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

Algebra of slice regular functions
on non-symmetric domains in several quaternionic variables

Xinyuan Dou douxinyuan@ustc.edu.cn Department of Mathematics, University of Science and Technology of China, Hefei 230026, China Institute of Mathematics, AMSS, Chinese Academy of Sciences, Beijing 100190, China Ming Jin mjin@must.edu.mo Faculty of Innovation Engineering, Macau University of Science and Technology, Macau, China Guangbin Ren rengb@ustc.edu.cn Department of Mathematics, University of Science and Technology of China, Hefei 230026, China  and  Ting Yang tingy@aqnu.edu.cn School of Mathematics and Physics, Anqing Normal University, Anqing 246133, China
Abstract.

The primary objective of this paper is to establish an algebraic framework for the space of weakly slice regular functions over several quaternionic variables. We recently introduced a *-product that maintains the path-slice property within the class of path-slice functions. It is noteworthy that this *-product is directly applicable to weakly slice regular functions, as every slice regular function defined on a slice-open set inherently possesses path-slice properties. Building on this foundation, we propose a precise definition of an open neighborhood for a path γ\gamma in the path space 𝒫(n)\mathscr{P}(\mathbb{C}^{n}). This definition is pivotal in establishing the holomorphism of stem functions. Consequently, we demonstrate that the *-product of two weakly slice regular functions retains its weakly slice regular nature. This retention is facilitated by holomorphy of stem functions and their relationship with weakly slice regular functions, providing a comprehensive algebraic structure for this class of functions.

Key words and phrases:
several quaternionic variables; slice regular functions; *-product; non-axially symmetric domains; slice topology
2020 Mathematics Subject Classification:
Primary: 30G35; Secondary: 32A30
This work was supported by the China Postdoctoral Science Foundation (2021M703425), the NNSF of China (12171448), the Faculty Research Grants of the Macau University of Science and Technology (FRG-23-034-FIE), and Xiaomi Young Talents Program.

1. Introduction

Slice analysis extends the theory of holomorphic functions in one complex variable to higher dimensions, offering a bridge to multifaceted mathematical disciplines. Initially introduced in the quaternions context by Gentili and Struppa in 2006 [Gentili2006001, Gentili2007001], this theory has since expanded, revealing that power series within non-commutative quaternion algebras qualify as slice regular functions, thus broadening the scope of slice analysis.

For those interested in the diverse applications and advancements in slice analysis, we recommend various sources: for geometric function theory, see [Ren2017001, Ren2017002, Wang2017001]; for quaternionic Schur analysis, refer to [Alpay2012001]; for insights into quaternionic operator theory, consult [Alpay2015001, MR3887616, MR3967697, Gantner2020001, MR4496722]. Further extensions of slice analysis include its application to real Clifford algebras [Colombo2009002], octonions [Gentili2010001], real alternative *-algebras [Ghiloni2011001], and 2n2n-dimensional Euclidean spaces [Dou2023002].

Two principal methodologies have emerged for exploring the several-variable aspect of slice analysis, both originating from the theories in one variable. The first, strongly slice regular functions, focuses on real alternative *-algebras in one variable [Ghiloni2011001] and is characterized by its reliance on intrinsic stem functions and its applicability to axially symmetric domains. For several-variable extensions, see [Colombo2012002] for the quaternionic context, and [Ghiloni2012001] for real Clifford algebras and [Ghiloni2020001] for real alternative *-algebras. The second approach, weakly slice regular functions, involves functions satisfying the Cauchy-Riemann equations within each complex plane slice, a concept originating from the initial definition of slice regular functions [Gentili2007001]. Recent advancements in this area, such as the introduction of slice topology, have facilitated the study of non-axially symmetric domains, leading to comprehensive global theories [Dou2023001]. For a broader understanding, refer to [Dou2023001, Dou2021001] for general settings, and [Dou2021001] for the octonions context.

This paper delves into the algebraic structure of weakly slice regular functions across multiple quaternionic variables, with a focus on identifying a suitable multiplication method that maintains slice regularity. Historically, two approaches to defining the *-product have been proposed: one based on the unique slice regular extension of the *-product between restricted functions on specific complex planes [Gentili2008001, Colombo2009001], and another defined via a corresponding multiplication on their stem functions [Ghiloni2011001]. Both approaches, traditionally confined to axially symmetric slice domains, are equivalent in such contexts. However, the conventional extension theorem [Colombo2009001] proves insufficient for non-axially symmetric slice domains. Consequently, we propose utilizing the second method for defining the *-product, fulfilling the necessary conditions. In [Dou2023003], we introduced a *-product that preserves the path-slice property, thus encompassing *-products of slice regular functions, as each slice regular function on a slice-open set is inherently path-slice. It’s important to note that the criteria for a function to qualify as a stem function of a slice function are straightforward in axially symmetric slice domains. In these domains, the stem function is uniquely representable through the slice function, as per the representation formula. However, this becomes significantly more complex in non-axially symmetric slice domains. We addressed this challenge in [Dou2023003], offering auxiliary conditions to facilitate the process.

The primary focus of this paper is the slice regularity of the *-product of weakly slice regular functions. Central to this discussion is the holomorphy of stem functions. To apply differential operators to stem functions, we define an appropriate open neighborhood of a path γ\gamma in the path space 𝒫(n)\mathscr{P}(\mathbb{C}^{n}). In weak slice analysis, the corresponding stem function of a path-slice function is defined along paths, and the compatibility of stem function holomorphy with weak slice regular functions is a crucial consideration.

The paper is structured as follows: Section 2 revisits key results from [Dou2023003], including the relationship between path-slice functions and their stem functions, and the *-product between path-slice functions. Section 3 introduces holomorphic stem functions, defining a suitable open neighborhood for paths in the path space 𝒫(n)\mathscr{P}(\mathbb{C}^{n}) and establishing the holomorphy of stem functions corresponding to slice regular functions. Finally, Section 4 demonstrates that the *-product of two slice regular functions retains slice regularity, thereby confirming that the space of weakly slice regular functions, when equipped with the *-product, forms an algebra.

