some cφ,k~φ>0c_{\varphi},\,\tilde{k}_{\varphi}>0 associated with the growth of φ\varphi near infinity. A direct consequence of the definition of the function M~\tilde{M}, and the definition of the radius R=R(|z|)R=R(|z|) yield

Rexp(M~(Rk~φ))exp(M~(cRk~φ))=exp(M~(cρ(2)k|z|k~k~φ)),R\exp\left(\tilde{M}\left(\frac{R}{\tilde{k}_{\varphi}}\right)\right)\leq\exp\left(\tilde{M}\left(\frac{cR}{\tilde{k}_{\varphi}}\right)\right)=\exp\left(\tilde{M}\left(\frac{c\rho(2)k|z|}{\tilde{k}\tilde{k}_{\varphi}}\right)\right),

which allows to end this part of the proof. We get that

(28) 12πiΓ2(R)φ(w)0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ𝑑w𝔼D4D5nme(n)exp(M~(D6|z|)),\left\|\frac{1}{2\pi i}\int_{\Gamma_{2}(R)}\varphi(w)\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi dw\right\|_{\mathbb{E}}\leq D_{4}D_{5}^{n}m_{e}(n)\exp\left(\tilde{M}(D_{6}|z|)\right),

for some D4,D5,D6>0D_{4},\,D_{5},\,D_{6}>0.

The case wΓ4(R)w\in\Gamma_{4}(R) can be treated analogously, to arrive at

(29) 12πiΓ4(R)φ(w)0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ𝑑w𝔼D7D8nme(n)exp(M~(D9|z|)),\left\|\frac{1}{2\pi i}\int_{\Gamma_{4}(R)}\varphi(w)\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi dw\right\|_{\mathbb{E}}\leq D_{7}D_{8}^{n}m_{e}(n)\exp\left(\tilde{M}(D_{9}|z|)\right),

for some D7,D8,D9>0D_{7},\,D_{8},\,D_{9}>0.

We conclude the proof with the case that wΓ3(R)w\in\Gamma_{3}(R). We parametrize Γ3(R)\Gamma_{3}(R) by [dθ2/2,d+θ2/2]tReit[d-\theta_{2}/2,d+\theta_{2}/2]\ni t\mapsto Re^{it} and choose wΓ3(R)w\in\Gamma_{3}(R). Let τ(arg(w)ω(me)π2,arg(w)+ω(me)π2)\tau\in\left(-\arg(w)-\frac{\omega(m_{e})\pi}{2},-\arg(w)+\frac{\omega(m_{e})\pi}{2}\right). Then, one has

|0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ|\displaystyle\left|\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi\right| =|0(seiτ)nE(zseiτ)e(wseiτ)wseiτeiτ𝑑s|\displaystyle=\left|\int_{0}^{\infty}(se^{i\tau})^{n}E(zse^{i\tau})\frac{e(wse^{i\tau})}{wse^{i\tau}}e^{i\tau}ds\right|
1R0sn1|E(zseiτ)||e(wseiτ)|𝑑s.\displaystyle\leq\frac{1}{R}\int_{0}^{\infty}s^{n-1}|E(zse^{i\tau})||e(wse^{i\tau})|ds.

In view of (10) and (9), together with the application of the same argument as in (19) (for r1r_{1} substituted by RR), the previous expression is bounded from above by

(30) cc~R0sn1exp(M(|z|sk~))exp(M(Rsk))𝑑scc~R0sn1exp(M(s|z|k~)2M(sRρ(2)k))𝑑s.\frac{c\tilde{c}}{R}\int_{0}^{\infty}s^{n-1}\exp\left(M\left(\frac{|z|s}{\tilde{k}}\right)\right)\exp\left(-M\left(\frac{Rs}{k}\right)\right)ds\\ \leq\frac{c\tilde{c}}{R}\int_{0}^{\infty}s^{n-1}\exp\left(M\left(\frac{s|z|}{\tilde{k}}\right)-2M\left(\frac{sR}{\rho(2)k}\right)\right)ds.

The function MM is monotone increasing in [0,)[0,\infty). We recall that R=ρ(2)kk~|z|R=\frac{\rho(2)k}{\tilde{k}}|z|, so the previous expression is bounded from above by

cc~R0sn1exp(M(sRρ(2)k))𝑑s.\frac{c\tilde{c}}{R}\int_{0}^{\infty}s^{n-1}\exp\left(-M\left(\frac{sR}{\rho(2)k}\right)\right)ds.

At this point, one can take into account (20), together with R=ρ(2)kk~|z|ρ(2)kk~r~R=\frac{\rho(2)k}{\tilde{k}}|z|\geq\frac{\rho(2)k}{\tilde{k}}\tilde{r}, and

sup|w|=Rφ(w)𝔼cφexp(M~(Rk~φ))=cφexp(M~(ρ(2)kk~k~φ|z|))\sup_{|w|=R}\left\|\varphi(w)\right\|_{\mathbb{E}}\leq c_{\varphi}\exp\left(\tilde{M}\left(\frac{R}{\tilde{k}_{\varphi}}\right)\right)=c_{\varphi}\exp\left(\tilde{M}\left(\frac{\rho(2)k}{\tilde{k}\tilde{k}_{\varphi}}|z|\right)\right)

to get that

(31) 12πiΓ3(R)φ(w)0(τ)ξnE(zξ)e(wξ)wξ𝑑ξ𝑑w𝔼D10D11nme(n)exp(M~(D12|z|)),\left\|\frac{1}{2\pi i}\int_{\Gamma_{3}(R)}\varphi(w)\int_{0}^{\infty(\tau)}\xi^{n}E(z\xi)\frac{e(w\xi)}{w\xi}d\xi dw\right\|_{\mathbb{E}}\leq D_{10}D_{11}^{n}m_{e}(n)\exp\left(\tilde{M}(D_{12}|z|)\right),

for some D10,D11,D12>0D_{10},\,D_{11},\,D_{12}>0.

