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

Zeroes of weakly slice regular functions of several quaternionic variables on non-axially symmetric domains

Xinyuan Dou [email protected] 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 [email protected] Faculty of Innovation Engineering, Macau University of Science and Technology, Macau, China Guangbin Ren [email protected] Department of Mathematics, University of Science and Technology of China, Hefei 230026, China  and  Ting Yang [email protected] School of Mathematics and Physics, Anqing Normal University, Anqing 246133, China
Abstract.

In this research, we study zeroes of weakly slice regular functions within the framework of several quaternionic variables, specifically focusing on non-axially symmetric domains. Our recent work introduces path-slice stem functions, along with a novel *-product, tailored for weakly slice regular functions. This innovation allows us to explore new techniques for conjugating and symmetrizing path-slice functions. A key finding of our study is the discovery that the zeroes of a path-slice function are comprehensively encapsulated within the zeroes of its symmetrized counterpart. This insight is particularly significant in the context of path-slice stem functions. We establish that for weakly slice regular functions, the processes of conjugation and symmetrization gain prominence once the function’s slice regularity is affirmed. Furthermore, our investigation sheds light on the intricate nature of the zeroes of a slice regular function. We ascertain that these zeroes constitute a path-slice analytic set. This conclusion is drawn from the observed phenomenon that the zeroes of the symmetrization of a slice regular function also form a path-slice analytic set. This finding marks an advancement in understanding the complex structure and properties of weakly slice regular functions in quaternionic analysis.

Key words and phrases:
Quaternions; slice regular functions; zero set; symmetrization; conjugation
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

Quaternions, conceptualized by Hamilton in 1843, represent an extension of complex numbers, made possible by the Cayley-Dickson construction as noted in Dickson’s work [Dickson1919001]. This development led to a unique form of quaternionic analysis, aiming to broaden the scope of holomorphic function theory into quaternionic variable domains. One significant branch of quaternionic analysis is slice analysis over quaternions, introduced by Gentili and Struppa [Gentili2007001], which bases itself on the idea of representing the quaternionic field \mathbb{H} as a collective of complex planes.

The slice regular functions, central to this theory, are functions that conform to the Cauchy-Riemann equations across these complex planes. This categorizes them as vector-valued holomorphic functions when analyzed within these planes. Notably, while simple functions like identities and polynomials qualify as slice regular functions, this classification does not extend to Fueter-regular functions, another quaternionic analysis model predating slice analysis, as discussed in Fueter’s work [Fueter1934001].

The growth and development of slice analysis have led to its integration into several mathematical disciplines, including geometric function theory [Ren2017001, Ren2017002, Wang2017001], quaternionic Schur analysis [Alpay2012001], and quaternionic operator theory [Alpay2015001, MR3887616, MR3967697, Gantner2020001, MR4496722]. Its expansion further encompasses higher dimensions through real Clifford algebras [Colombo2009002], octonions [Gentili2010001], real alternative *-algebras [Ghiloni2011001], and 2n2n-dimensional Euclidean spaces [Dou2023002]. Additionally, the study of slice analysis extends to several variables across various frameworks [Colombo2012002, Dou2023002, Ghiloni2012001, Ghiloni2020001], presenting it as an evolution of complex analysis in several variables.

Two distinct forms of slice regular functions have emerged within this field. The first, the weakly slice regular functions, were introduced by Gentili and Struppa [Gentili2007001] and initially focused on Euclidean open sets, with the representation formula playing a pivotal role in their study [Colombo2009001]. The second, the strongly regular functions, were later introduced by Ghiloni and Perotti [Ghiloni2011001] to extend the concept to quadratic cones in real alternative *-algebras. This extension brought about the concept of stem functions, intrinsic to the structure of slice functions and pivotal for holomorphy and multiplication in slice analysis.

One of recent advancements in slice analysis is the slice topology [Dou2023001], a nuanced approach that transcends the limitations of the Euclidean topology and facilitates the study of slice analysis in non-axially symmetric domains. This has led to the emergence of path-slice functions and a deeper exploration of the convergence domains of quaternionic power series.

The primary objective of this paper is to investigate the zero sets of weakly slice regular functions in several quaternionic variables, particularly within non-axially symmetric slice-domains. While earlier studies [Gentili2008001, MR3026135, MR4182982] have focused on Euclidean domains, this paper aims to understand the properties of zeros on domains in slice topology. Central to this study is the concept of symmetrization of path-slice and subsequently slice regular functions. By examining the conjugation and symmetrization processes [Colombo2009001], and ensuring the preservation of slice regularity and slice-preserving properties, this paper aims to provide a thorough understanding of the zeros of these functions and their analytical nature in quaternionic variable domains.

This paper unfolds as follows: Section 2 serves as a foundation, where we delve into the fundamentals of weak slice regular functions and their associated stem functions within the context of multiple quaternionic variables. In Section 3, we introduce the concept of Ω1\Omega_{1}-slice conjugation for path-slice functions. We demonstrate that this conjugation maintains slice regularity when derived from a slice regular function. Section 4 is dedicated to the exploration of symmetrization within the realm of path-slice functions. Here, we establish that this symmetrization retains the critical attribute of being slice-preserving. Finally, Section 5 focuses on the zeros of a path-slice function. By juxtaposing these zeros with those of their symmetrization, we reveal that the zero set of a slice regular function forms an analytic set, thereby providing a deeper understanding of its structural properties.

2. Preliminaries

This section lays the groundwork for our analysis, building upon foundational concepts and results as detailed in [Dou2023003, Dou2023004].

Denote by \mathbb{H} the algebra of quaternions. Define

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

where

𝕊:={I:I2=1},andIn:=(I)n.\mathbb{S}:=\{I\in\mathbb{H}:I^{2}=-1\},\qquad\mbox{and}\qquad\mathbb{C}_{I}^{n}:=(\mathbb{C}_{I})^{n}.

The slice topology is defined by

τs(sn):={Usn:UIτ(In)}\tau_{s}(\mathbb{H}_{s}^{n}):=\{U\in\mathbb{H}_{s}^{n}:U_{I}\in\tau(\mathbb{C}_{I}^{n})\}

where τ(In)\tau(\mathbb{C}_{I}^{n}) is the Euclidean topology in nn-dimensional complex plane In\mathbb{C}_{I}^{n} and

UI:=UI.U_{I}:=U\cap\mathbb{C}_{I}.

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

Extending the classical complex space n\mathbb{C}^{n} with quaternionic imaginary units II from 𝕊\mathbb{S} leads to the space In\mathbb{C}_{I}^{n}. The transition employs the mapping ΨiI:n\xlongrightarrowIn\Psi_{i}^{I}:\mathbb{C}^{n}\xlongrightarrow{}\mathbb{C}_{I}^{n}, defined by

ΨiI(x+yi)=x+yI,\Psi_{i}^{I}(x+yi)=x+yI,

for real vectors x,ynx,y\in\mathbb{R}^{n}, replacing the standard imaginary unit ii with II.

To understand the structure of sn\mathbb{H}_{s}^{n} and the behavior of related functions, we define a set of paths 𝒫(n)\mathscr{P}(\mathbb{C}^{n}) as continuous paths γ\gamma from [0,1][0,1] to n\mathbb{C}^{n}, starting in the real subspace n\mathbb{R}^{n} at γ(0)\gamma(0). For a subset Ω\Omega of sn\mathbb{H}_{s}^{n}, we define 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega) as the paths δ\delta in 𝒫(n)\mathscr{P}(\mathbb{C}^{n}) for which there exists I𝕊I\in\mathbb{S} such that the path in sn\mathbb{H}_{s}^{n}

δI:=ΨiI(δ)\delta^{I}:=\Psi_{i}^{I}(\delta)

lies in Ω\Omega.

Furthermore, for a fixed path γ\gamma in 𝒫(n)\mathscr{P}(\mathbb{C}^{n}), we define the set 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma) as

𝕊(Ω,γ):={I𝕊γIΩ}.\mathbb{S}(\Omega,\gamma):=\left\{I\in\mathbb{S}\mid\gamma^{I}\subset\Omega\right\}.

This set contains all quaternionic imaginary units II from 𝕊\mathbb{S} for which the path γ\gamma, when transformed to the slice In\mathbb{C}_{I}^{n} of sn\mathbb{H}_{s}^{n}, is entirely contained within the subset Ω\Omega.

Path-slice functions and their stem functions are central to our study.

Definition 2.1.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n}. A function f:Ωf:\Omega\rightarrow\mathbb{H} is termed path-slice if there exists a function

F:𝒫(n,Ω)2×1F:\mathscr{P}(\mathbb{C}^{n},\Omega)\rightarrow\mathbb{H}^{2\times 1}

satisfying

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

for any 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 referred to as a path-slice stem function of ff. The set of all path-slice functions defined on Ω\Omega is denoted by 𝒫𝒮(Ω)\mathcal{PS}(\Omega), and the set of all path-slice stem functions of a function f𝒫𝒮(Ω)f\in\mathcal{PS}(\Omega) is denoted by 𝒫𝒮𝒮(f)\mathcal{PSS}(f).