2. Preliminaries

This section revisits certain key concepts in the study of slice regular functions and elaborates on the *-product of path-slice functions as presented in [Dou2023003]. The quaternion algebra is denoted by \mathbb{H}. The set of imaginary units of quaternions is represented as:

𝕊:={I:I2=1}.\mathbb{S}:=\{I\in\mathbb{H}:I^{2}=-1\}.

For nn\in\mathbb{N}, the nn-dimensional weakly slice cone is defined as:

sn:=I𝕊In,\mathbb{H}_{s}^{n}:=\bigcup_{I\in\mathbb{S}}\mathbb{C}_{I}^{n},

with slice topology

τs(sn):={Ωsn:ΩIτ(In),I𝕊},\tau_{s}(\mathbb{H}^{n}_{s}):=\{\Omega\subset\mathbb{H}^{n}_{s}:\Omega_{I}\in\tau(\mathbb{C}_{I}^{n}),\ \forall\ I\in\mathbb{S}\},

where I=+I\mathbb{C}_{I}=\mathbb{R}+\mathbb{R}I and In=(I)n\mathbb{C}_{I}^{n}=(\mathbb{C}_{I})^{n}

ΩI:=ΩIn.\Omega_{I}:=\Omega\cap\mathbb{C}_{I}^{n}.

Open sets, connected sets, and paths in τs\tau_{s} are termed slice-open sets, slice-connected sets, and slice-paths in Ω\Omega, respectively.

For I𝕊I\in\mathbb{S}, define:

ΨiI:n\xlongrightarrow[]In,x+yi\xlongrightarrow[]x+yI.\begin{split}\Psi_{i}^{I}:\quad\mathbb{C}^{n}\quad&\xlongrightarrow[\hskip 28.45274pt]{}\quad\mathbb{C}_{I}^{n},\\ x+yi\ &\shortmid\!\xlongrightarrow[\hskip 28.45274pt]{}\ x+yI.\end{split}

The set of continuous paths with real-valued start points in n\mathbb{C}^{n} is denoted as:

𝒫(n):={γ:[0,1]n:γ is continuous with γ(0)n}.\mathscr{P}(\mathbb{C}^{n}):=\{\gamma:[0,1]\rightarrow\mathbb{C}^{n}:\gamma\text{ is continuous with }\gamma(0)\in\mathbb{R}^{n}\}.

For a subset Ωsn\Omega\subset\mathbb{H}_{s}^{n}, we define:

𝒫(n,Ω):={δ𝒫(n):I𝕊, such that δIΩ},\mathscr{P}(\mathbb{C}^{n},\Omega):=\left\{\delta\in\mathscr{P}(\mathbb{C}^{n}):\exists\ I\in\mathbb{S},\text{ such that }\delta^{I}\subset\Omega\right\},

where δI:=ΨiI(δ)\delta^{I}:=\Psi_{i}^{I}(\delta) is a path in ΩI\Omega_{I}. For a fixed path γ𝒫(n)\gamma\in\mathscr{P}(\mathbb{C}^{n}), we define:

𝕊(Ω,γ):={I𝕊:γIΩ}.{\mathbb{S}(\Omega,\gamma)}:=\{I\in\mathbb{S}:\gamma^{I}\subset\Omega\}.
Definition 2.1.

A function f:Ωf:\Omega\rightarrow\mathbb{H}, for Ωsn\Omega\subset\mathbb{H}_{s}^{n}, is termed path-slice if it correlates with a function F:𝒫(n,Ω)2×1F:\mathscr{P}(\mathbb{C}^{n},\Omega)\rightarrow\mathbb{H}^{2\times 1} such that:

(2.1) fγI(1)=(1,I)F(γ),f\circ\gamma^{I}(1)=(1,I)F(\gamma),

for every γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma). The function FF is called a path-slice stem function of ff. The set of path-slice functions defined on Ω\Omega is denoted as 𝒫𝒮(Ω)\mathcal{PS}(\Omega), and 𝒫𝒮𝒮(f)\mathcal{PSS}(f) denotes the set of path-slice stem functions of f𝒫𝒮(Ω)f\in\mathcal{PS}(\Omega).

Definition 2.2.

For Ω\Omega in the slice topology τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}), a function f:Ωf:\Omega\rightarrow\mathbb{H} is considered (weakly) slice regular if, for every I𝕊I\in\mathbb{S}, the restriction fI:=f|ΩIf_{I}:=f|{\Omega_{I}} is (left II-) holomorphic. This means fIf_{I} is real differentiable and satisfies:

12(x+Iy)fI(x+yI)=0\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)f_{I}(x+yI)=0

on ΩI\Omega_{I} for each =1,2,,d\ell=1,2,...,d. The class of weakly slice regular functions on Ω\Omega is denoted as 𝒮(Ω)\mathcal{SR}(\Omega).

According to [Dou2023002, Corollary 5.9], weakly slice regular functions qualify as path-slice.

Definition 2.3.

A subset Ωsn\Omega\subset\mathbb{H}_{s}^{n} is defined as real-path-connected if, for each point qΩq\in\Omega, there exists a path γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and an element I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma) such that γI(1)=q\gamma^{I}(1)=q.

Given a subset Ωsn\Omega\subset\mathbb{H}_{s}^{n}, denote

(2.2) 𝒫2(n,Ω):={(α,β)[𝒫(n,Ω)]2:α(1)=β(1)}.\mathscr{P}_{*}^{2}(\mathbb{C}^{n},\Omega):=\{(\alpha,\beta)\in[\mathscr{P}(\mathbb{C}^{n},\Omega)]^{2}:\alpha(1)=\beta(1)\}.
Definition 2.4.

For subsets Ω1,Ω2sn\Omega_{1},\Omega_{2}\subset\mathbb{H}_{s}^{n}, Ω2\Omega_{2} is termed Ω1\Omega_{1}-stem-preserving if it satisfies the following conditions:

  1. (i)

    |𝕊(Ω2,γ)|2\left|\mathbb{S}(\Omega_{2},\gamma)\right|\geqslant 2, for every γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}).

