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

A higher Dimensional Marcinkiewicz Exponent and the Riemann Boundary Value Problems for Polymonogenic Functions on Fractals Domains

Carlos Daniel Tamayo Castroa, Juan Bory Reyesb
( aInstituto de Matemáticas. Universidad Nacional Autónoma de México, Mexico City, Mexico.
cdtamayoc@comunidad.unam.mx
bSEPI-ESIME-Zacatenco, Instituto Politécnico Nacional, Mexico City, Mexico
juanboryreyes@yahoo.com
)
Abstract

We use a high-dimensional version of the Marcinkiewicz exponent, a metric characteristic for non-rectifiable plane curves, to present a direct application to the solution of some kind of Riemann boundary value problems on fractal domains of Euclidean space n+1,n2\mathbb{R}^{n+1},n\geq 2 for Clifford algebra-valued polymonogenic functions with boundary data in classes of higher order Lipschitz functions. Sufficient conditions to guarantee the existence and uniqueness of solution to the problems are proved. To illustrate the delicate nature of this theory we described a class of hypersurfaces where the results are more refined than those that exist in literature.

Keywords: Clifford analysis, Riemann boundary value problem, polymonogenic functions, fractal boundaries.
MSC(2020): Primary 30G35, 28A80, Secondary 30E20, 30G30

1 Introduction

Riemann boundary value problems (RBVP for short) for analytic functions in a bounded Jordan domain of the complex plane are widely discussed and applied to many branches of mathematics, physics and engineering. Along classical lines, the Cauchy type integral is used as the main tool in the treatment of these boundary value problems, see [13, 23, 27] for more details.

Although the Cauchy type integral loses its meaning on fractal curves, the RBVP is equally valid. Pioneering works in the field were given by Kats in [15]. In particular, solvability conditions involving the Hölder exponent of the data function associated with the problem and the upper Minkowski dimension of the boundary are established. For a recent account of this approach we refer the reader to [3, 16, 17].

The Marcinkiewicz exponent was introduced in [18, 20, 19], once again restricting the discussion to the complex analysis context. The use of this new metric characteristic of the boundary of the domains, considerably sharp the solvability conditions in the mentioned above study of the RBVPs.

Clifford analysis [9, 14] offers an elegant generalization of analytic functions from the complex plane to higher-dimensions. Monogenic functions on Euclidean space are basic to Clifford analysis, they are defined as smooth functions with values in the corresponding Clifford algebra, which are null solutions to a generalized Cauchy-Riemann operator. In view of the fact that the generalized Cauchy-Riemann operator factorizes the higher Laplace operator, monogenic functions are also harmonic.

As was early pointed out in [28, pp 22, 24], significant obstacles exist when trying to give a thorough treatment to the RBVP for monogenic functions. These are influenced by the fact that the product of two monogenic functions is not necessarily a monogenic. This is due to the non-commutativity properties of the Clifford algebra. This explains why an explicit solution to these boundary value problems has been found only for the jump problem and some slight modifications, see [2, 7] and the references given there.

Almost all uniqueness and existence theorems for the solutions of RBVPs for analytic functions on the complex plane involving the Minkowski dimension of the boundary can be set in the context of Clifford analysis, see for instance [4, 5, 30].

In [1, 10, 32, 26, 21] a kind of RBVPs have been studied for polymonogenic functions, i.e., null solutions to iterated generalized Cauchy-Riemann operator.

However, to the best of our knowledge, a full description of the relations between the high-dimensional version of the Marcinkiewicz exponent, introduced in [30] and the RBVPs in fractal domains of Euclidean spaces for polymonogenic functions with boundary data in classes of higher order Lipschitz functions still remains open.

This paper aims to obtain solvability conditions to RBVPs for Clifford algebra-valued polymonogenic functions involving the high-dimensional Marcinkiewicz exponent of the hypersurface. Besides, to show that this condition strictly improves those requiring the Minkowski dimension, we need to describe an uncountable class of hypersurfaces, such that the relationship between its Minkowski dimension and Marcinkiewicz exponent is with strict inequality.

The content of this paper is structured as follows: Section 2 presents the basic notions and terminology of the theory of monogenic functions. Moreover, we give a brief review of the definitions of higher order Lipschitz functions, fractal dimensions and Marcinkiewicz exponent. Also, a Whitney type extension theorem is stated. In Section 3, we obtain solvability and uniqueness conditions to a kind of RBVPs for Clifford algebra-valued polymonogenic functions with data higher order Lipschitz functions. Section 4 provides a description of a class of hypersurfaces, where the conditions obtained in Section 3 improve those involving the Minkowski dimension that exist in literature.

2 Preliminaries and Notations

This section presents the essential background needed for developing the results in the subsequent sections. It is divided into three subsections, each devoted to a fundamental component of the strategy for proving the results. The first subsection contains some basic facts about Clifford analysis. In second subsection, we give a brief exposition of the data classes of functions required in the problems considered, and a corresponding Whitney-type extension theorem is presented. Finally, we introduce the fractal dimensions and characteristics of the boundaries that will be treated in the work.

2.1 Clifford Algebras and Monogenic Functions

In this section, we recall some basic facts about Clifford analysis which will be needed in the sequel. It could be seen in more detail in the literature, for instance [9, 14].

Definition 1.

We consider the Clifford algebra associated with n\mathbb{R}^{n} and endowed with the usual Euclidean metric, as the minimal extension of n\mathbb{R}^{n} to a unitary, associative algebra 𝒞(n)\mathcal{C}\ell(n) over the real numbers, satisfying

x2=|x|2,\displaystyle x^{2}=-\arrowvert x\arrowvert^{2},

for any xnx\in\mathbb{R}^{n}.

It thus follows that, the standard orthonormal basis of n\mathbb{R}^{n}, denoted by {ej}j=1n\{e_{j}\}_{j=1}^{n}, is subject to the basic multiplication rules

eiej+ejei=2δij,e_{i}e_{j}+e_{j}e_{i}=-2\delta_{ij},

were δij\delta_{ij} is the Kronecker delta.

For arbitrary a𝒞(n)a\in\mathcal{C}\ell(n) we have a=ANaAeAa=\sum\limits_{A\subseteq N}a_{A}e_{A}, N={1,,n}N=\{1,\ldots,n\}, aAa_{A}\in\mathbb{R} where e=e0=1e_{\emptyset}=e_{0}=1, e{j}=eje_{\{j\}}=e_{j} and eA=eβ1eβke_{A}=e_{\beta_{1}}\cdots e_{\beta_{k}} for A={β1,,βk}A=\{\beta_{1},\ldots,\beta_{k}\} with βj{1,,n}\beta_{j}\in\{1,\ldots,n\} and β1<<βk\beta_{1}<\ldots<\beta_{k}.

An important subspace of 𝒞(n)\mathcal{C}\ell(n) is the so-called space of paravectors and its elements have the form x=j=0nxjej,x=\sum_{j=0}^{n}x^{j}e_{j},. Each x=(x0,x1,,xn)n+1x=(x^{0},x^{1},\ldots,x^{n})\in\mathbb{R}^{n+1} will be identified with a paravector.

The algebra norm of arbitrary aa is defined by |a|:=(AaA2)12\arrowvert a\arrowvert:=\left(\sum_{A}a_{A}^{2}\right)^{\frac{1}{2}} and 𝒞(n)\mathcal{C}\ell(n) becomes an Euclidean space.

The conjugation in 𝒞(n)\mathcal{C}\ell(n) is defined as the anti-involution aa¯:=AaAe¯Aa\mapsto\overline{a}:=\sum_{A}a_{A}\overline{e}_{A} such that

e¯A:=(1)keβkeβ2eβ1.\overline{e}_{A}:=(-1)^{k}e_{\beta_{k}}\cdots e_{\beta_{2}}e_{\beta_{1}}.

For every paravector xx the relation

xx¯=x¯x=|x|2x\overline{x}=\overline{x}x=|x|^{2}

holds.

An 𝒞(n)\mathcal{C}\ell(n)-valued function uu defined over Ωn+1\Omega\subset\mathbb{R}^{n+1} has the representation

u(x)=AuA(x)eA,u(x)=\sum_{A}u_{A}(x)e_{A},

where uAu_{A} are \mathbb{R}-valued components.