Exploring the complex structures of slice quaternionic analysis leads us to the concept of slice regularity.

Definition 2.2.

For a subset Ω\Omega in the slice topology τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}), a function f:Ωf:\Omega\rightarrow\mathbb{H} is called (weakly) slice regular if and only if, for each I𝕊I\in\mathbb{S}, the restriction

fI:=f|ΩIf_{I}:=f|_{\Omega_{I}}

is (left II-)holomorphic. This means that fIf_{I} is real differentiable and, for each =1,2,,d\ell=1,2,\ldots,d,

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

The set of all weakly slice regular functions defined on Ω\Omega is denoted by 𝒮(Ω)\mathcal{SR}(\Omega).

By [Dou2023002, Corollary 5.9], weakly slice regular functions are path-slice.

In our exploration of specific subsets within the quaternionic cone sn\mathbb{H}_{s}^{n}, it becomes necessary to introduce and define certain key concepts.

Definition 2.3.

A subset Ωsn\Omega\subset\mathbb{H}_{s}^{n} is called real-path-connected if, for each point qΩq\in\Omega, there exists a path γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) and an imaginary unit I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma) such that γI(1)=q\gamma^{I}(1)=q.

We introduce the set 𝒫2(n,Ω)\mathscr{P}_{*}^{2}(\mathbb{C}^{n},\Omega), which comprises pairs of continuous paths (α,β)(\alpha,\beta) within 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega) and sharing the same endpoint at α(1)=β(1)\alpha(1)=\beta(1). Formally, it is represented as

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

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

  1. (i)

    The cardinality of the set 𝕊(Ω2,γ)\mathbb{S}(\Omega_{2},\gamma) is at least 2 for each path γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}).

  2. (ii)

    The intersection 𝕊(Ω2,α)𝕊(Ω2,β)\mathbb{S}(\Omega_{2},\alpha)\cap\mathbb{S}(\Omega_{2},\beta) does not contain exactly one element for each pair (α,β)(\alpha,\beta) in 𝒫2(n,Ω1)\mathscr{P}_{*}^{2}(\mathbb{C}^{n},\Omega_{1}).

Definition 2.5.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving, and f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} be path-slice. Then

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

is well defined and does not depend on the choice of G𝒫𝒮𝒮(f)G\in\mathcal{PSS}(f).

This function FΩ1fF_{\Omega_{1}}^{f} is well-defined and its formulation is independent of the particular choice of the stem function FF within the set 𝒫𝒮𝒮(f)\mathcal{PSS}(f), as established in [Dou2023003, Proposition 3.9].

Lemma 2.6.

[Dou2023003, Proposition 3.3]. Let Ωsn\Omega\subset\mathbb{H}_{s}^{n}, f𝒫𝒮(Ω)f\in\mathcal{PS}(\Omega), γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega), FF be a path-slice stem function of ff, and I,J𝕊(Ω,γ)I,J\in\mathbb{S}(\Omega,\gamma) with IJI\neq J. Then

(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}.

For any quaternion q=(q1,,qn)q=(q_{1},\ldots,q_{n}) within the nn-dimensional weakly slice cone sn\mathbb{H}_{s}^{n}, there exists a specific imaginary unit II from the set 𝕊\mathbb{S} such that each component qiq_{i} of qq belongs to the corresponding complex plane I\mathbb{C}_{I}. This particular imaginary unit II is denoted as (q)\mathfrak{I}(q). To define (q)\mathfrak{I}(q) precisely, consider the function:

:sn𝕊{0}\begin{split}\mathfrak{I}:\mathbb{H}_{s}^{n}&\longrightarrow\mathbb{S}\cup\{0\}\end{split}

defined by

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

where ı\imath is the smallest positive integer from {1,,n}\{1,\ldots,n\} such that the component qıq_{\imath} of qq is not a real number. This formulation assigns to each quaternion in sn\mathbb{H}_{s}^{n} an appropriate imaginary unit from 𝕊\mathbb{S} or the value 0 if the quaternion lies entirely in the real space n\mathbb{R}^{n}.

By [Dou2023003, Proposition 3.11], the following definition is well-defined.

Definition 2.7.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving, and f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} be path-slice. Then the function

(2.5) Ω1f:Ω1\xlongrightarrow[]2×1\begin{split}\mathscr{F}_{\Omega_{1}}^{f}:\quad\Omega_{1}\quad&\xlongrightarrow[\hskip 28.45274pt]{}\quad\mathbb{H}^{2\times 1}\end{split}

is defined by

(2.6) Ω1f(q)={FΩ1f(γ),qn,(f(q),0)T,otherwise,\begin{split}\mathscr{F}_{\Omega_{1}}^{f}(q)=\begin{cases}F_{\Omega_{1}}^{f}(\gamma),\qquad&q\notin\mathbb{R}^{n},\\ \left(f(q),0\right)^{T},\qquad&\mbox{otherwise},\end{cases}\end{split}

for any γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) with γ(q)Ω1\gamma^{\mathfrak{I}(q)}\subset\Omega_{1} and γ(q)(1)=q\gamma^{\mathfrak{I}(q)}(1)=q.

Now we introduce the *-product and its properties, a fundamental operation that underpins the algebraic structure of the quaternionic function space. This operation not only enriches the algebraic framework but also provides a tool for examining the interactions between quaternionic functions in more complex scenarios.

Definition 2.8.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\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}). We call

(2.7) fg:=(f,f)Ω1g:Ω1f*g:=(f,\mathfrak{I}f)\mathscr{F}_{\Omega_{1}}^{g}:\Omega_{1}\rightarrow\mathbb{H}

the *-product of ff and gg.

In the context of quaternionic analysis, the standard product in \mathbb{H}\otimes_{\mathbb{R}}\mathbb{C} can be extended to a unique product within the space of 2×12\times 1 quaternionic column vectors, 2×1\mathbb{H}^{2\times 1}. This specialized product is referred to as the *-product. To define this product more concretely, consider two vectors p=(p1,p2)Tp=(p_{1},p_{2})^{T} and q=(q1,q2)Tq=(q_{1},q_{2})^{T} in 2×1\mathbb{H}^{2\times 1}. The pointwise *-product of these vectors is given by the formula:

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

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

𝕀:=(11),σ:=(11),e1:=(10).\mathbb{I}:=\begin{pmatrix}1\\ &1\end{pmatrix},\quad\sigma:=\begin{pmatrix}&-1\\ 1\end{pmatrix},\quad e_{1}:=\begin{pmatrix}1\\ 0\end{pmatrix}.

Extending this concept to functional spaces, let Ω\Omega be a subset of sn\mathbb{H}_{s}^{n}, and consider two functions F,G:𝒫(n,Ω)2×1F,G:\mathscr{P}(\mathbb{C}^{n},\Omega)\rightarrow\mathbb{H}^{2\times 1}. The functional *-product of FF and GG is then defined as a mapping from the set of paths 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega) to 2×1\mathbb{H}^{2\times 1}, where each path γ\gamma is mapped to the pointwise *-product of F(γ)F(\gamma) and G(γ)G(\gamma):

(2.9) FG:𝒫(n,Ω)\displaystyle F*G:\mathscr{P}(\mathbb{C}^{n},\Omega) \displaystyle\rightarrow 2×1,\displaystyle\mathbb{H}^{2\times 1},
γ\displaystyle\gamma \displaystyle\mapsto F(γ)G(γ).\displaystyle F(\gamma)*G(\gamma).
Proposition 2.9.

[Dou2023003, Proposition 4.5]. Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\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 fg𝒫𝒮(Ω1)f*g\in\mathcal{PS}(\Omega_{1}). Moreover, if FF is a path-slice stem function of ff, then FFΩ1gF*F_{\Omega_{1}}^{g} is a path-slice stem function of fgf*g.

We will illustrate that within non-axially symmetric domains, the roles typically filled by axially symmetric sets can be successfully taken over by those classified as self-stem-preserving domains.

Definition 2.10.

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

Lemma 2.11.

[Dou2023003, Lemma 3.12]. Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving, and f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} be path-slice. Then

(2.10) Ω1f|(Ω1)=(f|(Ω1)0).\left.\mathscr{F}_{\Omega_{1}}^{f}\right|_{(\Omega_{1})_{\mathbb{R}}}=\begin{pmatrix}f|_{(\Omega_{1})_{\mathbb{R}}}\\ 0\end{pmatrix}.
Lemma 2.12.

[Dou2023003, Lemma 3.14]. 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. If f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} is path-slice and γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) with γ(1)n\gamma(1)\in\mathbb{R}^{n}, then

(2.11) FΩ1f(γ)=(fγ(1)0).F_{\Omega_{1}}^{f}(\gamma)=\begin{pmatrix}f\circ\gamma(1)\\ 0\end{pmatrix}.
Lemma 2.13.