  2. (ii)

    |𝕊(Ω2,α)𝕊(Ω2,β)|1\left|\mathbb{S}(\Omega_{2},\alpha)\cap\mathbb{S}(\Omega_{2},\beta)\right|\neq 1, for each (α,β)𝒫2(n,Ω1)(\alpha,\beta)\in\mathscr{P}_{*}^{2}(\mathbb{C}^{n},\Omega_{1}).

Based on [1, Proposition 3.7], the subsequent definition is deemed well-defined:

Definition 2.5.

For a real-path-connected subset Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n}, an Ω1\Omega_{1}-stem-preserving subset Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n}, and a path-slice function f:Ω2f:\Omega_{2}\rightarrow\mathbb{H}, the following mapping

(2.3) FΩ1f:𝒫(n,Ω1)\xlongrightarrow[]2×1γ\xlongrightarrow[]G|𝒫(n,Ω1),\begin{split}F_{\Omega_{1}}^{f}:\mathscr{P}(\mathbb{C}^{n},\Omega_{1})&\xlongrightarrow[\hskip 28.45274pt]{}\mathbb{H}^{2\times 1}\\ \gamma&\shortmid\!\xlongrightarrow[\hskip 28.45274pt]{}G|_{\mathscr{P}(\mathbb{C}^{n},\Omega_{1})},\end{split}

is well defined and independent of the chosen G𝒫𝒮𝒮(f)G\in\mathcal{PSS}(f).

Proposition 2.6.

[Dou2023003, Proposition 3.3]. Given a domain Ωsn\Omega\subset\mathbb{H}_{s}^{n}, a function f𝒫𝒮(Ω)f\in\mathcal{PS}(\Omega), a path γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega), and a path-slice stem function FF associated with ff, for any distinct I,J𝕊(Ω,γ)I,J\in\mathbb{S}(\Omega,\gamma), it holds that

(2.4) F(γ)=(1I1J)1(fγI(1)fγJ(1)).F(\gamma)=\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}f\circ\gamma^{I}(1)\\ f\circ\gamma^{J}(1)\end{pmatrix}.

Define the function :sn𝕊0\mathfrak{I}:\mathbb{H}_{s}^{n}\rightarrow\mathbb{S}\cup{0} for q=(q1,,qn)snq=(q_{1},...,q_{n})\in\mathbb{H}_{s}^{n} by

(q)={0,if qn,qıRe(qı)|qıRe(qı)|,otherwise,\mathfrak{I}(q)=\begin{cases}0,&\text{if }q\in\mathbb{R}^{n},\\ \displaystyle\frac{q_{\imath}-Re(q_{\imath})}{|q_{\imath}-Re(q_{\imath})|},&\text{otherwise},\end{cases}

where ı{1,,n}\imath\in\{1,...,n\} is the smallest index such that qıq_{\imath}\notin\mathbb{R}.

From [Dou2023003, Proposition 3.11], we define:

Definition 2.7.

For a real-path-connected domain Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} and an Ω1\Omega_{1}-stem-preserving domain Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n}, if f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} is path-slice, then

(2.5) Ω1f(q)={FΩ1f(γ),if qn,(f(q),0)T,otherwise,\mathscr{F}_{\Omega_{1}}^{f}(q)=\begin{cases}F_{\Omega_{1}}^{f}(\gamma),&\text{if }q\notin\mathbb{R}^{n},\\ (f(q),0)^{T},&\text{otherwise},\end{cases}

is well-defined for qΩ1q\in\Omega_{1}, where γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) satisfies γ(q)Ω1\gamma^{\mathfrak{I}(q)}\subset\Omega_{1} and γ(q)(1)=q\gamma^{\mathfrak{I}(q)}(1)=q.

Definition 2.8.

Given real-path-connected Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n}, an Ω1\Omega_{1}-stem-preserving Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n}, and functions f𝒫𝒮(Ω1)f\in\mathcal{PS}(\Omega_{1}), g𝒫𝒮(Ω2)g\in\mathcal{PS}(\Omega_{2}), the *-product of ff and gg is defined as

(2.6) fg:=(f,f)Ω1g:Ω1.f*g:=(f,\mathfrak{I}f)\mathscr{F}_{\Omega_{1}}^{g}:\Omega_{1}\rightarrow\mathbb{H}.

For p:=(p1,p2)T,q:=(q1,q2)T2×1p:=(p_{1},p_{2})^{T},q:=(q_{1},q_{2})^{T}\in\mathbb{H}^{2\times 1}, denote the product pqp*q as

pq:=(p1𝕀+p2σ)(q1𝕀+q2σ)e1,p*q:=(p_{1}\mathbb{I}+p_{2}\sigma)(q_{1}\mathbb{I}+q_{2}\sigma)e_{1},

where 𝕀\mathbb{I}, σ\sigma, and e1e_{1} are defined as

𝕀:=(1001),σ:=(0110),e1:=(10).\mathbb{I}:=\begin{pmatrix}1&0\\ 0&1\end{pmatrix},\ \sigma:=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix},\ e_{1}:=\begin{pmatrix}1\\ 0\end{pmatrix}.

For functions F,G:𝒫(n,Ω)2×1F,G:\mathscr{P}(\mathbb{C}^{n},\Omega)\rightarrow\mathbb{H}^{2\times 1} for a domain Ωsn\Omega\subset\mathbb{H}_{s}^{n}, define

FG(γ):=F(γ)G(γ),γ𝒫(n,Ω).F*G(\gamma):=F(\gamma)*G(\gamma),\qquad\forall\ \gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega).
Lemma 2.9.