Unless otherwise states we assume that uu to be 𝒞(n)\mathcal{C}\ell(n)-valued. Properties such as continuity or differentiability have to be understood component-wise.

Clifford analysis is mainly centered around the concept of monogenic functions which are null solutions of the generalized Cauchy-Riemann operator in n+1\mathbb{R}^{n+1} defined by

𝒟:=j=0nejxj.\mathcal{D}:=\sum^{n}_{j=0}e_{j}\frac{\partial}{\partial x_{j}}.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be an open set and uC1(Ω)u\in C^{1}(\Omega) then uu will be called left (respectively right) monogenic in Ω\Omega if 𝒟u=0\mathcal{D}u=0 (respectively u𝒟=0u\mathcal{D}=0) in Ω\Omega.

The fundamental solution for the Cauchy-Riemann operator is

E(x)=1σn+1x¯|x|n+1,xn+1{0}\begin{array}[]{cc}E(x)=\dfrac{1}{\sigma_{n+1}}\dfrac{\overline{x}}{|x|^{n+1}},&x\in\mathbb{R}^{n+1}\setminus\{0\}\end{array}

where σn+1\sigma_{n+1} stands for the hypersurface area of the unit sphere in n+1\mathbb{R}^{n+1}.

Polymonogenic functions fCk(Ω)f\in C^{k}(\Omega) of order kk, emerge as the solutions of the iterated action of the operator 𝒟\mathcal{D}

𝒟kf=0,\mathcal{D}^{k}f=0,

in Ωn+1\Omega\subset\mathbb{R}^{n+1}. This notion goes back as far as [6, 8, 11].

For arbitrary kNk\in N, let us introduce

Ek(x)=1σn+1x¯(x¯+x)k12k1(k1)!|x|n+1.E^{k}(x)=\dfrac{1}{\sigma_{n+1}}\dfrac{\overline{x}(\overline{x}+x)^{k-1}}{2^{k-1}(k-1)!\lvert x\rvert^{n+1}}.

Note that for k=1k=1

E1(x)=E(x),\displaystyle E^{1}(x)=E(x),

as it is easy to check.

A direct computation shows that

𝒟Ek(x)=Ek1(x).\displaystyle\mathcal{D}E^{k}(x)=E^{k-1}(x).

Consequently, by decreasing induction on kk,

𝒟kEk(x)=𝒟E1(x)=0,\displaystyle\mathcal{D}^{k}E^{k}(x)=\mathcal{D}E^{1}(x)=0, xn+1{0}.\displaystyle x\in\mathbb{R}^{n+1}\setminus\{0\}.

Let us now introduce the Teodorescu transform related to the theory of monogenic functions, see [14] for more details.

Definition 2.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be a domain and let uC1(Ω¯)u\in C^{1}(\overline{\Omega}), the so-called Teodorescu transform is defined by

(TΩu)(x):=ΩE(yx)u(y)𝑑V(y),\displaystyle(T_{\Omega}u)(x):=-\int\limits_{\Omega}E(y-x)u(y)dV(y), xn+1,\displaystyle x\in\mathbb{R}^{n+1},

where dV(y)dV(y) is the volume element.

Sufficient conditions for the Hölder continuity of TΩuT_{\Omega}u are established by our next theorem.

Theorem 1.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be a domain and let uLp(Ω)u\in L^{p}(\Omega) for p>n+1p>n+1. Then

  • The integral (TΩu)(x)(T_{\Omega}u)(x) exists in the whole n+1\mathbb{R}^{n+1} and tends to zero as |x|\arrowvert x\arrowvert\rightarrow\infty. Besides, TΩuT_{\Omega}u is a monogenic function in n+1Ω¯\mathbb{R}^{n+1}\setminus\overline{\Omega}.

  • For x,yn+1x,y\in\mathbb{R}^{n+1}, and xyx\neq y, we have

    |(TΩu)(x)(TGu)(y)|C2(Ω,p,n)up|xy|pn1p.\arrowvert(T_{\Omega}u)(x)-(T_{G}u)(y)\arrowvert\leq C_{2}(\Omega,p,n)\|u\|_{p}\arrowvert x-y\arrowvert^{\frac{p-n-1}{p}}.

Conditions for the derivability of TΩuT_{\Omega}u on Ω\Omega are given as follows

Theorem 2.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be as stated above. Then TΩuT_{\Omega}u is a differentiable function for every xΩx\in\Omega with

xj(TΩu)(x)=Ωxj[En(yx)]u(y)𝑑V(y)+ej¯u(x)n+1.\frac{\partial}{\partial x_{j}}(T_{\Omega}u)(x)=-\int\limits_{\Omega}\frac{\partial}{\partial x_{j}}[E_{n}(y-x)]u(y)dV(y)+\overline{e_{j}}\dfrac{u(x)}{n+1}.

In particular, the identity

𝒟(TΩu)(x)=u(x),xΩ.\displaystyle\mathcal{D}(T_{\Omega}u)(x)=u(x),\ \ x\in\Omega.

holds.

Now, we introduce a kind of polymonogenic Teodorescu transform following [6]. Due to its good properties, this integral operator will play an essential role in the strategies described below. For a recent account of the subject we refer the reader to [22].

Definition 3.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be a bounded domain and uL1(Ω¯)u\in L^{1}(\overline{\Omega}). Then for kk\in\mathbb{N} we define the kk-polymonogenic Teodorescu transform as follow

TΩku(x):=(1)kΩEk(yx)u(y)𝑑V(y),T^{k}_{\Omega}u(x):=(-1)^{k}\int\limits_{\Omega}E^{k}(y-x)u(y)dV(y),

where dV(y)dV(y) is the volume element.

By derivability properties of TΩku(x)T^{k}_{\Omega}u(x), the following equality is achieved

𝒟TΩku=TΩk1u,\displaystyle\mathcal{D}T^{k}_{\Omega}u=T^{k-1}_{\Omega}u, k2.\displaystyle k\geq 2.

Therefore, decreasing induction on kk combining with Theorems 1 and 2 gives

𝒟kTΩku=𝒟TΩ1u={u,inΩ0,inn+1Ω¯.\mathcal{D}^{k}T^{k}_{\Omega}u=\mathcal{D}T^{1}_{\Omega}u=\left\{\begin{array}[]{ccc}u,&in&\Omega\\ 0,&in&\mathbb{R}^{n+1}\setminus\overline{\Omega}.\end{array}\right. (1)

2.2 Function Classes and a kind of a Whitney Type Theorem

In order to state the main problems that we address, it is necessary to define the suitable classes where the data functions will be considered. These are the higher order Lipschitz classes, on the basis of which some Whitney-type extension theorems will be presented.

As a first step, we shall recall the class of pp-integrables functions, see [14].

Definition 4.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be a domain and let u:Ω𝒞(n)u:\Omega\mapsto\mathcal{C}\ell(n). Lp(Ω),0<p<L^{p}(\Omega),0<p<\infty denotes the space of all equivalence classes of Lebesgue measurable functions equal almost everywhere, such that

up<.\|u\|_{p}<\infty.

where

up:=(Ω|u|p𝑑V)1p,\|u\|_{p}:=\left(\int\limits_{\Omega}|u|^{p}dV\right)^{\frac{1}{p}},

Let 𝐄{\bf E} be a closed subset of n+1\mathbb{R}^{n+1}. We write j=(j0,j1,,jn)j=(j_{0},j_{1},\cdots,j_{n}) a n-dimensional multi-index of order |j|=j0+j1++jn|j|=j_{0}+j_{1}+\cdots+j_{n}, where j0,j1,,jnj_{0},j_{1},\dots,j_{n} are non-negative integers. In addition, we have, j!=j0!j1!jn!j!=j_{0}!j_{1}!\cdots j_{n}! and xj=xj0x1j1xnjnx^{j}=x_{j_{0}}x_{1}^{j_{1}}\cdots x_{n}^{j_{n}}.

Definition 5.