[Dou2023003, Lemma 3.13]. Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving, and f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} be path-slice. Then

(2.12) FΩ1f(γ)=(11)FΩ1f(γ¯),γ𝒫(n,Ω1).F_{\Omega_{1}}^{f}(\gamma)=\begin{pmatrix}1\\ &-1\end{pmatrix}F_{\Omega_{1}}^{f}(\overline{\gamma}),\qquad\forall\ \gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}).

Consider complex numbers z,wz,w\in\mathbb{C} and denote zw\mathcal{L}_{z}^{w} as the linear segment connecting zz to ww. Given a path γ\gamma in the set 𝒫(n)\mathscr{P}(\mathbb{C}^{n}) and a positive radius rr, we define the set B𝒫(n)(γ,r)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r) as follows:

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

Here, B(γ(1),r)B_{\mathbb{C}}(\gamma(1),r) represents the ball in the complex plane \mathbb{C} centered at γ(1)\gamma(1) with a radius of rr. This set essentially forms a collection of paths derived from γ\gamma by extending its endpoint γ(1)\gamma(1) along all possible directions within the radius rr in the complex plane. For simplicity, we make the convention: when r=0r=0, we set B𝒫(n)(γ,r)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r) as the empty set ϕ\phi.

Define for γ𝒫(n)\gamma\in\mathscr{P}(\mathbb{C}^{n}) and r>0r>0,

(2.14) γ:B(γ(1),r)\xlongrightarrow[]B𝒫(n)(γ,r),z\xlongrightarrow[]γγ(1)z.\begin{split}\mathscr{L}_{\gamma}\quad:\quad B_{\mathbb{C}}(\gamma(1),r)\quad&\xlongrightarrow[\hskip 28.45274pt]{}\quad B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r),\\ z\qquad\ &\shortmid\!\xlongrightarrow[\hskip 28.45274pt]{}\ \quad\gamma\circ\mathcal{L}_{\gamma(1)}^{z}.\end{split}

In our study, we introduce several constants that are crucial for understanding the behavior of functions within quaternionic spaces. Let’s consider a subset Ω\Omega of the nn-dimensional weakly slice cone sn\mathbb{H}_{s}^{n}, a path γ\gamma in 𝒫(n)\mathscr{P}(\mathbb{C}^{n}), and a subset 𝕊\mathbb{S}^{\prime} of the set 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma), which includes specific quaternionic imaginary units. We define the following constants:

  • rγ,Ωr_{\gamma,\Omega} is defined as the supremum of all radii rr such that the set of paths B𝒫(n)(γ,r)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r) is entirely contained within 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega):

    rγ,Ω:=sup{r[0,):B𝒫(n)(γ,r)𝒫(n,Ω)}.r_{\gamma,\Omega}:=\sup\left\{r\in[0,\infty):B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r)\subset\mathscr{P}(\mathbb{C}^{n},\Omega)\right\}.
  • rγ,Ω𝕊r_{\gamma,\Omega}^{\mathbb{S}^{\prime}} is defined as the supremum of all radii rr such that, for every imaginary unit II in 𝕊\mathbb{S}^{\prime}, the ball BI(γI(1),r)B_{I}(\gamma^{I}(1),r) is contained within the slice ΩI\Omega_{I}:

    rγ,Ω𝕊:=sup{r[0,):BI(γI(1),r)ΩI,I𝕊}.r_{\gamma,\Omega}^{\mathbb{S}^{\prime}}:=\sup\left\{r\in[0,\infty):B_{I}(\gamma^{I}(1),r)\subset\Omega_{I},\forall I\in\mathbb{S}^{\prime}\right\}.
  • rγ,Ω2r_{\gamma,\Omega}^{2} is defined as the supremum among all rγ,Ω𝕊′′r_{\gamma,\Omega}^{\mathbb{S}^{\prime\prime}}, where 𝕊′′\mathbb{S}^{\prime\prime} is a subset of 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma) containing at least two elements:

    rγ,Ω2:=sup{rγ,Ω𝕊′′:𝕊′′𝕊(Ω,γ) with |𝕊′′|2}.r_{\gamma,\Omega}^{2}:=\sup\left\{r_{\gamma,\Omega}^{\mathbb{S}^{\prime\prime}}:\mathbb{S}^{\prime\prime}\subset\mathbb{S}(\Omega,\gamma)\text{ with }|\mathbb{S}^{\prime\prime}|\geqslant 2\right\}.

In these definitions, BI(γI(1),r)B_{I}(\gamma^{I}(1),r) represents the set of all points qq in the complex plane I\mathbb{C}_{I} that are within a distance rr from the point γI(1)\gamma^{I}(1):

BI(γI(1),r):={qI:|qγI(1)|<r}.B_{I}(\gamma^{I}(1),r):=\{q\in\mathbb{C}_{I}:|q-\gamma^{I}(1)|<r\}.

These constants play a pivotal role in determining the extent to which paths in complex space can be extended while remaining within the quaternionic subset Ω\Omega.

We are expanding the idea of holomorphy of stem functions defined in complex plane n\mathbb{C}^{n} to encompass the one in path spaces.

Definition 2.14.

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 exists 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

(2.15) 12(x+σy)(Fγ)(x+yi)=0\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)(F\circ\mathscr{L}_{\gamma})(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,\ldots,n\}. Furthermore, FF is termed holomorphic in U𝒫(n,Ω)U\subset\mathscr{P}(\mathbb{C}^{n},\Omega) if it is holomorphic at each γU\gamma\in U, and holomorphic in 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega) if it is holomorphic at every γ\gamma.

We revisit some key properties of slice regular functions.

Proposition 2.15.

[Dou2023004, 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.

Theorem 2.16.

[Dou2023004, Theorem 4.4]. Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} be self-stem-preserving. Then (𝒮(Ω),+,)(\mathcal{SR}(\Omega),+,*) forms an associative unitary real algebra.

3. Slice conjugation of path-slice functions

In this section, we delve into the intriguing concept of Ω1\Omega_{1}-slice conjugation for path-slice functions. This exploration is driven by the quest to identify an effective form of ‘conjugation’ for a given slice regular function within the confines of specific domain restrictions. While conventional notions of conjugation may not always be applicable, we discover that under certain conditions, a unique slice regular function can embody this role of conjugation. This function, termed the path-slice conjugation of the original function, emerges as a pivotal element in our analysis. Furthermore, we extend this concept to encompass path-slice functions as well, broadening the scope of our study.

This innovative approach to conjugation of slice regular functions not only enhances our understanding of their structure but also paves the way for further explorations into the properties and applications of these functions within slice quaternionic analysis.

Definition 3.1.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving, and f𝒫𝒮(Ω2)f\in\mathcal{PS}(\Omega_{2}). Then we call

(3.1) fΩ1c:=(1,)Ω1f,cf_{\scriptscriptstyle{\Omega_{1}}}^{c}:=(1,\mathfrak{I})\mathscr{F}_{\Omega_{1}}^{f,c}

the Ω1\Omega_{1}-slice conjugation of ff on Ω1\Omega_{1}, where

Ω1f,c:=(Ω1f,c,1Ω1f,c,2):=ConjΩ1f=(Ω1f,1¯Ω1f,2¯).\mathscr{F}_{\Omega_{1}}^{f,c}:=\begin{pmatrix}\mathscr{F}_{\Omega_{1}}^{f,c,1}\\ \\ \mathscr{F}_{\Omega_{1}}^{f,c,2}\end{pmatrix}:={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ\mathscr{F}_{\Omega_{1}}^{f}=\begin{pmatrix}\overline{\mathscr{F}_{\Omega_{1}}^{f,1}}\\ \\ \overline{\mathscr{F}_{\Omega_{1}}^{f,2}}\end{pmatrix}.
Remark 3.2.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be real-path-connected, Ω2sn\Omega_{2}\subset\mathbb{H}_{s}^{n} be Ω1\Omega_{1}-stem-preserving, and f𝒫𝒮(Ω2)f\in\mathcal{PS}(\Omega_{2}). By (5.7), we have

fΩ1c=f¯f^{c}_{\scriptscriptstyle{\Omega_{1}}}=\overline{f}

as the conjugation of slice regular functions on axially symmetric domains. In fact, if Ω2=σ(I,2)\Omega_{2}=\sigma(I,2) and

f=ı(qI2)2ı,f=\sum_{\imath\in\mathbb{N}}\left(\frac{{q-I}}{2}\right)^{*2^{\imath}},

then there is no slice regular function gg on Ω2\Omega_{2} such that

(3.2) g=f¯on(Ω2).g=\overline{f}\qquad\mbox{on}\qquad(\Omega_{2})_{\mathbb{R}}.

However, if let

Ω1:=Σ(I,2)Σ(I,2),\Omega_{1}:=\Sigma(I,2)\cap\Sigma(-I,2),

then Ω1\Omega_{1} is real-path-connected, Ω2\Omega_{2} is Ω1\Omega_{1}-stem-preserving and fΩ1cf^{c}_{\scriptscriptstyle{\Omega_{1}}} is a slice regular function (see Theorem 3.8) with fΩ1c=f¯f^{c}_{\scriptscriptstyle{\Omega_{1}}}=\overline{f} on (Ω1)=(Ω2)(\Omega_{1})_{\mathbb{R}}=(\Omega_{2})_{\mathbb{R}}.