[Dou2023003, Lemma 4.2]. Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, and Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving domain. For cc\in\mathbb{H}, g𝒫𝒮(Ω2)g\in\mathcal{PS}(\Omega_{2}), a path γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), and I𝕊(Ω1,γ)I\in\mathbb{S}(\Omega_{1},\gamma) where q:=γI(1)q:=\gamma^{I}(1), it follows that

(2.7) (c,Ic)FΩ1g(γ)=(c,Ic)FΩ1g(γ¯)=(c,(q)c)Ω1g(q).(c,Ic)F_{\Omega_{1}}^{g}(\gamma)=(c,-Ic)F_{\Omega_{1}}^{g}(\overline{\gamma})=\left(c,\mathfrak{I}(q)c\right)\mathscr{F}_{\Omega_{1}}^{g}(q).
Definition 2.10.

A domain Ωsn\Omega\subset\mathbb{H}_{s}^{n} is termed self-stem-preserving if it is real-path-connected and Ω\Omega-stem-preserving.

Theorem 2.11.

[Dou2023003, Theorem 5.3]. If a domain Ωsn\Omega\subset\mathbb{H}_{s}^{n} is self-stem-preserving, then the structure (𝒫𝒮(Ω),+,)(\mathcal{PS}(\Omega),+,*) forms an associative unitary real algebra.

3. Holomorphic stem functions

In this section, we explore the concept of holomorphy for path-dependent stem functions.

We discover that a path-slice stem function FF derived from a path-slice function ff is not inherently holomorphic. However, the specific restriction FΩ1fF_{\Omega_{1}}^{f} of FF exhibits holomorphy.

For any two points z,wz,w in the complex space n\mathbb{C}^{n}, we define a standard path connecting zz to ww. This is a function zw:[0,1]n\mathcal{L}_{z}^{w}:[0,1]\to\mathbb{C}^{n} given by the formula

zw(t)=(1t)z+tw.\mathcal{L}_{z}^{w}(t)=(1-t)z+tw.

For a path γ𝒫(n)\gamma\in\mathscr{P}(\mathbb{C}^{n}) and a positive radius rr, Bn(γ(1),r)B_{\mathbb{C}^{n}}(\gamma(1),r) represents a ball in n\mathbb{C}^{n} centered at the endpoint of the path γ(1)\gamma(1) with radius rr.

The ball in the path space 𝒫(n)\mathscr{P}(\mathbb{C}^{n}) is defined as the set of all paths one can form by taking the standard path γ(1)z\mathcal{L}_{\gamma(1)}^{z} from the end point of γ\gamma to the point zz inside the ball Bn(γ(1),r)B_{\mathbb{C}^{n}}(\gamma(1),r) in n\mathbb{C}^{n}. Essentially, this process creates new paths in 𝒫(n)\mathscr{P}(\mathbb{C}^{n}) by extending γ\gamma to reach any point within a radius rr from its endpoint, and each of these new paths starts from the origin of γ\gamma and ends at some point within this spherical region in n\mathbb{C}^{n}. This concept is mathematically expressed as

(3.1) B𝒫(n)(γ,r):={γγ(1)z:zB(γ(1),r)}.B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r):=\left\{\gamma\circ\mathcal{L}_{\gamma(1)}^{z}:z\in B_{\mathbb{C}}(\gamma(1),r)\right\}.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} and γ𝒫(n)\gamma\in\mathscr{P}(\mathbb{C}^{n}). We consider a subset 𝕊\mathbb{S}^{\prime} of 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma), which could be specific selections or categories within a larger set that are related to the region Ω\Omega and the path γ\gamma.

We need to introduce some concepts which are about finding the largest possible balls in different spaces, and how these balls fit within certain regions.

  • rγ,Ωr_{\gamma,\Omega}: This represents the supremum of radii rr for which the ball in the path space centered at γ\gamma with radius rr is entirely contained within the path space related to Ω\Omega. Geometrically, this is finding the largest possible radius of a ball in the path space, centered at γ\gamma, that still fits entirely within the region defined by Ω\Omega.

  • rγ,Ωr_{\gamma,\scriptscriptstyle\Omega}: This is the supremum of radii rr for which the ball centered at the endpoint of γ\gamma (in the image space of γ\gamma) with radius rr is contained within ΩI\Omega_{I} for all selections II in 𝕊\mathbb{S}^{\prime}. Geometrically, it is about finding the largest radius of a ball in the image space that fits within the regions ΩI\Omega_{I} for every category in 𝕊\mathbb{S}^{\prime}.

  • rγ,Ω2r_{\gamma,\scriptscriptstyle\Omega}^{2}: This represents the supremum of rγ,Ω𝕊′′r_{\gamma,\scriptscriptstyle\Omega}^{\mathbb{S}^{\prime\prime}} over all subsets 𝕊′′\mathbb{S}^{{}^{\prime\prime}} of 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma) that have at least two elements. Geometrically, this is finding the largest ball (in terms of radius) that fits within the ΩI\Omega_{I} regions for every possible subset of 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma) that contains two or more elements.

These concepts can be expressed as

rγ,Ω\displaystyle r_{\gamma,\scriptscriptstyle\Omega} :=\displaystyle:= sup{r[0,+):B𝒫(n)(γ,r)𝒫(n,Ω)},\displaystyle\sup\left\{r\in[0,+\infty):B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r)\subset\mathscr{P}(\mathbb{C}^{n},\Omega)\right\},
rγ,Ω𝕊\displaystyle r_{\gamma,\scriptscriptstyle\Omega}^{\mathbb{S}^{\prime}} :=\displaystyle:= sup{r[0,+):BI(γI(1),r)ΩI,I𝕊}.\displaystyle\sup\left\{r\in[0,+\infty):B_{I}\left(\gamma^{I}(1),r\right)\subset\Omega_{I},\ \forall\ I\in\mathbb{S}^{\prime}\right\}.
rγ,Ω2\displaystyle r_{\gamma,\scriptscriptstyle\Omega}^{2} :=\displaystyle:= sup{rγ,Ω𝕊′′:𝕊′′𝕊(Ω,γ) with |𝕊′′|2}.\displaystyle\sup\left\{r_{\gamma,\scriptscriptstyle\Omega}^{\mathbb{S}^{\prime\prime}}:\mathbb{S}^{\prime\prime}\subset\mathbb{S}(\Omega,\gamma)\mbox{ with }|\mathbb{S}^{\prime\prime}|\geqslant 2\right\}.