Let 0<ν10<\nu\leq 1. We shall say that a real-valued function ff, defined in 𝐄{\bf E}, belongs to Lip(𝐄,k+ν){\mbox{Lip}}({\bf E},k+\nu) if there exist real-valued bounded functions f(j)f^{(j)}, 0<|j|k0<|j|\leq k, defined on 𝐄{\bf E}, with f(0)=ff^{(0)}=f, and so that

f(j)(x)=|j+l|kf(j+l)(y)l!(xy)l+Rj(x,y),x,y𝐄,f^{(j)}(x)=\sum_{|j+l|\leq k}\frac{f^{(j+l)}(y)}{l!}(x-y)^{l}+R_{j}(x,y),\,\,x,y\in{\bf E}, (2)

where

|f(j)(x)|M,|Rj(x,y)|M|xy|k+ν|j|,x,y𝐄,|j|k,|f^{(j)}(x)|\leq M,\,\,\,|R_{j}(x,y)|\leq M|x-y|^{k+\nu-|j|},\,\,x,y\in{\bf E},|j|\leq k, (3)

being MM a positive constant.

The space of all these functions is named higher order Lipschitz spaces, see for instance [29]. When k=1k=1 and 𝐄{\bf E} is a compact set, these functions reduce to the classical Hölder continuous functions.

A norm in Lip(𝐄,k+ν){\mbox{Lip}}({\bf E},k+\nu) is defined as the smallest MM satisfying (3). In [31] was proved that Lip(𝐄,k+ν){\mbox{Lip}}({\bf E},k+\nu) endowed with this norm is a Banach space. Also, conditions for continuous and compact embeddings of generalized higher-order Lipschitz classes on a compact subset of Euclidean spaces were obtained, showing that these spaces are not only a generalization but also a refinement of the classical Lipschitz classes.

In general, an element of Lip(𝐄,k+ν){\mbox{Lip}}({\bf E},k+\nu) should be interpreted as a collection

{f(j):𝐄,|j|k}.\left\{f^{(j)}:{\bf E}\mapsto\mathbb{R},\,|j|\leq k\right\}.

In order to present a kind of a Whitney type extension theorem for polymonogenic functions, let us recall the following classical theorem by Whitney, see [29, pag. 177].

Theorem 3.

Let En+1\textbf{E}\subset\mathbb{R}^{n+1} be a closed set and let fLip(E,k+ν)f\in{\mbox{Lip}}(\textbf{E},k+\nu) with values in \mathbb{R}. Then, there exists a \mathbb{R}-valued function f~Lip(n+1,k+ν)\widetilde{f}\in{\mbox{Lip}}(\mathbb{R}^{n+1},k+\nu) satisfying

  • (j)f~|E=f(j)\partial^{(j)}\widetilde{f}\arrowvert_{E}=f^{(j)},

  • f~C(n+1E)\widetilde{f}\in C^{\infty}(\mathbb{R}^{n+1}\setminus\textbf{E}),

  • |(j)f~(x)|Cdist(x,E)ν1\arrowvert\partial^{(j)}\widetilde{f}(x)\arrowvert\leq C\mathrm{dist}(x,\textbf{E})^{\nu-1} for |j|=k+1\arrowvert j\arrowvert=k+1 and xn+1Ex\in\mathbb{R}^{n+1}\setminus\textbf{E}.

Here and subsequently the symbol

(j):=|j|x0j0x1j1xkjk,\partial^{(j)}:=\frac{\partial^{|j|}}{\partial x_{0}^{j_{0}}\partial x_{1}^{j_{1}}\cdots\partial x_{k}^{j_{k}}},

stands for the higher-order partial derivatives.

Let fLip(𝒮,k1+ν)f\in{\mbox{Lip}}(\mathcal{S},k-1+\nu) be a 𝒞(n)\mathcal{C}\ell(n)-valued function, interpreted as the collection

{f(j):𝒮𝒞(n),|j|k1}\left\{f^{(j)}:\mathcal{S}\mapsto\mathcal{C}\ell(n),\,|j|\leq k-1\right\}

with f(0)=ff^{(0)}=f satisfying (2) and (3).

In order to present a suitable version of Whitney extension theorem for 𝒞(n)\mathcal{C}\ell(n)-valued function, in [1] are constructed the following functions,

𝐟(i)=r1,,ri=0ner1erif1r1++1ri,\displaystyle\mathbf{f}^{(i)}=\sum_{r_{1},\cdots,r_{i}=0}^{n}e_{r_{1}}\cdots e_{r_{i}}f^{\textbf{1}_{r_{1}}+\cdots+\textbf{1}_{r_{i}}}, i=0,1,,k1.\displaystyle i=0,1,\dots,k-1. (4)

Here 1ri\textbf{1}_{r_{i}} is the multi-index (j0,j1,,jn)(j_{0},j_{1},\cdots,j_{n}) with

jp={1,p=ri0,pri.j_{p}=\left\{\begin{array}[]{cc}1,&p=r_{i}\\ 0,&p\neq r_{i}.\end{array}\right. (5)

We should note that the functions 𝐟(i)\mathbf{f}^{(i)} are an appropriate arranged of every function f(j)f^{(j)} with |j|=i\left|j\right|=i. In addition, 𝐟(0)\mathbf{f}^{(0)} = f(0)f^{(0)} = ff.

Let us mention an important consequence of Theorem 3. This can be found in [1].

Theorem 4.

Let En+1\textbf{E}\subset\mathbb{R}^{n+1} be a closed set and let fLip(E,k1+ν)f\in{\mbox{Lip}}(\textbf{E},k-1+\nu) with values in 𝒞(n)\mathcal{C}\ell(n). Then, there exists a 𝒞(n)\mathcal{C}\ell(n)-valued function f~Lip(n+1,k1+ν)\widetilde{f}\in{\mbox{Lip}}(\mathbb{R}^{n+1},k-1+\nu) satisfying

  • 𝒟if~|E=𝐟(i)\mathcal{D}^{i}\widetilde{f}\arrowvert_{E}=\mathbf{f}^{(i)},   i=0,1,k1i=0,1,\cdots k-1

  • f~C(n+1E)\widetilde{f}\in C^{\infty}(\mathbb{R}^{n+1}\setminus\textbf{E}),

  • |𝒟kf~(x)|Cdist(x,E)ν1\arrowvert\mathcal{D}^{k}\widetilde{f}(x)\arrowvert\leq C\mathrm{dist}(x,\textbf{E})^{\nu-1} for xn+1Ex\in\mathbb{R}^{n+1}\setminus\textbf{E}.

2.3 Fractal Dimensions and Characteristics

To make the presentation self-contained, we review some basic ideas about fractal dimensions of sets that will be required to work with domains with fractal boundaries. For a deeper discussion of this topic, we refer the reader to [12, 24, 25].

Let En+1\textbf{E}\in\mathbb{R}^{n+1} a non-empty set. For any s0s\geq 0 and δ>0\delta>0, δs(E)\mathcal{H}_{\delta}^{s}(\textbf{E}) is defined as,

δs(E):=inf{i=1diam(Ui)s:{Ui}isaδcoveringofE},\mathcal{H}_{\delta}^{s}(\textbf{E}):=\inf\{\sum_{i=1}^{\infty}\mathrm{diam}(U_{i})^{s}:\{U_{i}\}\ is\ a\ \delta-covering\ of\ \textbf{E}\},

where diam(U)\mathrm{diam}(U) is the diameter of the set UU. The infimum here is taken over all countable δ\delta-coverings Ui{U_{i}} of E for open or closed balls. Now, we can present the Hausdorff measure.

Definition 6.

The ss-dimensional Hausdorff measure is defined by the limit

s(E):=limδ0δs(E).\mathcal{H}^{s}(\textbf{E}):=\lim_{\delta\rightarrow 0}\mathcal{H}_{\delta}^{s}(\textbf{E}).

It can be shown that the ss-dimensional Hausdorff measure of a set E is almost always 0 or \infty. There is only one value of ss where the measure change between these two values. Therefore, it looks natural to define the Hausdorff dimension as this value.

Definition 7.