To show that conjugation maintains the holomorphic nature of functions on path spaces, it is necessary to refer to a few important lemmas.

Lemma 3.3.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} be real-path-connected, I𝕊I\in\mathbb{S} and zz\in\mathbb{C} with zIΩz^{I}\in\Omega. Then there is γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) such that

(3.3) γIΩ,andγI(1)=zI.\gamma^{I}\subset\Omega,\qquad\mbox{and}\qquad\gamma^{I}(1)=z^{I}.
Proof.

Since Ω\Omega is real-path-connected, there is β𝒫(n,Ω)\beta\in\mathscr{P}(\mathbb{C}^{n},\Omega) and J𝕊J\in\mathbb{S} such that βJΩ\beta^{J}\subset\Omega and βJ(1)=zI\beta^{J}(1)=z^{I}.

If J±IJ\neq\pm I, then zIJnIn=nz^{I}\in\mathbb{C}_{J}^{n}\cap\mathbb{C}_{I}^{n}=\mathbb{R}^{n}. Let

γ:[0,1]\xlongrightarrow[]n,t\xlongrightarrow[]zI.\begin{split}\gamma\quad:\quad[0,1]\quad&\xlongrightarrow[\hskip 28.45274pt]{}\quad\mathbb{C}^{n},\\ t\quad&\shortmid\!\xlongrightarrow[\hskip 28.45274pt]{}\quad z^{I}.\end{split}

It is easy to check that (3.3) holds.

Otherwise, J=±IJ=\pm I. Let

γ:={β¯,JI,β,otherwise.\gamma:=\begin{cases}\overline{\beta},&\qquad J\neq I,\\ \beta,&\qquad\mbox{otherwise}.\end{cases}

It follows from

γI={β¯J=βJ,JI,βI=βJ,otherwise,\gamma^{I}=\begin{cases}\overline{\beta}^{-J}=\beta^{J},&\qquad J\neq I,\\ \beta^{I}=\beta^{J},&\qquad\mbox{otherwise},\end{cases}

that γIΩ\gamma^{I}\subset\Omega and (3.3) holds. ∎

Lemma 3.4.

Let Ωτs(sn)\Omega\in\tau_{s}\left(\mathbb{H}_{s}^{n}\right), γ𝒫(n,γ)\gamma\in\mathscr{P}(\mathbb{C}^{n},\gamma) and 𝕊𝕊(Ω,γ)\mathbb{S}^{\prime}\subset\mathbb{S}(\Omega,\gamma) with |𝕊|<+|\mathbb{S}^{\prime}|<+\infty. Then rγ,Ω𝕊>0r_{\gamma,\Omega}^{\mathbb{S}^{\prime}}>0.

Proof.

Let I𝕊𝕊(Ω,γ)I\in\mathbb{S}^{\prime}\subset\mathbb{S}(\Omega,\gamma). Then γI(1)ΩI\gamma^{I}(1)\in\Omega_{I}. Since Ω\Omega is slice-open, ΩI\Omega_{I} is open in I\mathbb{C}_{I}. There is rI>0r^{I}>0 such that BI(γI(1),rI)ΩIB_{I}(\gamma^{I}(1),r^{I})\subset\Omega_{I}. Let r:=min{rI:I𝕊}r:=\min\{r^{I}:I\in\mathbb{S}^{\prime}\}. By definition, rγ,Ω𝕊>r>0r_{\gamma,\Omega}^{\mathbb{S}^{\prime}}>r>0. ∎

Lemma 3.5.

Consider any real 2×22\times 2 matrix M2×2M\in\mathbb{R}^{2\times 2}. The following relation holds when acting on 2×1\mathbb{H}^{2\times 1}:

(3.4) MConj=ConjM.M\cdot{\mathop{\mathrm{Conj}}}_{\mathbb{H}}={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ M.
Proof.

Consider a real 2×22\times 2 matrix

M=(abcd)M=\begin{pmatrix}a&b\\ c&d\end{pmatrix}

within the space of quaternions 2×22×2\mathbb{R}^{2\times 2}\subset\mathbb{H}^{2\times 2}, and let (p,q)T(p,q)^{T} be a column vector in 2×1\mathbb{H}^{2\times 1}. We analyze the action of MM in conjunction with quaternion conjugation on this vector:

MConj(pq)\displaystyle M\cdot{\mathop{\mathrm{Conj}}}_{\mathbb{H}}\begin{pmatrix}p\\ q\end{pmatrix} =(abcd)(p¯q¯)\displaystyle=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}\overline{p}\\ \overline{q}\end{pmatrix}
=(ap¯+bq¯cp¯+dq¯)\displaystyle=\begin{pmatrix}a\overline{p}+b\overline{q}\\ c\overline{p}+d\overline{q}\end{pmatrix}
=(ap+bq¯cp+dq¯)\displaystyle=\begin{pmatrix}\overline{ap+bq}\\ \overline{cp+dq}\end{pmatrix}
=Conj(ap+bqcp+dq)\displaystyle={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\begin{pmatrix}ap+bq\\ cp+dq\end{pmatrix}
=ConjM(pq).\displaystyle={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ M\begin{pmatrix}p\\ q\end{pmatrix}.

This completes the proof. ∎

Lemma 3.6.

Suppose we have subsets Ω1\Omega_{1} and Ω2\Omega_{2} of sn\mathbb{H}_{s}^{n}, where Ω1\Omega_{1} is real-path-connected and Ω2\Omega_{2} is Ω1\Omega_{1}-stem-preserving. Let cc be a quaternion in \mathbb{H}, ff a path-slice function on Ω2\Omega_{2}, and γ\gamma a continuous path in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}). For an imaginary unit I𝕊(Ω1,γ)I\in\mathbb{S}(\Omega_{1},\gamma) and a point q:=γI(1)q:=\gamma^{I}(1), the following relation holds:

(3.5) (c,Ic)FΩ1f,c(γ)=(c,Ic)FΩ1f,c(γ¯)=(c,(q)c)Ω1f,c(q),(c,Ic)F_{\Omega_{1}}^{f,c}(\gamma)=(c,-Ic)F_{\Omega_{1}}^{f,c}(\overline{\gamma})=\left(c,\mathfrak{I}(q)c\right)\mathscr{F}_{\Omega_{1}}^{f,c}(q),

where

FΩ1f,c:=ConjFΩ1f.F_{\Omega_{1}}^{f,c}:={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ F_{\Omega_{1}}^{f}.
Proof.

Let us consider a path γ\gamma in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) and an imaginary unit II in 𝕊(Ω1,γ)\mathbb{S}(\Omega_{1},\gamma). By (2.12) and (3.4),

FΩ1f,c(γ)=ConjFΩ1f(γ)=Conj[(11)FΩ1f(γ¯)]=(11)ConjFΩ1f(γ¯)=(11)FΩ1f,c(γ¯).\begin{split}F_{\Omega_{1}}^{f,c}(\gamma)&={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ F_{\Omega_{1}}^{f}(\gamma)\\ &={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\left[\begin{pmatrix}1\\ &-1\end{pmatrix}F_{\Omega_{1}}^{f}(\overline{\gamma})\right]\\ &=\begin{pmatrix}1\\ &-1\end{pmatrix}{\mathop{\mathrm{Conj}}}_{\mathbb{H}}F_{\Omega_{1}}^{f}(\overline{\gamma})\\ &=\begin{pmatrix}1\\ &-1\end{pmatrix}F_{\Omega_{1}}^{f,c}(\overline{\gamma}).\end{split}

This implies

(c,Ic)FΩ1f,c(γ)=(c,Ic)(11)FΩ1f,c(γ)=(c,Ic)FΩ1f,c(γ¯).(c,Ic)F_{\Omega_{1}}^{f,c}(\gamma)=(c,Ic)\begin{pmatrix}1\\ &-1\end{pmatrix}F_{\Omega_{1}}^{f,c}(\gamma)=(c,-Ic)F_{\Omega_{1}}^{f,c}(\overline{\gamma}).

(i) For any point qq not in n\mathbb{R}^{n}, (q)\mathfrak{I}(q) is either II or I-I. Depending on (q)\mathfrak{I}(q) being II or I-I, we have γ(q)=γI\gamma^{\mathfrak{I}(q)}=\gamma^{I} or γ¯(q)=γI\overline{\gamma}^{\mathfrak{I}(q)}=\gamma^{I}, respectively.