Statement (15) follows from the application of (26), (28), (29) and (31). We observe that the identity in (14) is of analytic nature after the deformation path with respect to ww and the appropriate choice of τ\tau described in the proof, for each zS^d(θ1;r~)z\in\hat{S}_{d}(\theta_{1};\tilde{r}). ∎

Corollary 1.

Let mem_{e} be a sequence of moments, and let 𝕄~\tilde{\mathbb{M}} be a strongly regular sequence admitting a nonzero proximate order. Given dd\in\mathbb{R}, the space 𝔼{z}𝕄~,d\mathbb{E}\{z\}_{\tilde{\mathbb{M}},d} is closed under mem_{e}-differentiation.

Proof.

Let e~,E~\tilde{e},\,\tilde{E} be a pair of kernel functions for 𝕄~\tilde{\mathbb{M}}-summability, whose existence is guaranteed (see Remark at page 3.2). Let me~m_{\tilde{e}} be its associated sequence of moments. We write m¯:=(me(p)me~(p))p0\overline{m}:=(m_{e}(p)m_{\tilde{e}}(p))_{p\geq 0}, which is a sequence of moments in view of Lemma 3.

Let f^𝔼{z}𝕄~,d\hat{f}\in\mathbb{E}\{z\}_{\tilde{\mathbb{M}},d}. Then, it holds that ^me~,zf^\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{f} defines a function on some neighborhood of the origin, say UU, which can be extended to an infinite sector of bisecting direction dd, say SdS_{d}. Therefore, ^me~,zf^𝒪𝕄(S^d,𝔼)\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{f}\in\mathcal{O}^{\mathbb{M}}(\hat{S}_{d},\mathbb{E}). We apply the second part of Theorem 3 to the strongly regular sequence m¯\overline{m} and n=1n=1 in order to achieve that m¯,z^me~,zf^\partial_{\overline{m},z}\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{f}, which is an element in 𝒪(U)\mathcal{O}(U), is such that m¯,z^me~,zf^𝒪𝕄(S^d,𝔼)\partial_{\overline{m},z}\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{f}\in\mathcal{O}^{\mathbb{M}}(\hat{S}_{d},\mathbb{E}). Lemma 3 yields that

m¯,z^me~,zf^^me~,zme,zf^.\partial_{\overline{m},z}\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{f}\equiv\hat{\mathcal{B}}_{m_{\tilde{e}},z}\partial_{m_{e},z}\hat{f}.

We conclude that me,zf^\partial_{m_{e},z}\hat{f} defines a holomorphic function on some neighborhood of the origin, and admits an analytic extension to an infinite sector of bisecting direction dd, with adequate growth at infinity. This entails that me,zf^𝔼{z}𝕄~,d\partial_{m_{e},z}\hat{f}\in\mathbb{E}\{z\}_{\tilde{\mathbb{M}},d}. ∎

As a consequence of Corollary 1 and the alternative definition of summable formal power series stated in Theorem 2 the following definition makes sense.

Definition 10.

Let 𝕄~\tilde{\mathbb{M}} be a strongly regular sequence admitting a nonzero proximate order. Assume that u^𝔼{z}𝕄~,d\hat{u}\in\mathbb{E}\{z\}_{\tilde{\mathbb{M}},d}, for some dd\in\mathbb{R}. Let mem_{e} be a sequence of moments. Then, we define the operator of mem_{e}-moment differentiation of 𝒮M~,d(u^)\mathcal{S}_{\tilde{M},d}(\hat{u}) by

me,z(𝒮M~,d(u^)):=𝒮𝕄~,d(me,z(u^)).\partial_{m_{e},z}(\mathcal{S}_{\tilde{M},d}(\hat{u})):=\mathcal{S}_{\tilde{\mathbb{M}},d}(\partial_{m_{e},z}(\hat{u})).

The previous definition allows to determine upper bounds for the sum of a formal power series in the same way as in Theorem 3.

Proposition 2.

Let me=(me(p))p0m_{e}=(m_{e}(p))_{p\geq 0} be a sequence of moments. Let 𝕄~\tilde{\mathbb{M}} be a strongly regular sequence admitting a nonzero proximate order and dd\in\mathbb{R}. We choose u^𝔼{z}𝕄~,d\hat{u}\in\mathbb{E}\{z\}_{\tilde{\mathbb{M}},d} and write u=𝒮𝕄~,d(u^)𝒪(G,𝔼)u=\mathcal{S}_{\tilde{\mathbb{M}},d}(\hat{u})\in\mathcal{O}(G,\mathbb{E}) for some sectorial region G=Gd(θ)G=G_{d}(\theta) with θ>πω(𝕄~)\theta>\pi\omega(\tilde{\mathbb{M}}). Then for every GGG^{\prime}\prec G there exist C4,C5>0C_{4},\,C_{5}>0 such that

(32) me,znu(z)𝔼C4C5nme(n)M~n,\left\|\partial_{m_{e},z}^{n}u(z)\right\|_{\mathbb{E}}\leq C_{4}C_{5}^{n}m_{e}(n)\tilde{M}_{n},

for all n0n\in\mathbb{N}_{0} and zGz\in G^{\prime}.

Proof.