Under certain conditions, these constants remain positive, which proves to be valuable for practical applications.

Lemma 3.1.

For any set Ωτs(sn)\Omega\in\tau_{s}(\mathbb{H}_{s}^{n}), it holds that

rγ,Ω>0r_{\gamma,\scriptscriptstyle\Omega}>0

for every path γ\gamma in the set 𝒫(n,Ω).\mathscr{P}(\mathbb{C}^{n},\Omega).

Proof.

Assume γ\gamma is a path in the path space 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega) and let II be an element of 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma). Given the property that Ω\Omega is a slice-open set, we can find a positive radius rr such that the ball BI(γI(1),r)B_{I}(\gamma^{I}(1),r) is entirely contained within the subset (Ω)I(\Omega)_{I} of Ω\Omega.

We now consider any zB(γ(1),r)z\in B_{\mathbb{C}}(\gamma(1),r). For this choice of zz, we derive the following implications

(γ(1)z)IBI(γI(1),r)ΩI.\left(\mathcal{L}_{\gamma(1)}^{z}\right)^{I}\subseteq B_{I}(\gamma^{I}(1),r)\subseteq\Omega_{I}.

Subsequently, we deduce

(γγ(1)z)I=γI(γ(1)z)IΩI.\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right)^{I}=\gamma^{I}\circ\left(\mathcal{L}_{\gamma(1)}^{z}\right)^{I}\subseteq\Omega_{I}.

This leads to the inference that

γγ(1)z𝒫(n,Ω).\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\in\mathscr{P}(\mathbb{C}^{n},\Omega).

Since B𝒫(n)(γ,r)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r) is a subset of 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega), it follows that rγ,Ωr>0r_{\gamma,\scriptscriptstyle\Omega}\geqslant r>0. ∎

Lemma 3.2.

Suppose Ω1\Omega_{1} is a real-path-connected subset of τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}), and Ω2\Omega_{2}, also in τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}), is Ω1\Omega_{1}-stem-preserving. Then, for every path γ\gamma in the path space 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), the following inequality holds:

(3.2) rγ,Ω22>0.r_{\gamma,\scriptscriptstyle\Omega_{2}}^{2}>0.
Proof.

Assume γ\gamma is a path in the path space 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}). Given that Ω2\Omega_{2} in τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}) is Ω1\Omega_{1}-stem-preserving, the set 𝕊(Ω2,γ)\mathbb{S}(\Omega_{2},\gamma) contains at least two distinct elements. Let II and JJ be two such distinct elements in 𝕊(Ω2,γ)\mathbb{S}(\Omega_{2},\gamma). Since Ω2\Omega_{2} is slice-open, there exists a positive radius rr satisfying

BL(γL(1),r)(Ω2)L,B_{L}(\gamma^{L}(1),r)\subset(\Omega_{2})_{L},

for every L{I,J}.L\in\{I,J\}. This leads to the conclusion that

0<rrγ,Ω2{I,J}rγ,Ω22.0<r\leqslant r^{\{I,J\}}_{\gamma,\Omega_{2}}\leqslant r^{2}_{\gamma,\Omega_{2}}.

Lemma 3.3.

Consider Ω1\Omega_{1} as a subset of τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}) that is real-path-connected, and Ω2\Omega_{2} also in τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}), which is Ω1\Omega_{1}-stem-preserving. Let γ\gamma be a path in the path space 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) and r1r_{1} be arbitrary from the interval (0,min{rγ,Ω22,rγ,Ω1})\left(0,\min\{r_{\gamma,\scriptscriptstyle\Omega_{2}}^{2},r_{\gamma,\scriptscriptstyle\Omega_{1}}\}\right). Then, there exist distinct elements I,JI,J in 𝕊(Ω2,γ)\mathbb{S}(\Omega_{2},\gamma) such that for every path β\beta in the ball B𝒫(n)(γ,r1)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r_{1}), which is a subset of 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), the paths βI\beta^{I} and βJ\beta^{J} are contained within Ω2\Omega_{2}, i.e.,

(3.3) βI,βJΩ2,βB𝒫(n)(γ,r1).\beta^{I},\beta^{J}\subset\Omega_{2},\qquad\forall\ \beta\in B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r_{1}).
Proof.

Given that r1<rγ,Ω1r_{1}<r_{\gamma,\scriptscriptstyle\Omega_{1}}, we have the inclusion

B𝒫(n)(γ,r1)𝒫(n,Ω1).B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r_{1})\subseteq\mathscr{P}(\mathbb{C}^{n},\Omega_{1}).

From the condition r1<rγ,Ω22r_{1}<r_{\gamma,\scriptscriptstyle\Omega_{2}}^{2}, there exists a subset 𝕊\mathbb{S}^{\prime} of 𝕊(Ω2,γ)\mathbb{S}(\Omega_{2},\gamma) with |𝕊|2|\mathbb{S}^{\prime}|\geq 2 satisfying

BL(γL(1),r1)(Ω2)LB_{L}(\gamma^{L}(1),r_{1})\subseteq(\Omega_{2})_{L}

for each L𝕊.L\in\mathbb{S}^{\prime}. Select distinct elements I,J𝕊I,J\in\mathbb{S}^{\prime} and consider any path βB𝒫(n)(γ,r1)\beta\in B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r_{1}). By the definition of β\beta, we have β=γγ(1)β(1)\beta=\gamma\circ\mathcal{L}_{\gamma(1)}^{\beta(1)}. Then, for L{I,J}L\in\{I,J\}, the following holds:

(γ(1)β(1))LBL(γL(1),r)(Ω2)L.\left(\mathcal{L}_{\gamma(1)}^{\beta(1)}\right)^{L}\subseteq B_{L}(\gamma^{L}(1),r)\subseteq(\Omega_{2})_{L}.