If (q)=I\mathfrak{I}(q)=I, then we have

Ω1f,c(q)=ConjΩ1f(q)=ConjFΩ1f(γ)=FΩ1f,c(γ)\mathscr{F}_{\Omega_{1}}^{f,c}(q)={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ\mathscr{F}_{\Omega_{1}}^{f}(q)={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ F_{\Omega_{1}}^{f}(\gamma)=F_{\Omega_{1}}^{f,c}(\gamma)

Otherwise we have (q)=I\mathfrak{I}(q)=-I so that

Ω1f,c(q)=ConjΩ1f(q)=ConjFΩ1f(γ¯)=FΩ1f,c(γ¯).\mathscr{F}_{\Omega_{1}}^{f,c}(q)={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ\mathscr{F}_{\Omega_{1}}^{f}(q)={\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ F_{\Omega_{1}}^{f}(\overline{\gamma})=F_{\Omega_{1}}^{f,c}(\overline{\gamma}).

These results validate equation (3.5) in this particular setting.

(ii) In the case where qq is in n\mathbb{R}^{n}, utilizing (2.10) and (2.11), we deduce:

(c,Ic)FΩ1f,c(γ)=(c,Ic)ConjFΩ1f(γ)=cConjf(γ(1))=cConjf(q)=(c,(q)c)ConjΩ1f(q)=Ω1f,c(q).\begin{split}(c,Ic)F_{\Omega_{1}}^{f,c}(\gamma)&=(c,Ic){\mathop{\mathrm{Conj}}}_{\mathbb{H}}F_{\Omega_{1}}^{f}(\gamma)\\ &=c{\mathop{\mathrm{Conj}}}_{\mathbb{H}}f(\gamma(1))\\ &=c{\mathop{\mathrm{Conj}}}_{\mathbb{H}}f(q)\\ &=(c,\mathfrak{I}(q)c){\mathop{\mathrm{Conj}}}_{\mathbb{H}}\mathscr{F}_{\Omega_{1}}^{f}(q)\\ &=\mathscr{F}_{\Omega_{1}}^{f,c}(q).\end{split}

Hence, equation (3.5) holds in this scenario as well. ∎

Lemma 3.7.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} and F:𝒫(n,Ω)F:\mathscr{P}(\mathbb{C}^{n},\Omega)\rightarrow\mathbb{H} be a holomorphic function. Then FcF^{c} is also holomorphic.

Proof.

Let γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega). According to Definition 2.14, 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

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+yiBn(γ(1),r)x+yi\in B_{\mathbb{C}^{n}}(\gamma(1),r) and {1,,n}\ell\in\{1,...,n\}. It implies that

12(x+σy)(Fcγ)(x+yi)=12(x+σy)(ConjFγ)(x+yi)=Conj[12(x+σy)](Fγ)(x+yi)=0\begin{split}&\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\left(F^{c}\circ\mathscr{L}_{\gamma}\right)(x+yi)\\ =&\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\left({\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ F\circ\mathscr{L}_{\gamma}\right)(x+yi)\\ =&{\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ\left[\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\right]\left(F\circ\mathscr{L}_{\gamma}\right)(x+yi)=0\end{split}

Therefore, FcF^{c} is holomorphic. ∎

Let I𝕊I\in\mathbb{S}. Then

(3.6) I(1,I)=(I,1)=(1,I)(11)=(1,I)σ.I(1,I)=(I,-1)=(1,I)\begin{pmatrix}&-1\\ 1\end{pmatrix}=(1,I)\sigma.

We are now prepared to demonstrate how conjugation retains the holomorphic characteristics of functions within path spaces.

Theorem 3.8.

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 fΩ1c𝒮(Ω1)f_{\scriptscriptstyle{\Omega_{1}}}^{c}\in\mathcal{SR}(\Omega_{1}).

Proof.

Let zIΩ1z^{I}\in\Omega_{1}. According to Lemma 3.3, there is γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma) such that γI(1)=zI\gamma^{I}(1)=z^{I}. According to Proposition 2.15, FΩ1fF_{\Omega_{1}}^{f} is holomorphic. By Lemma 3.7,

FΩ1f,c=(FΩ1f)cF_{\Omega_{1}}^{f,c}=\left(F_{\Omega_{1}}^{f}\right)^{c}

is also holomorphic. According to Definition 2.14, there is r>0r>0 such that

B𝒫(n)(γ,r)𝒫(n,Ω1)B_{\mathscr{P}(\mathbb{C}^{n})}(\gamma,r)\subset\mathscr{P}(\mathbb{C}^{n},\Omega_{1})

and

(3.7) 12(x+σy)(FΩ1f,cγ)(x+yi)=0,\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\left(F_{\Omega_{1}}^{f,c}\circ\mathscr{L}_{\gamma}\right)(x+yi)=0,

for each x+yiBn(γ(1),r)x+yi\in B_{\mathbb{C}^{n}}(\gamma(1),r) and {1,,n}\ell\in\{1,...,n\}. Taking c=1c=1 in (3.5), we have

fΩ1c(x+yI)=(1,(x+yI))Ω1f,c(x+yI)=(1,I)FΩ1f,c(γγ(1)x+yi),f_{\scriptscriptstyle{\Omega_{1}}}^{c}(x+yI)=(1,\mathfrak{I}(x+yI))\mathscr{F}_{\Omega_{1}}^{f,c}(x+yI)=(1,I)F_{\Omega_{1}}^{f,c}\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{x+yi}\right),

for each x+yiBn(γ(1),r)x+yi\in B_{\mathbb{C}^{n}}(\gamma(1),r). According to (3.6) and (3.7),

12(x+Iy)fΩ1c(x+yI)=12(x+Iy)(1,I)FΩ1f,c(γγ(1)x+yi)=(1,I)[12(x+σy)](FΩ1f,cγ)(x+yi)=0,\begin{split}&\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)f_{\scriptscriptstyle{\Omega_{1}}}^{c}(x+yI)\\ =&\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+I\frac{\partial}{\partial y_{\ell}}\right)(1,I)F_{\Omega_{1}}^{f,c}\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{x+yi}\right)\\ =&(1,I)\left[\frac{1}{2}\left(\frac{\partial}{\partial x_{\ell}}+\sigma\frac{\partial}{\partial y_{\ell}}\right)\right]\left(F_{\Omega_{1}}^{f,c}\circ\mathscr{L}_{\gamma}\right)(x+yi)=0,\end{split}

for each x+yiBn(γ(1),r)x+yi\in B_{\mathbb{C}^{n}}(\gamma(1),r). It implies that (fΩ1c)I\left(f_{\scriptscriptstyle{\Omega_{1}}}^{c}\right)_{I} is holomorphic at zIz^{I} for each zz\in\mathbb{C} with zIΩz^{I}\in\Omega. Therefore, (fΩ1c)I\left(f_{\scriptscriptstyle{\Omega_{1}}}^{c}\right)_{I} is holomorphic, for each I𝕊I\in\mathbb{S}. Hence fΩ1cf_{\scriptscriptstyle{\Omega_{1}}}^{c} is slice regular. ∎

We are also able to develop a path-slice stem function for the conjugate of a slice regular function.

Lemma 3.9.

Consider Ω1\Omega_{1} and Ω2\Omega_{2} within the slice topology τs(sn)\tau_{s}(\mathbb{H}_{s}^{n}), where Ω1\Omega_{1} is real-path-connected and Ω2\Omega_{2} is Ω1\Omega_{1}-stem-preserving. For a slice regular function ff defined on Ω2\Omega_{2}, the function FΩ1f,cF_{\Omega_{1}}^{f,c} is a path-slice stem function for fΩ1cf_{\scriptscriptstyle{\Omega_{1}}}^{c}.

Proof.

For any path γ\gamma within 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) and any imaginary unit II from 𝕊(Ω1,γ)\mathbb{S}(\Omega_{1},\gamma), let qq denote the endpoint of the path γI\gamma^{I}, that is, q=γI(1)q=\gamma^{I}(1). Utilizing equation (3.5) and the definition of Ω1\Omega_{1}-slice conjugation as stated in (3.1), we find that

(1,I)FΩ1f,c(γ)=(1,(q))Ω1f,c(q)=fΩ1c(q).(1,I)F_{\Omega_{1}}^{f,c}(\gamma)=(1,\mathfrak{I}(q))\mathscr{F}_{\Omega_{1}}^{f,c}(q)=f_{\scriptscriptstyle{\Omega_{1}}}^{c}(q).

Given this relationship and based on the definition of a path-slice function (Definition 2.1), we can conclude that FΩ1f,cF_{\Omega_{1}}^{f,c} acts as a path-slice stem function for the function fΩ1cf_{\scriptscriptstyle{\Omega_{1}}}^{c}. ∎

4. Symmetrization of path-slice functions

This section focuses on the innovative concept of extending symmetrization to path-slice functions. Our exploration reveals that symmetrization of a path-slice function not only retains its intrinsic properties but also ensures that it is slice-preserving. This property is crucial, as it maintains the function’s integrity across different slices of the quaternionic space. We delve into the intricate process of symmetrizing path-slice functions and demonstrate the implications of this process, particularly in preserving the slice nature of the functions. This advancement in quaternionic analysis opens new avenues for exploring the symmetrical aspects of path-slice functions.