In view of Theorem 2 one can write u=𝒮𝕄~,d(u^)=Te~,τ(^me~,z(u^))u=\mathcal{S}_{\tilde{\mathbb{M}},d}(\hat{u})=T_{\tilde{e},\tau}(\hat{\mathcal{B}}_{m_{\tilde{e}},z}(\hat{u})), for some direction τ\tau close to dd, with e~\tilde{e} being any kernel for 𝕄~\tilde{\mathbb{M}}-summability and me~m_{\tilde{e}} its associated sequence of moments. Taking into account Definition 10 and Lemma 3, one has that

me,znu(z)=Te~,τ(^me~,z(me,znu^))(z)=Te~,τ(me~me,zn(^me~,zu^))(z),\partial_{m_{e},z}^{n}u(z)=T_{\tilde{e},\tau}(\hat{\mathcal{B}}_{m_{\tilde{e}},z}(\partial_{m_{e},z}^{n}\hat{u}))(z)=T_{\tilde{e},\tau}(\partial_{m_{\tilde{e}}m_{e},z}^{n}(\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{u}))(z),

for all n0n\in\mathbb{N}_{0} and zGd(θ)z\in G_{d}(\theta).

We observe that ^me~,zu^𝒪𝕄~(S^d(δ;r),𝔼)\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{u}\in\mathcal{O}^{\tilde{\mathbb{M}}}(\hat{S}_{d}(\delta;r),\mathbb{E}) for some δ>0\delta>0 and r>0r>0. Therefore, one may apply Theorem 3 to the sequence of moments meme~m_{e}m_{\tilde{e}} (see Lemma 3) to arrive at

(33) meme~,zn(^me~,zu^(z))𝔼C1C2nme(n)me~(n)exp(M~(C3|z|)),\left\|\partial_{m_{e}m_{\tilde{e}},z}^{n}(\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{u}(z))\right\|_{\mathbb{E}}\leq C_{1}C_{2}^{n}m_{e}(n)m_{\tilde{e}}(n)\exp\left(\tilde{M}(C_{3}|z|)\right),

for some C1,C2,C3>0C_{1},\,C_{2},\,C_{3}>0 and zS^d(δ1;r~)z\in\hat{S}_{d}(\delta_{1};\tilde{r}) for 0<δ1<δ0<\delta_{1}<\delta, 0<r~<r0<\tilde{r}<r. Let f(z):=meme~,zn(^me~,zu^(z))f(z):=\partial_{m_{e}m_{\tilde{e}},z}^{n}(\hat{\mathcal{B}}_{m_{\tilde{e}},z}\hat{u}(z)). Then, there exists C6>0C_{6}>0 such that

0(τ)e~(u/z)f(u)duu𝔼0t0e~(u/z)f(u)duu𝔼+t0(τ)e~(u/z)f(u)duu𝔼=I3+I4C6,\left\|\int_{0}^{\infty(\tau)}\tilde{e}(u/z)f(u)\frac{du}{u}\right\|_{\mathbb{E}}\leq\left\|\int_{0}^{t_{0}}\tilde{e}(u/z)f(u)\frac{du}{u}\right\|_{\mathbb{E}}+\left\|\int_{t_{0}}^{\infty(\tau)}\tilde{e}(u/z)f(u)\frac{du}{u}\right\|_{\mathbb{E}}=I_{3}+I_{4}\leq C_{6},

for some τarg(Sd(δ1))\tau\in\hbox{arg}(S_{d}(\delta_{1})). Usual estimates for e~\tilde{e}-Laplace transform lead to the conclusion: the integrability property of ee at the origin (see (8)) leads to upper bounds for I3I_{3} and I4I_{4} is bounded from above in view of (9) and the very definition of the function MM. More precisely, this holds for |arg(z)τ|<ω(M~)π/2|\hbox{arg}(z)-\tau|<\omega(\tilde{M})\pi/2 and small enough |z||z|. One may vary τ\tau among the arguments arg(Sd(δ1))\hbox{arg}(S_{d}(\delta_{1})) following the usual procedure.

Finally, the bounds in (32) are attained taking into account that 𝕄\mathbb{M} and me~m_{\tilde{e}} are equivalent sequences, in view of the remark after Definition 8. ∎

4. Application: Summability of formal solutions of moment Partial Differential Equations

This section is devoted to the study of summability properties of the formal solutions of a certain family of moment partial differential equations.

Let 𝕄\mathbb{M} be a strongly regular sequence which admits a nonzero proximate order. We assume that 𝕄1\mathbb{M}_{1} and 𝕄2\mathbb{M}_{2} are strongly regular sequences which admit nonzero proximate order. Let e1e_{1} (resp. e2e_{2}) be a kernel function for 𝕄1\mathbb{M}_{1}-summability (resp. for 𝕄2\mathbb{M}_{2}-summability), and we write m1m_{1} (resp. m2m_{2}) for its associated sequence of moments. Additionally, we assume that m1m_{1} and m2m_{2} are 𝕄\mathbb{M}-sequences of orders s1>0s_{1}>0 and s2>0s_{2}>0, respectively.

Let 1k<p1\leq k<p be integer numbers such that s2p>s1ks_{2}p>s_{1}k. Let r>0r>0. We denote D:=D(0,r)D:=D(0,r) and assume that a(z)𝒪(D¯)a(z)\in\mathcal{O}(\overline{D}) and a(z)1𝒪(D¯)a(z)^{-1}\in\mathcal{O}(\overline{D}). We also fix f^[[t,z]]\hat{f}\in\mathbb{C}[[t,z]] and φj(z)𝒪(D¯)\varphi_{j}(z)\in\mathcal{O}(\overline{D}) for j=0,,k1j=0,\ldots,k-1.

We consider the following Cauchy problem.

(34) {(m1,tka(z)m2,zp)u(t,z)=f^(t,z)m1,tju(0,z)=φj(z),j=0,,k1.\left\{\begin{array}[]{lcc}\left(\partial_{m_{1},t}^{k}-a(z)\partial_{m_{2},z}^{p}\right)u(t,z)=\hat{f}(t,z)&\\ \partial_{m_{1},t}^{j}u(0,z)=\varphi_{j}(z),&\quad j=0,\ldots,k-1.\end{array}\right.
Lemma 4.