Consequently, this implies

βL=(γγ(1)β(1))L=γL(γ(1)β(1))L(Ω2)L,\beta^{L}=\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{\beta(1)}\right)^{L}=\gamma^{L}\circ\left(\mathcal{L}_{\gamma(1)}^{\beta(1)}\right)^{L}\subseteq(\Omega_{2})_{L},

for L{I,J}L\in\{I,J\}. Thus, the condition (3.3) is satisfied. ∎

For any γ𝒫(n)\gamma\in\mathscr{P}(\mathbb{C}^{n}) and r>0r>0, we introduce a map

(3.4) γ:B(γ(1),r)\xlongrightarrow[]B𝒫(n)(γ,r),\begin{split}\mathscr{L}_{\gamma}:B_{\mathbb{C}}(\gamma(1),r)\xlongrightarrow[\hskip 14.22636pt]{}B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r),\end{split}

defined by

γ(z)=γγ(1)z.\mathscr{L}_{\gamma}(z)=\gamma\circ\mathcal{L}_{\gamma(1)}^{z}.

The transformation γγ(1)z\gamma\circ\mathcal{L}_{\gamma(1)}^{z} geometrically represents the extension of the original path to reach the point zz. It illustrates how a point-wise approach (selecting a point zz) can lead to a new understanding or interpretation in a path-dependent framework (creating a new path in 𝒫(n)\mathscr{P}(\mathbb{C}^{n})). This highlights the interconnectedness and complementary nature of point-wise and path-dependent perspectives in slice analysis.

Proposition 3.4.

Let Ω1τs(sn)\Omega_{1}\in\tau_{s}(\mathbb{H}_{s}^{n}) be real-path-connected, Ω2τs(sn)\Omega_{2}\in\tau_{s}(\mathbb{H}_{s}^{n}) be Ω1\Omega_{1}-stem-preserving, γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) and f𝒮(Ω2)f\in\mathcal{SR}(\Omega_{2}). Then, the following equation holds:

(3.5) FΩ1fγ(z)=(1I1J)1(f(zI)f(zJ))\displaystyle F_{\Omega_{1}}^{f}\circ\mathscr{L}_{\gamma}(z)=\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}f(z^{I})\\ f(z^{J})\end{pmatrix}

for any zz in the complex ball Bn(γ(1),r1)B_{\mathbb{C}^{n}}\left(\gamma(1),r_{1}\right), where r1(0,min{rγ,Ω22,rγ,Ω1})r_{1}\in\left(0,\min\left\{r_{\gamma,\scriptscriptstyle\Omega_{2}}^{2},r_{\gamma,\scriptscriptstyle\Omega_{1}}\right\}\right) and I,J𝕊(Ω2,γ)I,J\in\mathbb{S}(\Omega_{2},\gamma) with IJI\neq J and (3.3) holds.

Proof.

Suppose we take any zz from B(γ(1),r1)B_{\mathbb{C}}\left(\gamma(1),r_{1}\right). Define β\beta as

γ(z)=γγ(1)z\mathscr{L}_{\gamma}(z)=\gamma\circ\mathcal{L}_{\gamma(1)}^{z}

and let F𝒫𝒮𝒮(f)F\in\mathcal{PSS}(f). It follows that:

β(1)=[γγ(1)z](1)=z.\beta(1)=\left[\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right](1)=z.

Given (3.3), II and JJ are elements of 𝕊(Ω2,β)\mathbb{S}(\Omega_{2},\beta). Referring to (2.3) and (2.4), we can deduce

FΩ1fγ(z)=FΩ1f(β)=F(β)=(1I1J)1(fβI(1)fβJ(1))=(1I1J)1(f(zI)f(zJ)).\begin{split}F_{\Omega_{1}}^{f}\circ\mathscr{L}_{\gamma}(z)=F_{\Omega_{1}}^{f}(\beta)=F(\beta)=&\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}f\circ\beta^{I}(1)\\ f\circ\beta^{J}(1)\end{pmatrix}\\ =&\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}f\left(z^{I}\right)\\ f\left(z^{J}\right)\end{pmatrix}.\end{split}

This completes the proof. ∎

We are now in a position to define the notion of a function being holomorphic over a space of paths.

Definition 3.5.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} and γ𝒫(n)\gamma\in\mathscr{P}(\mathbb{C}^{n}). A function F:𝒫(n,Ω)2×1F:\mathscr{P}(\mathbb{C}^{n},\Omega)\rightarrow\mathbb{H}^{2\times 1} is called holomorphic at γ\gamma, if there is r>0r>0 such that B𝒫(n)(γ,r)𝒫(n,Ω)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r)\subset\mathscr{P}(\mathbb{C}^{n},\Omega) and

(3.6) 12(x+σy)(Fγ)(x+yi)=0,\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\left(F\circ\mathscr{L}_{\gamma}\right)(x+yi)=0,

for each x+yiB(γ(1),r)x+yi\in B_{\mathbb{C}}(\gamma(1),r) and {1,,n}\ell\in\{1,...,n\}.

Moreover, we call FF is holomorphic in U𝒫(n,Ω)U\subset\mathscr{P}(\mathbb{C}^{n},\Omega) if FF is holomorphic at each γU\gamma\in U.

Proposition 3.6.

Let Ω1τs(sn)\Omega_{1}\in\tau_{s}(\mathbb{H}_{s}^{n}) be real-path-connected, Ω2τs(sn)\Omega_{2}\in\tau_{s}(\mathbb{H}_{s}^{n}) be Ω1\Omega_{1}-stem-preserving, and f𝒮(Ω2)f\in\mathcal{SR}(\Omega_{2}). Then FΩ1fF_{\Omega_{1}}^{f} is holomorphic in 𝒫(n,Ω).\mathscr{P}(\mathbb{C}^{n},\Omega).

Proof.

For each path γ\gamma in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), let us select r1r_{1} from the interval (0,min(rγ,Ω22,rγ,Ω1))(0,\min(r_{\gamma,\scriptscriptstyle\Omega_{2}}^{2},r_{\gamma,\scriptscriptstyle\Omega_{1}})). We choose two distinct elements I,JI,J from 𝕊(Ω2,γ)\mathbb{S}(\Omega_{2},\gamma) such that equation (3.3) is satisfied.