The Hausdorff dimension of E is defined as

dimHE:=inf{s0:s(E)=0}=sup{s0:s(E)=}.\dim_{H}\textbf{E}:=\inf\{s\geq 0:\mathcal{H}^{s}(\textbf{E})=0\}=\sup\{s\geq 0:\mathcal{H}^{s}(\textbf{E})=\infty\}.

The following theorem can be found in [5].

Theorem 5.

Let Ω\Omega be a domain in n+1\mathbb{R}^{n+1} and EΩ\textbf{E}\subset\Omega be a compact set. Let be n+μ(E)=0\mathcal{H}^{n+\mu}(\textbf{E})=0 where 0<μ10<\mu\leq 1. If uLip(Ω,μ)u\in{\mbox{Lip}}(\Omega,\mu), and it is monogenic in ΩE\Omega\setminus\textbf{E}, then uu is also monogenic in Ω\Omega.

The following definition of fractal set is due to Mandelbrot [24]

Definition 8.

If an arbitrary set En+1\textbf{E}\subset\mathbb{R}^{n+1} with topological dimension nn has dimHE>n\dim_{H}\textbf{E}>n, then E is called a fractal set.

The results presented in this paper are intended to deal with sets such that n(E)=\mathcal{H}^{n}(\textbf{E})=\infty as well as fractals from Definition 8. Throughout this paper, the expression ‘fractal domain’ will always refer to a domain with a fractal boundary.

The Minkowski dimension is widely used when working with fractals. That is due to the fact that computations are easier than with other fractal dimensions. We will only consider the upper Minkowski dimension.

Definition 9.

Let E be a non-empty bounded subset of n+1\mathbb{R}^{n+1} and let Nδ(E)N_{\delta}(\textbf{E}) be the smallest number of sets of diameter at most δ\delta, covering E. The upper Minkowski dimension of E is defined as

dim¯ME:=lim supδ0logNδ(E)logδ.\overline{\dim}_{M}\textbf{E}:=\limsup_{\delta\rightarrow 0}\dfrac{\log N_{\delta}(\textbf{E})}{-\log\delta}.

We denote 0\mathcal{M}_{0} the grid covering n+1\mathbb{R}^{n+1} which consists of (n+1)(n+1)-dimensional cubes with vertices with integer coordinates and edges of length one. From 0\mathcal{M}_{0} is generated the grid k\mathcal{M}_{k} by dividing every cube in 0\mathcal{M}_{0} into 2(n+1)k2^{(n+1)k} different cubes with edges lengths 2k2^{-k}. Let Nk(E)N_{k}(\textbf{E}) be the amount of cubes of the grid k\mathcal{M}_{k} which intersect E. Then as can be found in [12] we have

dim¯ME=lim supklogNk(E)klog(2).\overline{\dim}_{M}\textbf{E}=\limsup_{k\rightarrow\infty}\dfrac{\log N_{k}(\textbf{E})}{k\log(2)}. (6)

In [25, pp 77] is given the next theorem relating the Hausdorff and Minkowski dimensions.

Theorem 6.

For the bounded set En+1\textbf{E}\subset\mathbb{R}^{n+1} with topological dimension nn, we have

ndimHEdim¯MEn+1.n\leq\dim_{H}\textbf{E}\leq\overline{\dim}_{M}\textbf{E}\leq n+1.

The following new metric characteristics of a fractal set in n+1\mathbb{R}^{n+1} are mainly included to keep the exposition as self-contained as possible. These can be found in [30].

Let 𝒮\mathcal{S} be a topologically compact hypersurface in n+1\mathbb{R}^{n+1}, which bounds a Jordan domain Ω+\Omega^{+}. We write Ω\Omega^{-} for the unbounded complement. It is assumed 𝒮\mathcal{S} to be fractal.

Let Dn+1D\subset\mathbb{R}^{n+1} be a bounded set, which does not touch the hypersurface 𝒮\mathcal{S}. We will consider the integral

Ip(D)=DdV(x)distp(x,𝒮).I_{p}(D)=\int\limits_{D}\dfrac{dV(x)}{\mathrm{dist}^{p}(x,\mathcal{S})}.

For completeness, we recall:

Definition 10.

Let 𝒮\mathcal{S} be a topologically compact hypersurface which is the boundary of a Jordan domain in n+1\mathbb{R}^{n+1}. We define the inner and outer Marcinkiewicz exponent of 𝒮\mathcal{S}, respectively, as

𝔪+(𝒮)=sup{p:Ip(Ω+)<},𝔪(𝒮)=sup{p:Ip(Ω)<},\begin{array}[]{cc}\mathfrak{m}^{+}(\mathcal{S})=\sup\{p:I_{p}(\Omega^{+})<\infty\},&\mathfrak{m}^{-}(\mathcal{S})=\sup\{p:I_{p}(\Omega^{*})<\infty\},\end{array}

and the (absolute) Marcinkiewicz exponent of 𝒮\mathcal{S} as,

𝔪(𝒮)=max{𝔪+(𝒮),𝔪(𝒮)}.\mathfrak{m}(\mathcal{S})=\max\{\mathfrak{m}^{+}(\mathcal{S}),\mathfrak{m}^{-}(\mathcal{S})\}.

Here, the domain Ω:=Ω{x:|x|<r}\Omega^{*}:=\Omega^{-}\bigcap\{x:\arrowvert x\arrowvert<r\}, where rr is chosen in a way that 𝒮\mathcal{S} is completely contained inside the ball of radius rr. We should note that the value of 𝔪(𝒮)\mathfrak{m}^{-}(\mathcal{S}) does not depend on the selection of the radius rr when constructing Ω\Omega^{*}, due to the fact that only the points closest to 𝒮\mathcal{S} influence the convergence of the integral Ip(D)I_{p}(D).

The following theorem expresses the relationship between the Minkowski dimension with the Marcinkiewicz exponent, it was proved in [19] and in a different way in [30].

Theorem 7.

Let 𝒮\mathcal{S} be a topologically compact hypersurface which is the boundary of a Jordan domain in n+1\mathbb{R}^{n+1}, then 𝔪(𝒮)n+1dim¯M(𝒮)\mathfrak{m}(\mathcal{S})\geq n+1-\overline{\dim}_{M}(\mathcal{S}).

3 Riemann Boundary Value Problems

In this section our main results concerning RBVPs for polymonogenic functions on fractal domains using the absolute Marcinkiewicz exponent are stated and proved.

Let fLip(𝒮,k1+ν)f\in{\mbox{Lip}}(\mathcal{S},k-1+\nu) be a 𝒞(n)\mathcal{C}\ell(n)-valued function. We are first interested in the following boundary value problem: To find a polymonogenic function Φ\Phi of order k on n+1𝒮\mathbb{R}^{n+1}\setminus\mathcal{S} continuously extendable from Ω±\Omega^{\pm} to 𝒮\mathcal{S} such that its boundary values Φ±\Phi^{\pm} on 𝒮\mathcal{S} satisfy the following conditions

(𝒟iΦ(x))+(𝒟iΦ(x))=𝐟(i)x𝒮0ik1(𝒟iΦ())=00ik1,\begin{array}[]{ccc}(\mathcal{D}^{i}\Phi(x))^{+}-(\mathcal{D}^{i}\Phi(x))^{-}=\mathbf{f}^{(i)}&x\in\mathcal{S}&0\leq i\leq k-1\\ (\mathcal{D}^{i}\Phi(\infty))^{-}=0&&0\leq i\leq k-1,\end{array} (7)

where the functions f(i)\textbf{f}^{(i)} were defined in (4).

As a special case when k=1k=1 the polymonogenic functions derive to monogenic functions. At the same time, the higher order Lipschitz class Lip(𝒮,k1+ν){\mbox{Lip}}(\mathcal{S},k-1+\nu) becomes the standard Lipschitz class with exponent ν\nu. Consequently, problem (7) reduces the classical jump problem for monogenic functions:

Φ+(x)Φ(x)=f,x𝒮,Φ()=0.\begin{array}[]{ccc}\Phi^{+}(x)-\Phi^{-}(x)=f,&x\in\mathcal{S},\\ \Phi^{-}(\infty)=0.&\end{array} (8)

Hence, problem (7) generalizes problem (8) presented and studied in [30].