Under the previous assumptions there exists a unique formal solution u^(t,z)[[t,z]]\hat{u}(t,z)\in\mathbb{C}[[t,z]] of the Cauchy problem (34). Moreover, in the case that f^(t,z)𝒪(D¯)[[t]]\hat{f}(t,z)\in\mathcal{O}(\overline{D})[[t]] we have u^(t,z)𝒪(D¯)[[t]]\hat{u}(t,z)\in\mathcal{O}(\overline{D})[[t]].

Proof.

Let u^(t,z)[[t,z]]\hat{u}(t,z)\in\mathbb{C}[[t,z]]. We write u^(t,z)=n0un,(z)m1(n)tn\hat{u}(t,z)=\sum_{n\geq 0}\frac{u_{n,\star}(z)}{m_{1}(n)}t^{n}, for some un,(z)[[z]]u_{n,\star}(z)\in\mathbb{C}[[z]]. The initial conditions of (34) determine uj,(z)=m1(0)φj(z)u_{j,\star}(z)=m_{1}(0)\varphi_{j}(z) in order that u^(t,z)\hat{u}(t,z) is a formal solution of (34). We plug the formal power series u^(t,z)\hat{u}(t,z) into the problem to arrive at the recurrence formula

(35) un+k,(z)=a(z)m2,zpun,(z)+f^n,(z),u_{n+k,\star}(z)=a(z)\partial_{m_{2},z}^{p}u_{n,\star}(z)+\hat{f}_{n,\star}(z),

where we write f^(t,z)=n0f^n,(z)m1(n)tn\hat{f}(t,z)=\sum_{n\geq 0}\frac{\hat{f}_{n,\star}(z)}{m_{1}(n)}t^{n}. Therefore, the elements un,(z)u_{n,\star}(z) for nkn\geq k are determined by the initial data. Furthermore, under the convergence assumption on f^\hat{f} the solution of (35) belongs to 𝒪(D¯)\mathcal{O}(\overline{D}). ∎

From now on, the pair (𝔼,𝔼)(\mathbb{E},\left\|\cdot\right\|_{\mathbb{E}}) denotes the Banach space of holomorphic functions in D¯\overline{D}, and 𝔼\left\|\cdot\right\|_{\mathbb{E}} stands for the norm

n=0anznr:=n=0|an|rn.\left\|\sum_{n=0}^{\infty}a_{n}z^{n}\right\|_{r}:=\sum_{n=0}^{\infty}|a_{n}|r^{n}.
Lemma 5.

Let m=(m(p))p0m=(m(p))_{p\geq 0} be a (lc) sequence and f𝒪(D¯)f\in\mathcal{O}(\overline{D}). If there exist C>0C>0 and n0n\in\mathbb{N}_{0} such that

(36) f(z)r~|z|nm(n)Cfor everyzD¯,r~=|z|\left\|f(z)\right\|_{\tilde{r}}\leq\frac{|z|^{n}}{m(n)}C\quad\textrm{for every}\quad z\in\overline{D},\quad\tilde{r}=|z|

then

m,zkf(z)r~|z|n+km(n+k)Cfor everyk0andzD¯,r~=|z|.\left\|\partial_{m,z}^{-k}f(z)\right\|_{\tilde{r}}\leq\frac{|z|^{n+k}}{m(n+k)}C\quad\textrm{for every}\quad k\in\mathbb{N}_{0}\quad\textrm{and}\quad z\in\overline{D},\quad\tilde{r}=|z|.
Proof.

By (36) we may write f(z)=j=nfjzj𝒪(D¯)f(z)=\sum_{j=n}^{\infty}f_{j}z^{j}\in\mathcal{O}(\overline{D}). We define the auxiliary function g(z)𝒪(D¯)g(z)\in\mathcal{O}(\overline{D}) as

g(z):=j=0|fj+n|m(n)zj.g(z):=\sum_{j=0}^{\infty}|f_{j+n}|m(n)z^{j}.

By (36) we get g(z)r~C\left\|g(z)\right\|_{\tilde{r}}\leq C and f(z)znm(n)g(z)f(z)\ll\frac{z^{n}}{m(n)}g(z). Since mm is a (lc) sequence we conclude that

m,zkf(z)=zn+km(n+k)j=0m(n+j)m(n+k)m(j+n+k)fj+nzjzn+km(n+k)j=0|fj+n|m(n)zj=zn+kg(z)m(n+k).\partial_{m,z}^{-k}f(z)=\frac{z^{n+k}}{m(n+k)}\sum_{j=0}^{\infty}\frac{m(n+j)m(n+k)}{m(j+n+k)}f_{j+n}z^{j}\ll\frac{z^{n+k}}{m(n+k)}\sum_{j=0}^{\infty}|f_{j+n}|m(n)z^{j}=\frac{z^{n+k}g(z)}{m(n+k)}.

Hence

m,zkf(z)r~|z|n+km(n+k)g(z)r~|z|n+km(n+k)C\left\|\partial_{m,z}^{-k}f(z)\right\|_{\tilde{r}}\leq\frac{|z|^{n+k}}{m(n+k)}\left\|g(z)\right\|_{\tilde{r}}\leq\frac{|z|^{n+k}}{m(n+k)}C

for every k0k\in\mathbb{N}_{0} and zD¯z\in\overline{D}, r~=|z|\tilde{r}=|z|. ∎

Theorem 4.

Under the assumptions made on the elements involved in the Cauchy problem (34) let u^(t,z)\hat{u}(t,z) be the formal solution of (34) and dd\in\mathbb{R}. We define the strongly regular sequence 𝕄¯=(Mns2pks1)n0\overline{\mathbb{M}}=(M_{n}^{\frac{s_{2}p}{k}-s_{1}})_{n\geq 0}. The following statements are equivalent:

  1. (i)

    u^(t,z)\hat{u}(t,z) is 𝕄¯\overline{\mathbb{M}}-summable along direction dd as a formal power series in 𝔼[[t]]\mathbb{E}[[t]].