The introduction of symmetrization in the context of path-slice functions represents a significant step forward in our understanding of the structural and functional dynamics of these functions. It not only maintains the essential characteristics of the original functions but also ensures their adaptability across various slices of quaternionic domains. This development is expected to contribute substantially to the field of slice quaternionic analysis, offering novel insights and methodologies for future research.

Definition 4.1.

For subsets Ω1\Omega_{1} and Ω2\Omega_{2} of sn\mathbb{H}s^{n}, where Ω1\Omega_{1} is real-path-connected and Ω2\Omega_{2} is Ω1\Omega_{1}-stem-preserving, let ff be a path-slice function defined on Ω2\Omega_{2}. The function

fΩ1s:=fΩ1cff^{s}_{\scriptscriptstyle{\Omega_{1}}}:=f^{c}_{\scriptscriptstyle{\Omega_{1}}}*f

is termed the symmetrization of ff on Ω1\Omega_{1}.

Definition 4.2.

For a subset Ω\Omega of sn\mathbb{H}_{s}^{n}, a function f:Ωf:\Omega\rightarrow\mathbb{H} is defined as slice-preserving if, for every imaginary unit II in 𝕊\mathbb{S}, the restriction of ff to the slice ΩI\Omega_{I}, denoted as fIf_{I}, maps ΩI\Omega_{I} into the complex plane I\mathbb{C}_{I}.

Theorem 4.3.

Given subsets Ω1\Omega_{1} and Ω2\Omega_{2} of the quaternion space sn\mathbb{H}_{s}^{n}, assume Ω1\Omega_{1} is real-path-connected and Ω2\Omega_{2} is Ω1\Omega_{1}-stem-preserving. Let f:Ω2f:\Omega_{2}\rightarrow\mathbb{H} be a path-slice function. Then the symmetrization fΩ1sf^{s}_{\Omega_{1}} is slice-preserving. Furthermore, the function

FΩ1f,s:=FΩ1f,cFΩ1fF_{\Omega_{1}}^{f,s}:=F_{\Omega_{1}}^{f,c}*F_{\Omega_{1}}^{f}

acts as a path-slice stem function for fΩ1sf^{s}_{\Omega_{1}}. It can be expressed explicitly as

(4.1) FΩ1f,s:=(FΩ1f,s,1FΩ1f,s,2):=(FΩ1f,1¯FΩ1f,1FΩ1f,2¯FΩ1f,2FΩ1f,2¯FΩ1f,1+FΩ1f,2¯FΩ1f,1¯)2×1,F_{\Omega_{1}}^{f,s}:=\begin{pmatrix}F_{\Omega_{1}}^{f,s,1}\\ \\ F_{\Omega_{1}}^{f,s,2}\end{pmatrix}:=\begin{pmatrix}\overline{F_{\Omega_{1}}^{f,1}}F_{\Omega_{1}}^{f,1}-\overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,2}\\ \\ \overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,1}+\overline{\overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,1}}\end{pmatrix}\in\mathbb{R}^{2\times 1},

where FΩ1fF_{\Omega_{1}}^{f} is defined as

FΩ1f=(FΩ1f,1FΩ1f,2).F_{\Omega_{1}}^{f}=\begin{pmatrix}F_{\Omega_{1}}^{f,1}\\ \\ F_{\Omega_{1}}^{f,2}\end{pmatrix}.
Proof.

First, consider the pointwise *-product as defined in equations (2.8) and (2.9). We apply these definitions to find the path-slice stem function of fΩ1sf^{s}_{\Omega_{1}}:

FΩ1f,s=FΩ1f,cFΩ1f=(FΩ1f,c,1𝕀+FΩ1f,c,2σ)(FΩ1f,1𝕀+FΩ1f,2σ)e1,=(FΩ1f,1¯𝕀+FΩ1f,2¯σ)(FΩ1f,1𝕀+FΩ1f,2σ)e1=[(FΩ1f,1¯FΩ1f,1FΩ1f,2¯FΩ1f,2)𝕀+(FΩ1f,2¯FΩ1f,1+FΩ1f,1¯FΩ1f,2)σ]e1=(FΩ1f,1¯FΩ1f,1FΩ1f,2¯FΩ1f,2FΩ1f,2¯FΩ1f,1+FΩ1f,2¯FΩ1f,1¯).\begin{split}F_{\Omega_{1}}^{f,s}=&F_{\Omega_{1}}^{f,c}*F_{\Omega_{1}}^{f}\\ =&\left(F_{\Omega_{1}}^{f,c,1}\cdot\mathbb{I}+F_{\Omega_{1}}^{f,c,2}\cdot\sigma\right)\left(F_{\Omega_{1}}^{f,1}\cdot\mathbb{I}+F_{\Omega_{1}}^{f,2}\cdot\sigma\right)e_{1},\\ =&\left(\overline{F_{\Omega_{1}}^{f,1}}\cdot\mathbb{I}+\overline{F_{\Omega_{1}}^{f,2}}\cdot\sigma\right)\left(F_{\Omega_{1}}^{f,1}\cdot\mathbb{I}+F_{\Omega_{1}}^{f,2}\cdot\sigma\right)e_{1}\\ =&\left[\left(\overline{F_{\Omega_{1}}^{f,1}}F_{\Omega_{1}}^{f,1}-\overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,2}\right)\cdot\mathbb{I}+\left(\overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,1}+\overline{F_{\Omega_{1}}^{f,1}}F_{\Omega_{1}}^{f,2}\right)\cdot\sigma\right]e_{1}\\ =&\begin{pmatrix}\overline{F_{\Omega_{1}}^{f,1}}F_{\Omega_{1}}^{f,1}-\overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,2}\\ \\ \overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,1}+\overline{\overline{F_{\Omega_{1}}^{f,2}}F_{\Omega_{1}}^{f,1}}\end{pmatrix}.\end{split}

This verifies equation (4.1). Furthermore, FΩ1f,cFΩ1fF_{\Omega_{1}}^{f,c}*F_{\Omega_{1}}^{f} is a path-slice stem function of fΩ1sf^{s}_{\Omega_{1}}, as indicated by Lemma 3.9 and Proposition 2.9.

Next, for any I𝕊I\in\mathbb{S} and zIΩIz^{I}\in\Omega_{I}, Lemma 3.3 ensures the existence of a path γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}) with γI(1)=zI\gamma^{I}(1)=z^{I}. Applying the relation in equation (2.1) to our scenario:

fΩ1s(zI)\displaystyle f^{s}_{\Omega_{1}}(z^{I}) =fΩ1sγI(1)=(1,I)(FΩ1f,cFΩ1f)(γ)I.\displaystyle=f^{s}_{\Omega_{1}}\circ\gamma^{I}(1)=(1,I)\left(F_{\Omega_{1}}^{f,c}*F_{\Omega_{1}}^{f}\right)(\gamma)\in\mathbb{C}_{I}.

This implies that fΩ1sf^{s}_{\Omega_{1}} maps ΩI\Omega_{I} into I\mathbb{C}_{I}. Since I𝕊I\in\mathbb{S} was chosen arbitrarily, fΩ1sf^{s}_{\Omega_{1}} is confirmed to be slice-preserving. ∎

5. Zeros of path-slice functions

This section is dedicated to exploring the zero sets of both path-slice and slice regular functions. A key discovery presented here is the path-slice analytic nature of the zero set of a slice regular function. This significant aspect is comprehensively elaborated in Theorem 5.8, highlighting the intricate relationship between the zeros of these functions and their underlying analytical properties.

Theorem 5.1.

(Representation formula) Consider a subset Ω\Omega within sn\mathbb{H}_{s}^{n} and let ff be a function from the class 𝒫𝒮(Ω)\mathcal{PS}(\Omega). The representation of ff when composed with γI\gamma^{I} can be expressed as

(5.1) fγI=(1,I)(1J1K)1(fγJfγK),f\circ\gamma^{I}=(1,I)\begin{pmatrix}1&J\\ 1&K\end{pmatrix}^{-1}\begin{pmatrix}f\circ\gamma^{J}\\ f\circ\gamma^{K}\end{pmatrix},

where this holds true for every path γ\gamma in the set 𝒫(n,Ω)\mathscr{P}(\mathbb{C}^{n},\Omega) and for every I,J,KI,J,K in the set 𝕊(Ω,γ)\mathbb{S}(\Omega,\gamma), given that JJ and KK are distinct.

Proof.

For any given t[0,1]t\in[0,1], we define a new path α:[0,1]n\alpha:[0,1]\to\mathbb{C}^{n} by

α(s)=γ(ts)\alpha(s)=\gamma(ts)

for each s[0,1]s\in[0,1]. The path αL\alpha^{L} for any LI,J,KL\in{I,J,K} is contained within γLΩ\gamma^{L}\subset\Omega, indicating α𝒫(n,Ω)\alpha\in\mathscr{P}(\mathbb{C}^{n},\Omega) and I,J,K𝕊(Ω,α)I,J,K\in\mathbb{S}(\Omega,\alpha).