The following lemma was proved in [1] and takes part of the proof of the upcoming theorem.

Lemma 1.

Let Ω\Omega be a bounded domain of n+1\mathbb{R}^{n+1} and let gLp(Ω)g\in L^{p}(\Omega) with p>n+1p>n+1. Then,

𝒟iTΩkgLip(n+1,α),\displaystyle\mathcal{D}^{i}T^{k}_{\Omega}g\in{\mbox{Lip}}(\mathbb{R}^{n+1},\alpha), i=0,1,,k1;\displaystyle i=0,1,\cdots,k-1;

with 0<αpn1p0<\alpha\leq\frac{p-n-1}{p}.

The following theorem provides a sufficient solvability condition to problem (7) and generalizes and strengthens [30, Theorem 9].

Theorem 8.

If fLip(𝒮,k1+ν)f\in{\mbox{Lip}}(\mathcal{S},k-1+\nu), with

ν>1𝔪(𝒮)n+1,\displaystyle\nu>1-\dfrac{\mathfrak{m}(\mathcal{S})}{n+1}, (9)

and k<n+1k<n+1, then the problem (7) is solvable.

Proof.

We need to show that the solution is given by

Φ(x)=f~(x)χ+(x)(TΩ+k𝒟kf~)(x),xn+1,\begin{array}[]{cc}\Phi(x)=\widetilde{f}(x)\chi^{+}(x)-(T^{k}_{\Omega^{+}}\mathcal{D}^{k}\widetilde{f})(x),&x\in\mathbb{R}^{n+1},\end{array} (10)

when 𝔪(𝒮)=𝔪+(𝒮)\mathfrak{m}(\mathcal{S})=\mathfrak{m}^{+}(\mathcal{S}), or by

Φ(x)=f(x)χ(x)+(TΩk𝒟kf)(x),\displaystyle\Phi(x)=-f^{*}(x)\chi^{*}(x)+(T^{k}_{\Omega^{*}}\mathcal{D}^{k}f^{*})(x), xn+1,\displaystyle x\in\mathbb{R}^{n+1}, (11)

when 𝔪(𝒮)=𝔪(𝒮)\mathfrak{m}(\mathcal{S})=\mathfrak{m}^{-}(\mathcal{S}).

Here, χ+(x)\chi^{+}(x) and χ(x)\chi^{*}(x) are the characteristic functions of Ω+\Omega^{+} and Ω\Omega^{*} respectively. Besides, f~\widetilde{f} is the Whitney extension of ff and f=f~ρf^{*}=\widetilde{f}\rho where ρ\rho is defined as follows. We will fix r1r_{1} large enough such that 𝒮\mathcal{S} is entirely contained inside the ball B1={x:|x|<r1}B_{1}=\{x:\arrowvert x\arrowvert<r_{1}\}. We choose r>r1r>r_{1}, and define B={z:|x|<r}B=\{z:\arrowvert x\arrowvert<r\}. Due to the fact that the value of 𝔪(𝒮)\mathfrak{m}^{-}(\mathcal{S}) do not depend on the selection of rr then, let be Ω=ΩB\Omega^{*}=\Omega^{-}\bigcap B. Thus, let ρ(x)\rho(x) be a real valued function in C(n+1)C^{\infty}(\mathbb{R}^{n+1}) such that 0ρ(x)10\leq\rho(x)\leq 1, equal to 0 outside of BB, and equal to 1 over B1B_{1}.

The same proof works for 𝔪+(𝒮)\mathfrak{m}^{+}(\mathcal{S}) and 𝔪(𝒮)\mathfrak{m}^{-}(\mathcal{S}), we will consider the first case.

We must have that 𝒟kf~\mathcal{D}^{k}\widetilde{f}\in L(Ω+)p{}^{p}(\Omega^{+}) with p>n+1p>n+1, being f~\widetilde{f} the Whitney extension of ff. Theorem 4 now show that

Ω+|𝒟kf~(x)|p𝑑V(x)CΩ+dV(x)dist(x,𝒮)p(1ν).\int\limits_{\Omega^{+}}\arrowvert\mathcal{D}^{k}\widetilde{f}(x)\arrowvert^{p}dV(x)\leq C\int\limits_{\Omega^{+}}\dfrac{dV(x)}{\mathrm{dist}(x,\mathcal{S})^{p(1-\nu)}}.

The right-hand integral above converges for p<𝔪+(𝒮)1νp<\frac{\mathfrak{m}^{+}(\mathcal{S})}{1-\nu}, which is a direct consequence of Definition 10,. Then the main requirement is

ν>1𝔪+(𝒮)n+1.\nu>1-\dfrac{\mathfrak{m}^{+}(\mathcal{S})}{n+1}.

From (1) it follows that Φ\Phi is a polymonogenic function of order kk on n+1𝒮\mathbb{R}^{n+1}\setminus\mathcal{S}. Combining Lemma 1 with the fact that f~Lip(n+1,k1+ν)\widetilde{f}\in{\mbox{Lip}}(\mathbb{R}^{n+1},k-1+\nu) we obtain that the functions 𝒟iΦ\mathcal{D}^{i}\Phi, i=0,1,,k1i=0,1,\dots,k-1; are continuous functions on Ω+¯\overline{\Omega^{+}} and Ω¯\overline{\Omega^{-}}.

Combining Lemma 1 with Theorem 4, we can conclude that the function Φ(x)\Phi(x) satisfies the boundary condition over 𝒮\mathcal{S}. Finally, as was stated in [1] when k<n+1k<n+1, we have that 𝒟iΦ\mathcal{D}^{i}\Phi^{-} vanishes at infinity for every i=0,1,,k1i=0,1,\cdots,k-1. A trivial verification shows that 𝒟kfLp(Ω)\mathcal{D}^{k}f^{*}\in L^{p}(\Omega^{*}) providing that ν>1𝔪(𝒮)n+1\nu>1-\frac{\mathfrak{m}^{-}(\mathcal{S})}{n+1}. This completes the proof. ∎

We can also prove a sufficient condition for unicity. The next theorem is a generalization of [30, Theorem 10].

Theorem 9.

Let be fLip(𝒮,k1+ν)f\in{\mbox{Lip}}(\mathcal{S},k-1+\nu) with ν>1𝔪(𝒮)n+1\nu>1-\dfrac{\mathfrak{m}(\mathcal{S})}{n+1} and k<n+1k<n+1, let

dimH𝒮n<μ<1(n+1)(1ν)𝔪(𝒮).\dim_{H}\mathcal{S}-n<\mu<1-\dfrac{(n+1)(1-\nu)}{\mathfrak{m}(\mathcal{S})}.

Then there is a unique solution Φ\Phi of the problem (7), such that 𝒟iΦ\mathcal{D}^{i}\Phi belongs to the classes Lip(Ω+¯,μ){\mbox{Lip}}(\overline{\Omega^{+}},\mu) and Lip(Ω¯,μ){\mbox{Lip}}(\overline{\Omega^{-}},\mu), for i=0,1,,k1i=0,1,\cdots,k-1.

Proof.

From Lemma 1 and the proof of Theorem 8 we deduce that the solution Φ\Phi to the problem (7), defined by (10) or (11), belongs to Lip(Ω+¯,μ){\mbox{Lip}}(\overline{\Omega^{+}},\mu) and Lip(Ω¯,μ){\mbox{Lip}}(\overline{\Omega^{-}},\mu) for μ<1(n+1)(1ν)𝔪(𝒮)\mu<1-\frac{(n+1)(1-\nu)}{\mathfrak{m}(\mathcal{S})}.