  2. (ii)

    f^(t,z)𝔼[[t]]\hat{f}(t,z)\in\mathbb{E}[[t]] and m2,zju^(t,0)[[t]]\partial_{m_{2},z}^{j}\hat{u}(t,0)\in\mathbb{C}[[t]] for j=0,,p1j=0,\ldots,p-1 are 𝕄¯\overline{\mathbb{M}}-summable along direction dd.

If one of the previous equivalent statements holds then the sum of u^(t,z)\hat{u}(t,z) is an actual solution of (34).

Proof.

(i)\Rightarrow (ii). Equation (35) entails that f^(t,z)=n0f^n,(z)m1(n)tn𝔼[[t]]\hat{f}(t,z)=\sum_{n\geq 0}\frac{\hat{f}_{n,\star}(z)}{m_{1}(n)}t^{n}\in\mathbb{E}[[t]] whenever u^(t,z)𝔼[[t]]\hat{u}(t,z)\in\mathbb{E}[[t]]. In addition to this, the space 𝔼{t}𝕄¯,d\mathbb{E}\{t\}_{\overline{\mathbb{M}},d} is a differential algebra, with 𝒮𝕄¯,d\mathcal{S}_{\overline{\mathbb{M}},d} respecting the operations of addition, product and derivation (see Proposition 6.20 (i) [25]) and also under the action of the operator m2,z\partial_{m_{2},z} (see Corollary 1). Regarding equation (34) we conclude that f^(t,z)𝔼{t}𝕄¯,d\hat{f}(t,z)\in\mathbb{E}\{t\}_{\overline{\mathbb{M}},d}.

Let 0jp10\leq j\leq p-1. The same argument as before yields that the formal power series m2,zju^(t,z)𝔼{t}𝕄¯,d\partial_{m_{2},z}^{j}\hat{u}(t,z)\in\mathbb{E}\{t\}_{\overline{\mathbb{M}},d}. A direct application of the definition of summable formal power series guarantees summability of its evaluation at z=0z=0 along direction dd.

(ii)\Rightarrow (i). Let ψ^0(t):=u^(t,0)\hat{\psi}_{0}(t):=\hat{u}(t,0), and for all 1jp11\leq j\leq p-1 let ψ^j(t):=m2(0)m2(j)m2,zju^(t,0)\hat{\psi}_{j}(t):=\frac{m_{2}(0)}{m_{2}(j)}\partial_{m_{2},z}^{j}\hat{u}(t,0). We also write ω^(t,z):=m2,zpu^(t,z)\hat{\omega}(t,z):=\partial_{m_{2},z}^{p}\hat{u}(t,z). We observe that the formal power series ω^(t,z)\hat{\omega}(t,z) satisfies the equation

(11a(z)m1,tkm2,zp)ω^(t,z)=g^(t,z),\left(1-\frac{1}{a(z)}\partial_{m_{1},t}^{k}\partial_{m_{2},z}^{-p}\right)\hat{\omega}(t,z)=\hat{g}(t,z),

with

g^(t,z):=1a(z)(m1,tkψ^0(t)+zm1,tkψ^1(t)++zp1m1,tkψ^p1(t)f^(t,z)).\hat{g}(t,z):=\frac{1}{a(z)}(\partial_{m_{1},t}^{k}\hat{\psi}_{0}(t)+z\partial_{m_{1},t}^{k}\hat{\psi}_{1}(t)+\ldots+z^{p-1}\partial_{m_{1},t}^{k}\hat{\psi}_{p-1}(t)-\hat{f}(t,z)).

We write ω^(t,z)\hat{\omega}(t,z) in the form

ω^(t,z)=q0ω^q(t,z),\hat{\omega}(t,z)=\sum_{q\geq 0}\hat{\omega}_{q}(t,z),

where

ω^0(t,z):=g^(t,z), and ω^q(t,z):=1a(z)m1,tkm2,zpω^q1(t,z) for all q1.\hat{\omega}_{0}(t,z):=\hat{g}(t,z),\quad\hbox{ and }\quad\hat{\omega}_{q}(t,z):=\frac{1}{a(z)}\partial_{m_{1},t}^{k}\partial_{m_{2},z}^{-p}\hat{\omega}_{q-1}(t,z)\hbox{ for all }q\geq 1.

Observe that the hypotheses in (ii) together with the properties of differential algebra of 𝔼{t}𝕄¯,d\mathbb{E}\{t\}_{\overline{\mathbb{M}},d} guarantee that ω^0(t,z)𝔼[[t]]\hat{\omega}_{0}(t,z)\in\mathbb{E}[[t]] is 𝕄¯\overline{\mathbb{M}}-summable in direction dd. Let ω0(t,z)𝒪(G×D¯)\omega_{0}(t,z)\in\mathcal{O}(G\times\overline{D}) denote its sum, where GG stands for a sectorial region of opening larger than πω(𝕄¯)\pi\omega(\overline{\mathbb{M}}) bisected by direction dd. By Proposition 2, for all GGG^{\prime}\prec G, there exist C4,C5>0C_{4},\,C_{5}>0 such that

m1,tnω0(t,z)rC4C5nm1(n)Mns2pks1C~1C~2nMns2pk,\left\|\partial_{m_{1},t}^{n}\omega_{0}(t,z)\right\|_{r}\leq C_{4}C_{5}^{n}m_{1}(n)M_{n}^{\frac{s_{2}p}{k}-s_{1}}\leq\tilde{C}_{1}\tilde{C}_{2}^{n}M_{n}^{\frac{s_{2}p}{k}},