By (2.4), the stem function FF of ff satisfies

F(α)=(1J1K)1(f(αJ(1))f(αK(1))).F(\alpha)=\begin{pmatrix}1&J\\ 1&K\end{pmatrix}^{-1}\begin{pmatrix}f(\alpha^{J}(1))\\ f(\alpha^{K}(1))\end{pmatrix}.

Since α(1)=γ(t)\alpha(1)=\gamma(t), by applying the definition of path-slice functions in (2.1), we find

fγI(t)=fαI(1)=(1,I)F(α)=(1,I)(1J1K)1(fαJ(1)fαK(1))=(1,I)(1J1K)1(fγJ(t)fγK(t)).\begin{split}f\circ\gamma^{I}(t)=&f\circ\alpha^{I}(1)=(1,I)F(\alpha)\\ =&(1,I)\begin{pmatrix}1&J\\ 1&K\end{pmatrix}^{-1}\begin{pmatrix}f\circ\alpha^{J}(1)\\ f\circ\alpha^{K}(1)\end{pmatrix}\\ =&(1,I)\begin{pmatrix}1&J\\ 1&K\end{pmatrix}^{-1}\begin{pmatrix}f\circ\gamma^{J}(t)\\ f\circ\gamma^{K}(t)\end{pmatrix}.\end{split}

This relationship is valid for any t[0,1]t\in[0,1], thus proving the representation formula for the path-slice function ff. ∎

To analyze the zeros of path-slice functions, the following lemmas are required.

For any subset Ω\Omega of sn\mathbb{H}_{s}^{n} and a path γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega), we denote

γ𝕊(Ω,γ)(1):={γI(1):I𝕊(Ω,γ)}.\gamma^{\mathbb{S}(\Omega,\gamma)}(1):=\{\gamma^{I}(1):I\in\mathbb{S}(\Omega,\gamma)\}.
Lemma 5.2.

Consider a subset Ω\Omega of sn\mathbb{H}_{s}^{n} and a path-slice function ff defined on Ω\Omega. For a given path γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and distinct quaternionic units J,K𝕊(Ω,γ)J,K\in\mathbb{S}(\Omega,\gamma), suppose that both γJ(1)\gamma^{J}(1) and γK(1)\gamma^{K}(1) belong to the zero set of ff, denoted as 𝒵(f)={qΩ:f(q)=0}\mathcal{Z}(f)=\{q\in\Omega:f(q)=0\}. Then, the following inclusion holds:

(5.2) 𝒵(f)γ𝕊(Ω,γ)(1).\mathcal{Z}(f)\supset\gamma^{\mathbb{S}(\Omega,\gamma)}(1).
Proof.

Since γJ(1)\gamma^{J}(1) and γK(1)\gamma^{K}(1) are in 𝒵(f)\mathcal{Z}(f), we have f(γJ(1))=0f(\gamma^{J}(1))=0 and f(γK(1))=0f(\gamma^{K}(1))=0. By the definition of path-slice functions and the representation formula, for any I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma), we have

f(γI(1))=(1,I)(1J1K)1(f(γJ(1))f(γK(1))).f(\gamma^{I}(1))=(1,I)\begin{pmatrix}1&J\\ 1&K\end{pmatrix}^{-1}\begin{pmatrix}f(\gamma^{J}(1))\\ f(\gamma^{K}(1))\end{pmatrix}.

Since both f(γJ(1))f(\gamma^{J}(1)) and f(γK(1))f(\gamma^{K}(1)) are zero, the right-hand side of the equation evaluates to zero. Therefore, f(γI(1))=0f(\gamma^{I}(1))=0 for all I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma), implying that γI(1)𝒵(f)\gamma^{I}(1)\in\mathcal{Z}(f). This completes the proof. ∎

Lemma 5.3.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} be self-stem-preserving, and f𝒫𝒮(Ω)f\in\mathcal{PS}(\Omega). Then

(5.3) 𝒵(f)𝒵(fΩ1s).\mathcal{Z}(f)\subset\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right).
Proof.

Let q𝒵(f)q\in\mathcal{Z}(f), and let FF be a path-slice stem function of ff. Given that Ω\Omega is self-stem-preserving, it follows that Ω\Omega is also real-path-connected. Consequently, there exists a path γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega) and a quaternionic unit I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma) such that γI(1)=q\gamma^{I}(1)=q. From the definition of path-slice functions and equation (2.1), we have:

0=f(q)=f(γI(1))=(1,I)F(γ)=(1,I)FΩf(γ)=FΩf,1(γ)+IFΩf,2(γ).0=f(q)=f(\gamma^{I}(1))=(1,I)F(\gamma)=(1,I)F_{\Omega}^{f}(\gamma)=F_{\Omega}^{f,1}(\gamma)+IF_{\Omega}^{f,2}(\gamma).

This implies

(5.4) FΩf,1(γ)¯FΩf,1(γ)=FΩf,2(γ)¯FΩf,2(γ),\overline{F_{\Omega}^{f,1}(\gamma)}F_{\Omega}^{f,1}(\gamma)=\overline{F_{\Omega}^{f,2}(\gamma)}F_{\Omega}^{f,2}(\gamma),

and

(5.5) FΩf,2(γ)¯FΩf,1(γ)=FΩf,2(γ)¯[IFΩf,2(γ)]=[IFΩf,2(γ)¯]FΩf,2(γ)=FΩf,1(γ)¯FΩf,2(γ)=FΩf,2(γ)¯FΩf,1(γ)¯.\begin{split}\overline{F_{\Omega}^{f,2}(\gamma)}F_{\Omega}^{f,1}(\gamma)=&\overline{F_{\Omega}^{f,2}(\gamma)}\left[-I\cdot F_{\Omega}^{f,2}(\gamma)\right]\\ =&-\left[\overline{-IF_{\Omega}^{f,2}(\gamma)}\right]\cdot F_{\Omega}^{f,2}(\gamma)\\ =&-\overline{F_{\Omega}^{f,1}(\gamma)}F_{\Omega}^{f,2}(\gamma)\\ =&-\overline{\overline{F_{\Omega}^{f,2}(\gamma)}F_{\Omega}^{f,1}(\gamma)}.\end{split}

Therefore, equation (5.3) is satisfied due to (5.4), (5.5), and the definition of fΩsf^{s}_{\scriptscriptstyle{\Omega}} as the symmetrization of ff on Ω\Omega. ∎

Lemma 5.4.

Given subsets Ω1\Omega_{1} and Ω2\Omega_{2} of sn\mathbb{H}_{s}^{n}, where Ω1\Omega_{1} is real-path-connected and Ω2\Omega_{2} is Ω1\Omega_{1}-stem-preserving, consider a path-slice function ff defined on Ω2\Omega_{2}. Suppose γ\gamma is a path in 𝒫(n,Ω1)\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), and II is a quaternionic unit in 𝕊(Ω1,γ)\mathbb{S}(\Omega_{1},\gamma) such that the endpoint γI(1)\gamma^{I}(1) lies in the zero set 𝒵(fΩ1s)\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right). Then we have

γ𝕊(Ω1,γ)(1)𝒵(fΩ1s).\gamma^{\mathbb{S}(\Omega_{1},\gamma)}(1)\subset\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right).
Proof.

Drawing upon Theorem 4.3, we can deduce:

0=fΩ1s(γI(1))\displaystyle 0=f^{s}_{\scriptscriptstyle{\Omega_{1}}}(\gamma^{I}(1)) =\displaystyle= (1,I)FΩ1f,s(γ)\displaystyle(1,I)F_{\Omega_{1}}^{f,s}(\gamma)
=\displaystyle= (1,I)(FΩ1f,s,1(γ)FΩ1f,s,2(γ))\displaystyle(1,I)\begin{pmatrix}F_{\Omega_{1}}^{f,s,1}(\gamma)\\ \\ F_{\Omega_{1}}^{f,s,2}(\gamma)\end{pmatrix}
=\displaystyle= FΩ1f,s,1(γ)+IFΩ1f,s,2(γ),\displaystyle F_{\Omega_{1}}^{f,s,1}(\gamma)+IF_{\Omega_{1}}^{f,s,2}(\gamma),

where FΩ1f,s,1(γ)F_{\Omega_{1}}^{f,s,1}(\gamma) and FΩ1f,s,2(γ)F_{\Omega_{1}}^{f,s,2}(\gamma) are real-valued. This implies that

FΩ1f,s,1(γ)=FΩ1f,s,2(γ)=0,F_{\Omega_{1}}^{f,s,1}(\gamma)=F_{\Omega_{1}}^{f,s,2}(\gamma)=0,

leading to FΩ1f,s(γ)F_{\Omega_{1}}^{f,s}(\gamma) being zero. As a result, for every J𝕊(Ω,γ)J\in\mathbb{S}(\Omega,\gamma):

fΩ1s(γJ(1))=(1,I)FΩ1f,s(γ)=0,f^{s}_{\scriptscriptstyle{\Omega_{1}}}(\gamma^{J}(1))=(1,I)F_{\Omega_{1}}^{f,s}(\gamma)=0,

thereby confirming that the endpoints γ𝕊(Ω,γ)(1)\gamma^{\mathbb{S}(\Omega,\gamma)}(1) are all within the zero set 𝒵(fΩ1s)\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right). ∎

Now we introduce a new concept of path-slice analytic set within a slice-domain Ω\Omega in sn\mathbb{H}_{s}^{n}.