Now, let us suppose that there exist two solutions Φ1\Phi_{1} and Φ2\Phi_{2} to the problem (7), and define Φ:=Φ2Φ1\Phi:=\Phi_{2}-\Phi_{1}. This function is a solution to the homogeneous problem

(𝒟iΦ(x))+(𝒟iΦ(x))=0x𝒮0ik1(𝒟iΦ())=00ik1.\begin{array}[]{ccc}(\mathcal{D}^{i}\Phi(x))^{+}-(\mathcal{D}^{i}\Phi(x))^{-}=0&x\in\mathcal{S}&0\leq i\leq k-1\\ (\mathcal{D}^{i}\Phi(\infty))^{-}=0&&0\leq i\leq k-1.\end{array} (12)

We shall prove that Φ0\Phi\equiv 0 is the unique solution to this problem such that 𝒟iΦ\mathcal{D}^{i}\Phi belongs to the classes Lip(Ω+¯,μ){\mbox{Lip}}(\overline{\Omega^{+}},\mu) and Lip(Ω¯,μ){\mbox{Lip}}(\overline{\Omega^{-}},\mu), for i=0,1,,k1i=0,1,\cdots,k-1. The proof is carried out by induction on kk, by a repeated application of [30, Theorem 10].

Now we assume that (7) has the unique solution Φ0\Phi\equiv 0 such that 𝒟iΦ\mathcal{D}^{i}\Phi belongs to the classes Lip(Ω+¯,μ){\mbox{Lip}}(\overline{\Omega^{+}},\mu) and Lip(Ω¯,μ){\mbox{Lip}}(\overline{\Omega^{-}},\mu) for i=0,1,,k1i=0,1,\cdots,k-1; for k=lk=l, and let us consider the problem for k=l+1k=l+1

(𝒟iΦ(x))+(𝒟iΦ(x))=0x𝒮0il(𝒟iΦ())=00il.\begin{array}[]{ccc}(\mathcal{D}^{i}\Phi(x))^{+}-(\mathcal{D}^{i}\Phi(x))^{-}=0&x\in\mathcal{S}&0\leq i\leq l\\ (\mathcal{D}^{i}\Phi(\infty))^{-}=0&&0\leq i\leq l.\end{array} (13)

Let Φ\Phi be a solution of (13). If we denote Ψ:=𝒟Φ\Psi:=\mathcal{D}\Phi, then 𝒟lΨ:=𝒟l+1Φ=0\mathcal{D}^{l}\Psi:=\mathcal{D}^{l+1}\Phi=0 in n+1𝒮\mathbb{R}^{n+1}\setminus\mathcal{S} and

(𝒟iΨ(x))+(𝒟iΨ(x))=0x𝒮0il1(𝒟iΨ())=00il1.\begin{array}[]{ccc}(\mathcal{D}^{i}\Psi(x))^{+}-(\mathcal{D}^{i}\Psi(x))^{-}=0&x\in\mathcal{S}&0\leq i\leq l-1\\ (\mathcal{D}^{i}\Psi(\infty))^{-}=0&&0\leq i\leq l-1.\end{array}

Consequently, Ψ\Psi represents a solution of (12) with k=lk=l. Then, by the induction hypothesis, Ψ0\Psi\equiv 0 is the only solution in this class. As a result 𝒟Φ=0\mathcal{D}\Phi=0 in n+1𝒮\mathbb{R}^{n+1}\setminus\mathcal{S}, and

Φ(x)+Φ(x)=0x𝒮Φ()=0.\begin{array}[]{ccc}\Phi(x)^{+}-\Phi(x)^{-}=0&x\in\mathcal{S}\\ \Phi(\infty)^{-}=0&.\end{array}

Therefore, as in the proof of [30, Theorem 10], we have Φ0\Phi\equiv 0 in n+1\mathbb{R}^{n+1}, and the proof is complete. ∎

4 A Class of hypersurfaces in n+1\mathbb{R}^{n+1}

A new solvability condition for RBVPs for polymonogenic functions via a high-dimensional Marcinkiewicz exponent has been proved. However, there naturally arises the question of whether this condition improves those involving the Minkowski dimension presented in [1]? As a matter of fact, Theorem 7 is sufficient to guarantee that new condition does not ever be worth the effort of formulating them based on the Minkowski dimension. Indeed, a class of hypersurfaces in three-dimensional spaces so constructed in [30] shows that strict inequality can occur in Theorem 7, what is an extension to the case of the complex plane given in [20].

In this section, we generalize these constructions to a class of hypersurfaces in the (n+1)(n+1)-dimensional space. Summarizing we have.

Theorem 10.

Let α1\alpha\geq 1 and βn\beta\geq n. For each value d(n,n+1)d\in(n,n+1), there exists a uncountable collection of topologically compact hypersurfaces 𝒮α,βn+1\mathcal{S}^{n+1}_{\alpha,\beta}, which are the boundary of a Jordan domain in n+1\mathbb{R}^{n+1} such that d=dim¯M(𝒮α,βn+1)d=\overline{\dim}_{M}(\mathcal{S}^{n+1}_{\alpha,\beta}) and 𝔪(𝒮α,βn+1)>(n+1)dim¯M(𝒮α,βn+1)\mathfrak{m}(\mathcal{S}^{n+1}_{\alpha,\beta})>(n+1)-\overline{\dim}_{M}(\mathcal{S}^{n+1}_{\alpha,\beta}) for suitable values of α\alpha and β\beta.

Proof.

The proof consists in the construction of hypersurfaces 𝒮α,βn+1\mathcal{S}^{n+1}_{\alpha,\beta} having the desired properties.

Let Q=[0,1]×[0,1]×[0,1]××[1,0]Q=[0,1]\times[0,1]\times[0,1]\times\cdots\times[-1,0] be a (n+1)(n+1)-dimensional cube. We will add infinitely many (n+1)(n+1)-dimensional rectangles with suitable dimensions to this cube. In order to do that, let us fix α1\alpha\geq 1 and βn\beta\geq n. Initially, we will conveniently divide the segment [0,1][0,1] in the x0x_{0} axis. We break down it into the sub-segments [2m,2m+1][2^{-m},2^{-m+1}] for every mm\in\mathbb{N}, and we divide each of this sub-segments into 2[mβ]2^{[m\beta]} equally spaced segments where [mβ][m\beta] is the integer part of mβm\beta. The endpoints at the right side of these segments will be denoted by ymjy_{mj}, where j=1,2,,2[mβ]j=1,2,...,2^{[m\beta]}. In addition, let am=2m[mβ]a_{m}=2^{-m-[m\beta]} be the distance between two consecutive points ymjy_{mj} and ym(j+1)y_{m(j+1)}, and Cm=12amαC_{m}=\frac{1}{2}a_{m}^{\alpha}. Then, let RmjR_{mj} be the (n+1)(n+1)-dimensional rectangles defined as

Rmj=[ymjCm,ymj]×[0,2m]××[0,2m].R_{mj}=[y_{mj}-C_{m},y_{mj}]\times[0,2^{-m}]\times\cdots\times[0,2^{-m}].

Hence, we define

Tα,βn+1:=Q(m=1j=12[mβ]Rmj).T^{n+1}_{\alpha,\beta}:=Q\bigcup\left(\bigcup_{m=1}^{\infty}\bigcup_{j=1}^{2^{[m\beta]}}R_{mj}\right).

The claimed hypersurfaces 𝒮α,βn+1\mathcal{S}^{n+1}_{\alpha,\beta} are the boundaries of the corresponding Tα,βn+1T^{n+1}_{\alpha,\beta}.

We first obtain a suitable lower bound on the Marcinkiewicz exponent of hypersurfaces 𝒮α,βn+1\mathcal{S}^{n+1}_{\alpha,\beta}. To do this, split up the integral as follows

Ω+dVdistp(x,𝒮α,βn+1)=QdVdistp(x,𝒮α,βn+1)+m=1j=12[mβ]RmjdVdistp(x,𝒮α,βn+1).\int\limits_{\Omega^{+}}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}=\int\limits_{Q}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}+\sum_{m=1}^{\infty}\sum_{j=1}^{2^{[m\beta]}}\int\limits_{R_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}. (14)

Let Lmj′′L^{{}^{\prime\prime}}_{mj} be the subset of RmjR_{mj} such that dist(x,𝒮α,β)=dist(x,Γmj′′)\mathrm{dist}(x,\mathcal{S}_{\alpha,\beta})=\mathrm{dist}(x,\Gamma^{{}^{\prime\prime}}_{mj}) where Γmj′′={ymj}×[0,2m]××[0,2m]\Gamma^{{}^{\prime\prime}}_{mj}=\left\{y_{mj}\right\}\times[0,2^{-m}]\times\cdots\times[0,2^{-m}], then

RmjdVdistp(x,𝒮α,βn+1)=RmjLmj′′dVdistp(x,𝒮α,βn+1)+Lmj′′dVdistp(x,𝒮α,βn+1).\int\limits_{R_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}=\int\limits_{R_{mj}\setminus L^{{}^{\prime\prime}}_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}+\int\limits_{L^{{}^{\prime\prime}}_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}.

and

Lmj′′dVdistp(x,𝒮α,βn+1)=Lmj′′dVdistp(x,Γmj′′).\displaystyle\int\limits_{L^{{}^{\prime\prime}}_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\mathcal{S}^{n+1}_{\alpha,\beta})}=\int\limits_{L^{{}^{\prime\prime}}_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\Gamma^{{}^{\prime\prime}}_{mj})}.