for some C~1,C~2>0\tilde{C}_{1},\,\tilde{C}_{2}>0, all n0n\in\mathbb{N}_{0} and tGt\in G^{\prime}. An induction argument allows to state that for every q0q\geq 0 the formal power series ω^q(t,z)𝔼[[t]]\hat{\omega}_{q}(t,z)\in\mathbb{E}[[t]] is 𝕄¯\overline{\mathbb{M}}-summable in direction dd, with

m1,tnωq(t,z)r~C~1CqC~2qk+nMqk+ns2pk|z|pqm2(pq),\left\|\partial_{m_{1},t}^{n}\omega_{q}(t,z)\right\|_{\tilde{r}}\leq\tilde{C}_{1}C^{q}\tilde{C}_{2}^{qk+n}M_{qk+n}^{\frac{s_{2}p}{k}}\frac{|z|^{pq}}{m_{2}(pq)},

for tGGt\in G^{\prime}\prec G, zD¯z\in\overline{D} with r~=|z|\tilde{r}=|z| and C=1a(z)rC=\left\|\frac{1}{a(z)}\right\|_{r}. Indeed, by Lemma 5 and by the inductive hypothesis we get

m1,tnωq+1(t,z)r~=1a(z)m2,zpm1,tk+nω^q(t,z)r~C|z|pq+pm2(pq+p)C~1CqC~2qk+n+kMqk+n+ks2pk\left\|\partial_{m_{1},t}^{n}\omega_{q+1}(t,z)\right\|_{\tilde{r}}=\left\|\frac{1}{a(z)}\partial_{m_{2},z}^{-p}\partial_{m_{1},t}^{k+n}\hat{\omega}_{q}(t,z)\right\|_{\tilde{r}}\leq C\frac{|z|^{pq+p}}{m_{2}(pq+p)}\tilde{C}_{1}C^{q}\tilde{C}_{2}^{qk+n+k}M_{qk+n+k}^{\frac{s_{2}p}{k}}

for tGGt\in G^{\prime}\prec G, zD¯z\in\overline{D} and r~=|z|\tilde{r}=|z|.

We have the following upper bound:

q0m1,tnωq(t,z)r~C~1C~2nq0(CC~2k|z|p)qMqk+ns2pk1m2(pq).\sum_{q\geq 0}\left\|\partial_{m_{1},t}^{n}\omega_{q}(t,z)\right\|_{\tilde{r}}\leq\tilde{C}_{1}\tilde{C}_{2}^{n}\sum_{q\geq 0}(C\tilde{C}_{2}^{k}|z|^{p})^{q}M_{qk+n}^{\frac{s_{2}p}{k}}\frac{1}{m_{2}(pq)}.

Due to (mg)(mg) condition, the fact that m2m_{2} is an 𝕄\mathbb{M}-sequence of order s2s_{2} (see also [15], Lemma 8) together with (3) yield

(37) Mqk+ns2pk1m2(pq)(A1qk+nMqkMn)s2pk1A3pqMpqs2=A1s2p(qk+n)kA3pqMns2pkMqks2pkMpqs2A1s2p(qk+n)kA3pqMns2pkA1qps2(k+1)k/2Mqps2Mqps2=A1s2p(qk+n)kA1qps2(k+1)k/2A3pqMns2pk,M_{qk+n}^{\frac{s_{2}p}{k}}\frac{1}{m_{2}(pq)}\leq(A_{1}^{qk+n}M_{qk}M_{n})^{\frac{s_{2}p}{k}}\frac{1}{A_{3}^{pq}M_{pq}^{s_{2}}}=\frac{A_{1}^{\frac{s_{2}p(qk+n)}{k}}}{A_{3}^{pq}}M_{n}^{\frac{s_{2}p}{k}}\frac{M_{qk}^{\frac{s_{2}p}{k}}}{M_{pq}^{s_{2}}}\\ \leq\frac{A_{1}^{\frac{s_{2}p(qk+n)}{k}}}{A_{3}^{pq}}M_{n}^{\frac{s_{2}p}{k}}\frac{A_{1}^{qps_{2}(k+1)k/2}M_{qp}^{s_{2}}}{M_{qp}^{s_{2}}}=\frac{A_{1}^{\frac{s_{2}p(qk+n)}{k}}A_{1}^{qps_{2}(k+1)k/2}}{A_{3}^{pq}}M_{n}^{\frac{s_{2}p}{k}},

for some A1,A3>0A_{1},\,A_{3}>0 We finally have

q0m1,tnωq(t,z)r~C~1C~4nMns2pkq0(A3pA1ps2+ps2(k+1)k/2CC~2k|z|p)q.\sum_{q\geq 0}\left\|\partial_{m_{1},t}^{n}\omega_{q}(t,z)\right\|_{\tilde{r}}\leq\tilde{C}_{1}\tilde{C}_{4}^{n}M_{n}^{\frac{s_{2}p}{k}}\sum_{q\geq 0}(A_{3}^{-p}A_{1}^{ps_{2}+ps_{2}(k+1)k/2}C\tilde{C}_{2}^{k}|z|^{p})^{q}.

The previous series is convergent for |z|<A3A1s2+s2(k+1)k/2(1CC~2k)1/p=:r|z|<\frac{A_{3}}{A_{1}^{s_{2}+s_{2}(k+1)k/2}}\left(\frac{1}{C\tilde{C}_{2}^{k}}\right)^{1/p}=:r^{\prime}. Therefore, one has that

ω(t,z):=q0ωq(t,z)\omega(t,z):=\sum_{q\geq 0}\omega_{q}(t,z)

defines an analytic function on G×D(0,r)G\times D(0,r^{\prime}). We reduce rr in order that rrr\leq r^{\prime}, if necessary, to arrive at

(38) q0m1,tnωq(t,z)𝔼C~3C~4nMns2pk,\sum_{q\geq 0}\left\|\partial_{m_{1},t}^{n}\omega_{q}(t,z)\right\|_{\mathbb{E}}\leq\tilde{C}_{3}\tilde{C}_{4}^{n}M_{n}^{\frac{s_{2}p}{k}},

for some C~3>0\tilde{C}_{3}>0, which is valid for all tGt\in G^{\prime}.