Observe the following identity:

σ(1I1J)1=(1I1J)1(IJ).\sigma\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}=\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}I&\\ &J\end{pmatrix}.

Upon direct computation, we arrive at

12(x+σy)(FΩ1fγ)(z)\displaystyle\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\left(F_{\Omega_{1}}^{f}\circ\mathscr{L}{\gamma}\right)(z)
=\displaystyle= 12(x+σy)(1I1J)1(f(zI)f(zJ))\displaystyle\frac{1}{2}\left(\frac{\partial}{\partial x\ell}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}f(z^{I})\\ f(z^{J})\end{pmatrix}
=\displaystyle= (1I1J)1[12(x+(IJ)y)](f(zI)f(zJ))\displaystyle\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\left[\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\begin{pmatrix}I\\ &J\end{pmatrix}\frac{\partial}{\partial y_{\ell}}\right)\right]\begin{pmatrix}f(z^{I})\\ f(z^{J})\end{pmatrix}
=\displaystyle= (1I1J)1(12(x+Iy)f(zI)12(x+Jy)f(zJ))\displaystyle\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)f(z^{I})\\ \frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+J\frac{\partial}{\partial y_{\ell}}\right)f(z^{J})\end{pmatrix}
=\displaystyle= (1I1J)1(00)=0,\displaystyle\begin{pmatrix}1&I\\ 1&J\end{pmatrix}^{-1}\begin{pmatrix}0\\ 0\end{pmatrix}=0,

for every z=x+yiz=x+yi in B(γ(1),r1)B_{\mathbb{C}}(\gamma(1),r_{1}) and for all 1,,n\ell\in{1,...,n}. By the definition in (3.6), it follows that FΩ1fF_{\Omega_{1}}^{f} is holomorphic in 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega). ∎

4. Star-product of slice regular functions

In this section, we show that the *-product of slice regular functions is still slice regular. It implies that the class of slice regular functions is an associative unitary real algebra. For this purpose, we will require the assistance of several results.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n}, γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma). Denote

rγ,ΩI:=sup{r[0,+):BI(γI(1),r)ΩI}.r_{\gamma,\Omega}^{I}:=\sup\left\{r\in[0,+\infty):B_{I}\left(\gamma^{I}(1),r\right)\subset\Omega_{I}\right\}.

Geometrically, it represents the largest radius rr for which a ball BI(γI(1),r)B_{I}(\gamma^{I}(1),r), centered at the point γI(1)\gamma^{I}(1) in the slice II of Ω\Omega, remains entirely inside the subset ΩI\Omega_{I} of Ω\Omega associated with the slice II. In simpler terms, it is the largest ball you can draw around the endpoint of the path γI\gamma^{I} without leaving the space ΩI\Omega_{I}.

Lemma 4.1.

Let Ωτs(sn)\Omega\in\tau_{s}(\mathbb{H}_{s}^{n}), γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma). It holds that

rγ,ΩI>0.r_{\gamma,\Omega}^{I}>0.

Furthermore, for any point zz within the complex ball B(γ(1),rγ,ΩI)B_{\mathbb{C}}(\gamma(1),r_{\gamma,\Omega}^{I}), we have

(4.1) γ(z)𝒫(n,Ω).\mathscr{L}_{\gamma}(z)\in\mathscr{P}(\mathbb{C}^{n},\Omega).
Proof.

Given that the point γI(1)\gamma^{I}(1) is contained in Ω\Omega and ΩI\Omega_{I} is a member of the topology τ(In)\tau(\mathbb{C}_{I}^{n}), there exists a positive radius rr satisfying

BI(γI(1),r)ΩI.B_{I}\left(\gamma^{I}(1),r\right)\subset\Omega_{I}.

This observation ensures that rγ,ΩIr_{\gamma,\Omega}^{I} is greater than zero.

From its definition, it follows that

BI(γI(1),r)ΩI,B_{I}\left(\gamma^{I}(1),r^{\prime}\right)\subseteq\Omega_{I},

for all r(0,rγ,ΩI)r^{\prime}\in\left(0,r_{\gamma,\Omega}^{I}\right). Consequently, we can assert that

(4.2) BI(γI(1),rγ,ΩI)=r(0,rγ,ΩI)BI(γI(1),r)ΩI.B_{I}\left(\gamma^{I}(1),r_{\gamma,\Omega}^{I}\right)=\bigcup_{r^{\prime}\in\left(0,r_{\gamma,\Omega}^{I}\right)}B_{I}\left(\gamma^{I}(1),r^{\prime}\right)\subset\Omega_{I}.

For any point zz in the complex ball B(γ(1),rγ,ΩI)B_{\mathbb{C}}\left(\gamma(1),r_{\gamma,\Omega}^{I}\right) and considering equation (4.2), we have that (γ(1)z)I\left(\mathcal{L}_{\gamma(1)}^{z}\right)^{I} is a subset of ΩI\Omega_{I}. This implies

(γ(z))I=(γγ(1)z)I=γI(γ(1)z)IΩI.\left(\mathscr{L}_{\gamma}(z)\right)^{I}=\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right)^{I}=\gamma^{I}\circ\left(\mathcal{L}_{\gamma(1)}^{z}\right)^{I}\subset\Omega_{I}.

Hence, equation (4.1) is satisfied, completing the proof. ∎

Proposition 4.2.

Let Ω1τs(sn)\Omega_{1}\in\tau_{s}(\mathbb{H}_{s}^{n}) be real-path-connected, Ω2τs(sn)\Omega_{2}\in\tau_{s}(\mathbb{H}_{s}^{n}) be Ω1\Omega_{1}-stem-preserving, f𝒫𝒮(Ω1)f\in\mathcal{PS}(\Omega_{1}) and g𝒫𝒮(Ω2)g\in\mathcal{PS}(\Omega_{2}). Then, the following relationship holds:

(4.3) fg(zI)=(f(zI),If(zI))(FΩ1gγ)(z),f*g\left(z^{I}\right)=\left(f\left(z^{I}\right),If\left(z^{I}\right)\right)\left(F_{\Omega_{1}}^{g}\circ\mathscr{L}_{\gamma}\right)(z),

for every path γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), each element I𝕊(Ω1,γ)I\in\mathbb{S}(\Omega_{1},\gamma), any radius r(0,rγ,Ω1I)r\in\left(0,r_{\gamma,\Omega_{1}}^{I}\right) and for all points zB(γ(1),r)z\in B_{\mathbb{C}}\left(\gamma(1),r\right).