Hence, using the estimate

Lmj′′dVdistp(x,Γmj′′)=Lmj′′dV|ymjx0|pRmjdV|ymjx1|p.\displaystyle\int\limits_{L^{{}^{\prime\prime}}_{mj}}\dfrac{dV}{\mathrm{dist}^{p}(x,\Gamma^{{}^{\prime\prime}}_{mj})}=\int\limits_{L^{{}^{\prime\prime}}_{mj}}\dfrac{dV}{\arrowvert y_{mj}-x_{0}\arrowvert^{p}}\leq\int\limits_{R_{mj}}\dfrac{dV}{\arrowvert y_{mj}-x_{1}\arrowvert^{p}}.

and after some straightforward computations, we obtain that if the sum

m=12mβnm(1p)α(m+mβ),\sum_{m=1}^{\infty}2^{m\beta-nm-(1-p)\alpha(m+m\beta)}, (15)

converges, then so is the integral in (14). This geometric sum (15) converges if and only if the condition

p<1βnα(β+1),p<1-\dfrac{\beta-n}{\alpha(\beta+1)},

is fulfilled. Thus, the following estimation for the inner Marcinkiewicz exponent holds

𝔪+(𝒮α,βn+1):=sup{p>0:Ip(G+)<}1βnα(β+1).\mathfrak{m}^{+}(\mathcal{S}^{n+1}_{\alpha,\beta}):=\sup\{p>0:I_{p}(G^{+})<\infty\}\geq 1-\dfrac{\beta-n}{\alpha(\beta+1)}.

We have proved more, namely that the absolute Marcinkiewicz exponent satisfies

𝔪(𝒮α,βn+1):=max{𝔪+(𝒮α,βn+1),𝔪(𝒮α,βn+1)}𝔪+(𝒮α,βn+1)1βnα(β+1).\mathfrak{m}(\mathcal{S}^{n+1}_{\alpha,\beta}):=\max\{\mathfrak{m}^{+}(\mathcal{S}^{n+1}_{\alpha,\beta}),\mathfrak{m}^{-}(\mathcal{S}^{n+1}_{\alpha,\beta})\}\geq\mathfrak{m}^{+}(\mathcal{S}^{n+1}_{\alpha,\beta})\geq 1-\dfrac{\beta-n}{\alpha(\beta+1)}.

We proceed to calculate the value of the Minkowski dimension of hypersurfaces 𝒮α,βn+1\mathcal{S}^{n+1}_{\alpha,\beta}. It is sufficient to show that its lower and upper bounds are the same.

First, look for a suitable upper bound for dim¯M(𝒮α,βn+1)\overline{\dim}_{M}(\mathcal{S}^{n+1}_{\alpha,\beta}). We define the sets Λm:=j=12[mβ][Rmj(Rmj)|xn=0]\Lambda_{m}:=\bigcup\limits_{j=1}^{2^{[m\beta]}}[\partial R_{mj}\setminus(\partial R_{mj})\arrowvert_{x_{n}=0}] and Λ:=m=1Λm\Lambda:=\bigcup\limits_{m=1}^{\infty}\Lambda_{m}. Furthermore, Q^:=Q[m=1j=12[mβ](Rmj)|xn=0]\widehat{Q}:=\partial Q\setminus[\bigcup\limits_{m=1}^{\infty}\bigcup\limits_{j=1}^{2^{[m\beta]}}(\partial R_{mj})\arrowvert_{x_{n}=0}], we can see that 𝒮α,βn+1=Q^Λ\mathcal{S}^{n+1}_{\alpha,\beta}=\widehat{Q}\cup\Lambda.

The following step makes use of the grid k\mathcal{M}_{k} defined in Subsection 2.3. Initially, note that with 2(n+1)(12k)n2(n+1)\left(\frac{1}{2^{-k}}\right)^{n} cubes of k\mathcal{M}_{k}, we can cover Q^\widehat{Q}. In order to study Λ\Lambda, we need to consider three cases. The first case is if m<km<k and Cm>2kC_{m}>2^{-k}, then 2[mβ]+1(2m2k)n2^{[m\beta]+1}\left(\frac{2^{-m}}{2^{-k}}\right)^{n} cubes have to be used to cover the faces of the (n+1)(n+1)-dimensional rectangles RmjR_{mj} parallel to x0=0x_{0}=0, in Λm\Lambda_{m}. No more than 2(2m2k)n2\left(\frac{2^{-m}}{2^{-k}}\right)^{n} cubes are required to cover the nn-dimensional rectangles in Λm\Lambda_{m} parallel to the coordinate plane xl=0x_{l}=0, for every l=1,2,,nl=1,2,\cdots,n.

The second, and main case, is if m<km<k and Cm2kC_{m}\leq 2^{-k}. Here we must study two more cases. When Cm2kC_{m}\leq 2^{-k}, k>mk>m, and also amCm>2ka_{m}-C_{m}>2^{-k}, thus analogously than in the previous step 2[mβ](2m2k)n2^{[m\beta]}(\frac{2^{-m}}{2^{-k}})^{n} cubes are sufficient to cover the faces of RmjR_{mj}’s parallel to x0=0x_{0}=0 in Λm\Lambda_{m}. While, to cover the nn-dimensional rectangles in Λm\Lambda_{m} parallel to the coordinate plane xl=0x_{l}=0 are not required more than 2(2m2k)n2\left(\frac{2^{-m}}{2^{-k}}\right)^{n} cubes of k\mathcal{M}_{k}, for every l=1,2,,nl=1,2,\cdots,n
If Cm2kC_{m}\leq 2^{-k}, k>mk>m, and amCm2ka_{m}-C_{m}\leq 2^{-k}, hence (2m2k)n+1\left(\frac{2^{-m}}{2^{-k}}\right)^{n+1} cubes in k\mathcal{M}_{k} are adequate to cover Λm\Lambda_{m}.

Finally, when mkm\geq k, by definition, the hypersurfaces Λm\Lambda_{m}, with m>km>k, are all covered by only one cube of the grid k\mathcal{M}_{k}. While the hypersurface Λk\Lambda_{k} is covered by another of these cubes. Therefore, we obtain

Nk(𝒮α,βn+1)2+2(n+1)2nk+2Cm>2k,k>m2[mβ]+nknm+2nCm>2k,k>m2nknm+\displaystyle N_{k}(\mathcal{S}^{n+1}_{\alpha,\beta})\leq 2+2(n+1)\cdotp 2^{nk}+2\sum\limits_{C_{m}>2^{-k},\,\,\,k>m}2^{[m\beta]+nk-nm}+2n\sum\limits_{C_{m}>2^{-k},\,\,\,k>m}2^{nk-nm}+
+Cm2k,amCm2k,k>m2(n+1)k(n+1)m+Cm2k<amCm,k>m2[mβ]+nknm+\displaystyle+\sum\limits_{C_{m}\leq 2^{-k},\,\,\,a_{m}-C_{m}\leq 2^{-k},\,\,\,k>m}2^{(n+1)k-(n+1)m}+\sum\limits_{C_{m}\leq 2^{-k}<a_{m}-C_{m},\,\,\,k>m}2^{[m\beta]+nk-nm}+
+2nCm2k<amCmk>m2nknm.+2n\sum\limits_{C_{m}\leq 2^{-k}<a_{m}-C_{m}\,\,\,k>m}2^{nk-nm}.