We show that ω(t,z)\omega(t,z) is the 𝕄¯\overline{\mathbb{M}}-sum of ω^(t,z)=q0ω^q(t,z)𝔼[[t]]\hat{\omega}(t,z)=\sum_{q\geq 0}\hat{\omega}_{q}(t,z)\in\mathbb{E}[[t]] along direction dd.

Let ee be a kernel function for 𝕄¯\overline{\mathbb{M}}-summability. Then, for all q0q\in\mathbb{N}_{0} it holds that ωq(t,z)=Te,d^me,tω^q(t,z)\omega_{q}(t,z)=T_{e,d}\hat{\mathcal{B}}_{m_{e},t}\hat{\omega}_{q}(t,z) and ω(t,z)=q0Te,d^me,tω^q(t,z)\omega(t,z)=\sum_{q\geq 0}T_{e,d}\hat{\mathcal{B}}_{m_{e},t}\hat{\omega}_{q}(t,z).

By (38) we get that Te,dω(t,z)𝒪(D×D)T^{-}_{e,d}\omega(t,z)\in\mathcal{O}(D^{\prime}\times D) for some disc at the origin DD^{\prime}. Proposition 1 can be applied to arrive at Te,dω(t,z)𝒪𝕄¯(Sd,𝔼)T^{-}_{e,d}\omega(t,z)\in\mathcal{O}^{\overline{\mathbb{M}}}(S_{d},\mathbb{E}), for some infinite sector SdS_{d} with bisecting direction dd. Hence, Te,dω(t,z)𝒪𝕄¯(S^d,𝔼)T^{-}_{e,d}\omega(t,z)\in\mathcal{O}^{\overline{\mathbb{M}}}(\hat{S}_{d},\mathbb{E}).

Finally, convergence of the series and Theorem 1 allow us to write

(39) Te,dω(t,z)=Te,dq0Te,d^me,t(ω^q(t,z))=Te,dTe,dq0^me,t(ω^q(t,z))=q0^me,t(ω^q(t,z))=^me,t(q0ω^q(t,z))=^me,tω^(t,z).T^{-}_{e,d}\omega(t,z)=T^{-}_{e,d}\sum_{q\geq 0}T_{e,d}\hat{\mathcal{B}}_{m_{e},t}(\hat{\omega}_{q}(t,z))=T^{-}_{e,d}T_{e,d}\sum_{q\geq 0}\hat{\mathcal{B}}_{m_{e},t}(\hat{\omega}_{q}(t,z))\\ =\sum_{q\geq 0}\hat{\mathcal{B}}_{m_{e},t}(\hat{\omega}_{q}(t,z))=\hat{\mathcal{B}}_{m_{e},t}\left(\sum_{q\geq 0}\hat{\omega}_{q}(t,z)\right)=\hat{\mathcal{B}}_{m_{e},t}\hat{\omega}(t,z).

Therefore, ^me,tω^(t,z)𝒪𝕄¯(S^d×D)\hat{\mathcal{B}}_{m_{e},t}\hat{\omega}(t,z)\in\mathcal{O}^{\overline{\mathbb{M}}}(\hat{S}_{d}\times D) and the formal power series ω^(t,z)\hat{\omega}(t,z) is 𝕄¯\overline{\mathbb{M}}-summable along direction dd (as an element in 𝔼[[t]]\mathbb{E}[[t]]), with sum given by ω(t,z)\omega(t,z).

Assume that one of the equivalent statements holds. Let f(t,z)f(t,z) (resp. u(t,z)u(t,z)) be the sum of f^(t,z)𝔼[[t]]\hat{f}(t,z)\in\mathbb{E}[[t]] (resp. u^(t,z)𝔼[[t]]\hat{u}(t,z)\in\mathbb{E}[[t]]) in direction dd. Then the function t(m1,tka(z)m2,zp)u(t,z)f(t,z)t\mapsto(\partial_{m_{1},t}^{k}-a(z)\partial_{m_{2},z}^{p})u(t,z)-f(t,z) with values in 𝔼\mathbb{E} admits null (𝕄¯)(\overline{\mathbb{M}})-asymptotic expansion in a sector of opening larger than ω(𝕄¯)π\omega(\overline{\mathbb{M}})\pi. Watson’s lemma (see Corollary 4.12 [24]) states that it is the null function, which entails that u(t,z)u(t,z) is an analytic solution of (34), which satisfies the Cauchy data. ∎

Analogous estimates as in the proof of Theorem 4 can be applied to achieve the next result.

Corollary 2.

Assume that s1ks2ps_{1}k\geq s_{2}p. Under the assumptions made on the elements involved in the Cauchy problem (34) let u^(t,z)\hat{u}(t,z) be the formal solution of (34) and dd\in\mathbb{R}. The following statements are equivalent:

  1. (i)

    u^(t,z)\hat{u}(t,z) is convergent in a neighborhood of the origin.

  2. (ii)

    f^(t,z)\hat{f}(t,z) and m2,zju^(t,0)\partial_{m_{2},z}^{j}\hat{u}(t,0) for j=0,,p1j=0,\ldots,p-1 are convergent in a neighborhood of the origin.

Remark: Theorem 4 is compatible with the results obtained in [15]. Indeed, equation (34) falls into the case Γ={(0,p)}\Gamma=\{(0,p)\} in Section 5 [15], and where the associated Newton polygon has one positive slope k1k_{1} if and only if s2p>s1ks_{2}p>s_{1}k and it has no positive slope in the opposite case. Indeed,

1k1=max{0,s2pks1}.\frac{1}{k_{1}}=\max\left\{0,\frac{s_{2}p}{k}-s_{1}\right\}.