Definition 5.5.

A subset AA of a slice-domain Ωsn\Omega\subset\mathbb{H}_{s}^{n} is termed path-slice analytic if it either encompasses the entire domain Ω\Omega or, for each path γ𝒫(n,Ω)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega), certain conditions are met. Specifically, there exists a radius r>0r>0, a set of radii r[I](0,r]r{[I]}\in(0,r] for each I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma), and an analytic subset EE of the ball Bn(γ(1),r)B_{\mathbb{C}^{n}}(\gamma(1),r), distinct from the entire ball, such that

ABI(γI(1),r[I])EIA\cap B_{I}\left(\gamma^{I}(1),r_{[I]}\right)\subseteq E^{I}

for any I𝕊(Ω,γ).I\in\mathbb{S}(\Omega,\gamma). Here, EIE^{I} denotes the slice of the set EE corresponding to I𝕊(Ω,γ)I\in\mathbb{S}(\Omega,\gamma).

We aim to establish that the zeros of a path-slice regular function constitute a path-slice analytic set. To achieve this, we will need several lemmas.

Lemma 5.6.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be a real-path-connected slice-domain, 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 𝒵(fΩ1s)\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right) is path-slice analytic.

Proof.

According to Theorem 3.8, fΩ1cf^{c}_{\scriptscriptstyle{\Omega_{1}}} is slice regular. It follows from Theorem 2.16 that fΩ1s=fΩ1cff^{s}_{\scriptscriptstyle{\Omega_{1}}}=f^{c}_{\scriptscriptstyle{\Omega_{1}}}*f is also slice regular. Let γ𝒫(n,Ω1)\gamma\in\mathscr{P}(\mathbb{C}^{n},\Omega_{1}), I𝕊(Ω1,γ)I\in\mathbb{S}(\Omega_{1},\gamma), r(0,rγ,Ω{I})r\in\left(0,r_{\gamma,\Omega}^{\{I\}}\right),

r[J]=min{r,rγ,Ω{J}},J𝕊(Ω1,γ),r_{[J]}=\min\left\{r,r_{\gamma,\Omega}^{\{J\}}\right\},\qquad\forall\ J\in\mathbb{S}(\Omega_{1},\gamma),

and

E[J]:=𝒵(f)BJ(γJ(1),r[J]),J𝕊(Ω1,γ).E_{[J]}:=\mathcal{Z}(f)\cap B_{J}(\gamma^{J}(1),r_{[J]}),\qquad\forall\ J\in\mathbb{S}(\Omega_{1},\gamma).

If there is no analytic set EE in Bn(γ(1),r)B_{\mathbb{C}^{n}}(\gamma(1),r) such that EBn(γ(1),r)E\neq B_{\mathbb{C}^{n}}(\gamma(1),r) and E[I]EIE_{[I]}\subset E^{I}. Then by [Dou2023002, Splitting Lemma 3.3] and Identity Principle in complex analysis, Bn(γ(1),r)𝒵(fΩ1s)B_{\mathbb{C}^{n}}(\gamma(1),r)\subset\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right). According to [Dou2023002, Indentity Principle 3.5], fΩ1s0f^{s}_{\scriptscriptstyle{\Omega_{1}}}\equiv 0. It implies that 𝒵(fΩ1s)=Ω\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right)=\Omega is path-slice analytic.

Otherwise, let EE be an analytic set in Bn(γ(1),r)B_{\mathbb{C}^{n}}(\gamma(1),r) with EBn(γ(1),r)E\neq B_{\mathbb{C}^{n}}(\gamma(1),r) and E[I]EIE_{[I]}\subset E^{I}, and let zJE[J]z^{J}\in E_{[J]}. Then

(γγ(1)z)J(1)=zJ𝒵(fΩ1s).\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right)^{J}(1)=z^{J}\in\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right).

By Lemma 5.4,

zI=(γγ(1)z)I(1)(γγ(1)z)𝕊(Ω,γ)(1)𝒵(fΩ1s).z^{I}=\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right)^{I}(1)\in\left(\gamma\circ\mathcal{L}_{\gamma(1)}^{z}\right)^{\mathbb{S}(\Omega,\gamma)}(1)\subset\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right).

It implies that zIE[I]z^{I}\in E_{[I]}, z(ΨiI)1(E[I])Ez\in\left(\Psi_{i}^{I}\right)^{-1}\left(E_{[I]}\right)\subset E and zJEJz^{J}\in E^{J}. Therefore, E[J]EJE_{[J]}\subset E^{J} and

𝒵(fΩ1s)BI(γI(1),r)=E[J]EJ.\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right)\cap B_{I}(\gamma^{I}(1),r)=E_{[J]}\subset E^{J}.

By Definition 5.5, 𝒵(fΩ1s)\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega_{1}}}\right) is path-slice analytic. ∎

Proposition 5.7.

Let Ω1sn\Omega_{1}\subset\mathbb{H}_{s}^{n} be a real-path-connected slice-domain, 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

(5.6) fΩ1c=f¯andfΩ1s=|f|2onΩ.f^{c}_{\scriptscriptstyle{\Omega_{1}}}=\overline{f}\qquad\mbox{and}\qquad f^{s}_{\scriptscriptstyle{\Omega_{1}}}=|f|^{2}\qquad\mbox{on}\qquad\Omega_{\mathbb{R}}.
Proof.

According to (2.10) and (3.1),

fΩ1c=(1,)Ω1f,c=(1,)ConjΩ1f=(1,)Conj(f0)=(1,)(f¯0)=f¯,onΩ.\begin{split}f^{c}_{\scriptscriptstyle{\Omega_{1}}}=&(1,\mathfrak{I})\mathscr{F}_{\Omega_{1}}^{f,c}=(1,\mathfrak{I}){\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ\mathscr{F}_{\Omega_{1}}^{f}\\ =&(1,\mathfrak{I}){\mathop{\mathrm{Conj}}}_{\mathbb{H}}\circ\begin{pmatrix}f\\ 0\end{pmatrix}=(1,\mathfrak{I})\begin{pmatrix}\overline{f}\\ 0\end{pmatrix}=\overline{f},\end{split}\qquad\mbox{on}\qquad\Omega_{\mathbb{R}}.

By (2.7) and (2.10),

fΩ1s=fΩ1cf=(fΩ1c,fΩ1c)Ω1f=(f¯,f¯)(f0)=f¯f=|f|2,onΩ.f^{s}_{\scriptscriptstyle{\Omega_{1}}}=f^{c}_{\scriptscriptstyle{\Omega_{1}}}*f=(f^{c}_{\scriptscriptstyle{\Omega_{1}}},\mathfrak{I}f^{c}_{\scriptscriptstyle{\Omega_{1}}})\mathscr{F}_{\Omega_{1}}^{f}=(\overline{f},\mathfrak{I}\overline{f})\begin{pmatrix}f\\ 0\end{pmatrix}=\overline{f}f=|f|^{2},\qquad\mbox{on}\qquad\Omega_{\mathbb{R}}.

Ultimately, we are able to demonstrate that the zeros of a path-slice regular function form a path-slice analytic set.

Theorem 5.8.

Let Ωsn\Omega\subset\mathbb{H}_{s}^{n} be a self-stem-preserving slice-domain, and f𝒮(Ω)f\in\mathcal{SR}(\Omega). Then 𝒵(f)\mathcal{Z}(f) is path-slice analytic.

Proof.

If 𝒵(fΩs)=Ω\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega}}\right)=\Omega, then by (5.6),

0fΩs=|f|2,onΩ.0\equiv f^{s}_{\scriptscriptstyle{\Omega}}=|f|^{2},\qquad\mbox{on}\qquad\Omega_{\mathbb{R}}.

It implies that f0f\equiv 0 on Ω\Omega_{\mathbb{R}}. According to [Dou2023002, Indentity Principle 3.5], f0f\equiv 0. It implies that 𝒵(f)=Ω\mathcal{Z}\left(f\right)=\Omega is path-slice analytic.

Otherwise, 𝒵(fΩs)Ω\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega}}\right)\neq\Omega. According to (5.3), 𝒵(f)𝒵(fΩs)\mathcal{Z}(f)\subset\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega}}\right). It follows from Proposition 5.6 that 𝒵(fΩs)\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega}}\right) is path-slice analytic with 𝒵(fΩs)Ω\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega}}\right)\neq\Omega, so is 𝒵(f)𝒵(fΩs)\mathcal{Z}(f)\subset\mathcal{Z}\left(f^{s}_{\scriptscriptstyle{\Omega}}\right). ∎

References