Proof.

Consider γ\gamma as a path in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), II as an element of 𝕊(Ω1,γ)\mathbb{S}(\Omega_{1},\gamma), rr within the range (0,rγ,Ω1I)\left(0,r_{\gamma,\Omega_{1}}^{I}\right), and let zz be a point in the complex ball B(γ(1),r)B_{\mathbb{C}}\left(\gamma(1),r\right). From equation (4.1), we have γ(z)\mathscr{L}_{\gamma}(z) residing in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}). Referring to equations (2.6) and (2.7), the computation proceeds as follows:

fg(zI)=(f(zI),(zI)f(zI))Ω1g(zI)=(f(zI),If(zI))FΩ1g(γγ(1)z)=(f(zI),If(zI))FΩ1g(γ(z))=(f(zI),If(zI))(FΩ1gγ)(z).\begin{split}f*g(z^{I})=&\left(f(z^{I}),\mathfrak{I}(z^{I})f(z^{I})\right)\mathscr{F}_{\Omega_{1}}^{g}(z^{I})\\ =&\left(f(z^{I}),If(z^{I})\right)F_{\Omega_{1}}^{g}\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right)\\ =&\left(f(z^{I}),If(z^{I})\right)F_{\Omega_{1}}^{g}\left(\mathscr{L}_{\gamma}(z)\right)\\ =&\left(f\left(z^{I}\right),If\left(z^{I}\right)\right)\left(F_{\Omega_{1}}^{g}\circ\mathscr{L}_{\gamma}\right)(z).\end{split}

This completes the proof. ∎

Let cc\in\mathbb{H} and I𝕊I\in\mathbb{S}. Then

(4.4) I(c,Ic)=(Ic,c)=(c,Ic)(11)=(c,Ic)σ.I(c,Ic)=(Ic,-c)=(c,Ic)\begin{pmatrix}&-1\\ 1\end{pmatrix}=(c,Ic)\sigma.
Theorem 4.3.

Let Ω1τs(sn)\Omega_{1}\in\tau_{s}(\mathbb{H}_{s}^{n}) be real-path-connected, Ω2τs(sn)\Omega_{2}\in\tau_{s}(\mathbb{H}_{s}^{n}) be Ω1\Omega_{1}-stem-preserving, f𝒮(Ω1)f\in\mathcal{SR}(\Omega_{1}) and g𝒮(Ω2)g\in\mathcal{SR}(\Omega_{2}). Then fg𝒮(Ω1)f*g\in\mathcal{SR}(\Omega_{1}).

Proof.

Consider II as an element of 𝕊\mathbb{S} and let qq be a point in (Ω1)I\left(\Omega_{1}\right)_{I}. There exists a complex number zz such that qq can be represented as zIz^{I}. Given that Ω1\Omega_{1} is real-path-connected, we can find a path γ\gamma in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) with II in 𝕊(Ω1,γ)\mathbb{S}(\Omega_{1},\gamma), such that γI\gamma^{I} lies in Ω1\Omega_{1} and qq equals γI(1)\gamma^{I}(1). Referring to equations (4.3), (4.4), and Proposition 3.6, the following derivation is obtained:

12(x+Iy)(fg)(zI)\displaystyle\frac{1}{2}\left(\frac{\partial}{\partial x\ell}+I\frac{\partial}{\partial y_{\ell}}\right)(fg)(z^{I})
=\displaystyle= 12(x+Iy)(f(zI),If(zI))(FΩ1gγ)(z)\displaystyle\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)\left(f\left(z^{I}\right),If\left(z^{I}\right)\right)\left(F_{\Omega_{1}}^{g}\circ\mathscr{L}\gamma\right)(z)
=\displaystyle= (12(x+Iy)f(zI),I[12(x+Iy)]f(zI))(FΩ1gγ)(z)\displaystyle\left(\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)f\left(z^{I}\right),I\left[\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)\right]f\left(z^{I}\right)\right)\left(F_{\Omega_{1}}^{g}\circ\mathscr{L}\gamma\right)(z)
+(f(zI),If(zI))[12(x+σy)](FΩ1gγ)(z)\displaystyle+\left(f\left(z^{I}\right),If\left(z^{I}\right)\right)\left[\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\right]\left(F_{\Omega_{1}}^{g}\circ\mathscr{L}\gamma\right)(z)
=\displaystyle= (0,I0)(FΩ1gγ)(z)+(f(zI),If(zI))0=0.\displaystyle(0,I\cdot 0)\left(F_{\Omega_{1}}^{g}\circ\mathscr{L}_{\gamma}\right)(z)+\left(f\left(z^{I}\right),If\left(z^{I}\right)\right)\cdot 0=0.

This result implies that (fg)I(fg)_{I} is holomorphic for every II in 𝕊\mathbb{S}. Consequently, it is established that fgfg is slice regular. ∎

Theorem 4.4.

Suppose Ω\Omega is a subset of sn\mathbb{H}_{s}^{n} that exhibits self-stem-preserving properties. In such a case, the structure (𝒮(Ω),+,)(\mathcal{SR}(\Omega),+,*) forms an associative unitary real algebra.

Proof.

The assertion of this theorem follows immediately from the application of Theorem 2.11 in conjunction with Theorem 4.3. ∎

In conclusion, our study has established that the *-product preserves the slice regularity of functions, thereby confirming that this class of functions forms an associative unitary real algebra. Our investigation into the star-product of slice regular functions over several quaternionic variables has enriched our comprehension of their algebraic and geometric properties. Moreover, it has opened avenues for future research and potential applications.

References