Theorem 1 in [15] states that the formal solution of the equation u^(t,z)=n0un(z)tn\hat{u}(t,z)=\sum_{n\geq 0}u_{n}(z)t^{n} satisfies that for some 0<r<r0<r^{\prime}<r there exist C,H>0C,\,H>0 such that

supzD(0,r)|un(z)|CHn(Mn)1/k1,n0.\sup_{z\in D(0,r^{\prime})}|u_{n}(z)|\leq CH^{n}(M_{n})^{1/k_{1}},\quad n\in\mathbb{N}_{0}.

The result remains coherent with Theorem 2, [15].

Remark: Theorem 4 is also coherent with the results obtained in [23] in the Gevrey classical setting. See Theorem 2 and Theorem 3, [23].

References

  • [1] W. Balser, Formal power series and linear systems of meromorphic ordinary differential equations, Universitext, Springer-Verlag, New York, 2000.
  • [2] W. Balser, M. Yoshino, Gevrey order of formal power series solutions of inhomogeneous partial differential equations with constant coefficients, Funkcial. Ekvac. 53 (2010), 411–434.
  • [3] J. Chaumat, A.M. Chollet, Surjectivité de l’application restriction à un compact dans des classes de fonctions ultradifférentiables, Math. Ann. 298(1) (1994), 7–40.
  • [4] A.A.Goldberg, I.V.Ostrovskii, Value Distribution of Meromorphic Functions, Transl. Math. Monogr., vol.236, Amer. Math. Soc., Providence, RI, 2008.
  • [5] T. Dreyfus, A. Lastra, S. Malek, Multiple-scale analysis for some linear partial q-difference and differential equations with holomorphic coefficients, Adv. Differ. Equ. 2019, 326 (2019).
  • [6] K. Ichinobe, S. Adachi, On kk-summability of formal solutions to the Cauchy problems for some linear qq-difference-differential equations, Complex Differential and Difference Equations, De Gruyter Proceedings in Mathematics (2020), 447–464.
  • [7] G.K. Immink, Exact asymptotics of nonlinear difference equations with levels 1 and 1+, Ann. Fac. Sci. Toulouse T.XVII (2) (2008), 309–356.
  • [8] G. K. Immink, Accelero-summation of the formal solutions of nonlinear difference equations, Ann. Inst. Fourier (Grenoble) 61 (2011), no. 1, 1–51.
  • [9] J. Jiménez-Garrido, S. Kamimoto, A. Lastra, J. Sanz, Multisummability in Carleman ultraholomorphic classes by means of nonzero proximate orders J. Math. Anal. Appl. 472(1) (2019), 627–686.
  • [10] J. Jiménez-Garrido, J. Sanz, G. Schindl, Log-convex sequences and nonzero proximate orders, J. Math. Anal. Appl. 448(2) (2017) 1572–1599.
  • [11] J. Jiménez-Garrido, J. Sanz, G. Schindl, Injectivity and surjectivity of the asymptotic Borel map in Carleman ultraholomorphic classes, J. Math. Anal. Appl. 469 (2019), 136–168.
  • [12] A. Lastra, S. Malek, On multiscale Gevrey and Gevrey asymptotics for some linear difference differential initial value Cauchy problems, Journal of Difference Equations and Applications, 23:8 (2017), 1397–1457.
  • [13] A. Lastra, S. Malek, J. Sanz, Summability in general Carleman ultraholomorphic classes, J. Math. Anal. Appl. 430 (2015), 1175–1206.
  • [14] A. Lastra, S. Malek, J. Sanz, Strongly regular multi-level solutions of singularly perturbed linear partial differential equations, Results. Math. 70 (2016), 581–614.
  • [15] A. Lastra, S. Michalik, M. Suwińska, Estimates of formal solutions for some generalized moment partial differential equations, 2020, submitted (arXiv:1911.11998).
  • [16] S. Michalik, Summability and fractional linear partial differential equations, J. Dyn. Control Syst. 16(4) (2010), 557–584.
  • [17] S. Michalik, Analytic solutions of moment partial differential equations with constant coefficients, Funkcial. Ekvac. 56 (2013), 19–50.
  • [18] S. Michalik, Summability of formal solutions of linear partial differential equations with divergent initial data, J. Math. Anal. Appl. 406 (2013), 243–260.
  • [19] S. Michalik, Analytic summable solutions of inhomogeneous moment partial differential equations, Funkcial. Ekvac. 60 (2017), 325–351.
  • [20] S. Michalik, M. Suwińska, Gevrey estimates for certain moment partial differential equations, Complex Differential and Difference Equations, De Gruyter Proceedings in Mathematics (2020), 391–408.
  • [21] S. Michalik, B. Tkacz, The Stokes Phenomenon for some moment partial differential equations, J. Dyn. Control Syst. 25 (2019), 573–598.
  • [22] H.-J. Petzsche, On E.Borel’s theorem, Math. Ann. 282(2) (1988) 299–313.
  • [23] P. Remy, Gevrey order and summability of formal series solutions of some classes of inhomogeneous linear partial differential equations with variable coefficients. J. Dyn. Control Syst. 22, 693–711 (2016).
  • [24] J. Sanz, Flat functions in Carleman ultraholomorphic classes via proximate orders, J. Math. Anal. Appl. 415(2) (2014), 623–643.
  • [25] J. Sanz, Asymptotic analysis and summability of formal power series, Analytic, algebraic and geometric aspects of differential equations, Trends Math., Birkhäuser/Springer, Cham, (2017) 199–262.
  • [26] M. Suwińska, Gevrey estimates of formal solutions for certain moment partial differential equations with variable coefficients, 2020, submitted (arXiv:2002.02845).
  • [27] V. Thilliez, Division by flat ultradifferentiable functions and sectorial extensions, Results Math. 44 (2003), 169–188.