Working with the conditions on the sums in the previous inequality, we are able to obtain the following greater estimates

Nk(𝒮α,βn+1)2+2(n+1)2nk+32k<am,k>m2[mβ]+nknm+4n2k<am,k>m2nknm+N_{k}(\mathcal{S}^{n+1}_{\alpha,\beta})\leq 2+2(n+1)\cdotp 2^{nk}+3\sum\limits_{2^{-k}<a_{m},\,\,\,k>m}2^{[m\beta]+nk-nm}+4n\sum\limits_{2^{-k}<a_{m},\,\,\,k>m}2^{nk-nm}+
+am22k,k>m2(n+1)k(n+1)m.+\sum\limits_{\frac{a_{m}}{2}\leq 2^{-k},\,\,\,k>m}2^{(n+1)k-(n+1)m}.

We will denote by BkB_{k} and HkH_{k} the integers defined by the conditions

k1+β1Bk<k1+β,\displaystyle\dfrac{k}{1+\beta}-1\leq B_{k}<\dfrac{k}{1+\beta}, (16)
k11+β1Hk<k11+β.\displaystyle\dfrac{k-1}{1+\beta}-1\leq H_{k}<\dfrac{k-1}{1+\beta}. (17)

It is easy to check that the condition am>2ka_{m}>2^{-k} is satisfied if and only if m=1,2,,Bkm=1,2,...,B_{k}. By taking into account HkH_{k} for the sum under the conditions am22k,k>m\frac{a_{m}}{2}\leq 2^{-k},k>m; and BkB_{k} for those under the conditions 2k<am,k>m2^{-k}<a_{m},k>m , we get through some estimates the following inequality

Nk(𝒮α,βn+1)D(k)2(n+1)kββ+1,N_{k}(\mathcal{S}^{n+1}_{\alpha,\beta})\leq D(k)2^{\frac{(n+1)k\beta}{\beta+1}},

where D(k)=ak+cD(k)=ak+c; here aa and cc only depend on β\beta and nn. Then,

dim¯M(𝒮α,βn+1)(n+1)ββ+1.\overline{\dim}_{M}(\mathcal{S}^{n+1}_{\alpha,\beta})\leq\dfrac{(n+1)\beta}{\beta+1}.

Now we will compute a lower bound. In order to do that, we will build a set AβA_{\beta} such that Aβ𝒮α,βn+1A_{\beta}\subset\mathcal{S}^{n+1}_{\alpha,\beta} and therefore dim¯M(Aβ)dim¯M(𝒮α,βn+1)\overline{\dim}_{M}(A_{\beta})\leq\overline{\dim}_{M}(\mathcal{S}^{n+1}_{\alpha,\beta}).
We will present the auxiliary nn-dimensional rectangles PmjP_{mj}

Pmj={ymj}×[0,2m]××[0,2m],P_{mj}=\{y_{mj}\}\times[0,2^{-m}]\times\cdots\times[0,2^{-m}],

and the set AβA_{\beta} is given by

Aβ=m=1j=12[mβ]Pmj.A_{\beta}=\bigcup_{m=1}^{\infty}\bigcup_{j=1}^{2^{[m\beta]}}P_{mj}.

Observe, by construction, that Aβ𝒮α,βA_{\beta}\subset\mathcal{S}_{\alpha,\beta}. The task is now to find a lower bound on dim¯M(Aβ)\overline{\dim}_{M}(A_{\beta}).

The distance between PmjP_{mj} and Pmj+1P_{mj+1} is am=2m[mβ]a_{m}=2^{-m-[m\beta]}. If k>mk>m, and am>2ka_{m}>2^{-k}, then two of these rectangles can not be intersected by the same cube in the grid k\mathcal{M}_{k}. Then, (2m2k)n(\frac{2^{-m}}{2^{-k}})^{n} cubes in k\mathcal{M}_{k} cover a single nn-dimensional rectangle PmjP_{mj}, due to the fact that the lengths of each edge in these nn-dimensional rectangles is 2m2^{-m}.

Therefore, the total number cubes required to cover the 2[mβ]2^{[m\beta]} rectangles PmjP_{mj} for a fixed mm is 2[mβ](2m2k)n2^{[m\beta]}(\frac{2^{-m}}{2^{-k}})^{n}. Thus we get

Nk(Aβ)2am>2k,k>m2[mβ]+nknm,N_{k}(A_{\beta})\geq 2\cdotp\sum\limits_{a_{m}>2^{-k},\,\,\,k>m}2^{[m\beta]+nk-nm},

where Nk(Aβ)N_{k}(A_{\beta}) is the smallest amount of cubes in k\mathcal{M}_{k} which cover AβA_{\beta}.
Being BkB_{k} as in (16) we get

am>2k,k>m2[mβ]+nknm=2nkn=1Bk2[mβ]nm2nk1n=1Bk2m(βn)C2(n+1)kββ+1,\sum\limits_{a_{m}>2^{-k},\,\,\,k>m}2^{[m\beta]+nk-nm}=2^{nk}\sum\limits_{n=1}^{B_{k}}2^{[m\beta]-nm}\geq 2^{nk-1}\sum\limits_{n=1}^{B_{k}}2^{m(\beta-n)}\geq C2^{\frac{(n+1)k\beta}{\beta+1}},

where CC does not depend on kk. Hence

dim¯M(Sα,βn+1)dim¯M(Aβ)(n+1)ββ+1.\overline{\dim}_{M}(S^{n+1}_{\alpha,\beta})\geq\overline{\dim}_{M}(A_{\beta})\geq\frac{(n+1)\beta}{\beta+1}.

Therefore,

dim¯M(Sα,βn+1)=(n+1)ββ+1.\overline{\dim}_{M}(S^{n+1}_{\alpha,\beta})=\frac{(n+1)\beta}{\beta+1}.

Having disposed the value of the Minkowski dimension and a lower bound on the Marcinkiewicz exponent of every hypersurface Sα,βn+1S^{n+1}_{\alpha,\beta}, we are in a position to finish the proof of Theorem 10.

Indeed, when α>1\alpha>1 and β>n\beta>n we have

𝔪(𝒮α,β)𝔪+(𝒮α,β)1βnα(β+1)>1βnβ+1=(n+1)(n+1)ββ+1=(n+1)dim¯M(𝒮α,β).\mathfrak{m}(\mathcal{S}_{\alpha,\beta})\geq\mathfrak{m}^{+}(\mathcal{S}_{\alpha,\beta})\geq 1-\dfrac{\beta-n}{\alpha(\beta+1)}>1-\dfrac{\beta-n}{\beta+1}=(n+1)-\dfrac{(n+1)\beta}{\beta+1}=(n+1)-\overline{\dim}_{M}(\mathcal{S}_{\alpha,\beta}).

Setting β=d(n+1)d\beta=\displaystyle\frac{d}{(n+1)-d}, for every d(n,n+1)d\in(n,n+1), we obtain dim¯M(𝒮α,β)=d\overline{\dim}_{M}(\mathcal{S}_{\alpha,\beta})=d for each α>1\alpha>1. This means that {Sα,βn+1}\{S^{n+1}_{\alpha,\beta}\} is an uncountable family, which completes the proof. ∎

Funding: C. D. Tamayo Castro gratefully acknowledges the financial support of the Postgraduate Study Fellowship of the Consejo Nacional de Ciencia y Tecnología (CONACYT) (Grant Number 957110). J. Bory Reyes was partially supported by Instituto Politécnico Nacional in the framework of SIP programs (SIP20230312).

References

  • [1] R. Abreu Blaya, R. Ávila Ávila, and J. Bory Reyes. Boundary value problems with higher order Lipschitz boundary data for polymonogenic functions in fractal domains. Appl. Math. Comput., 269:802–808, 2015.
  • [2] R. Abreu Blaya and J. Bory Reyes. On the Riemann-Hilbert type problems in Clifford analysis. Adv. Appl. Clifford Algebr., 11(1):15–26, 2001.
  • [3] R. Abreu Blaya, J. Bory Reyes, and B. A. Kats. Integration over non-rectifiable curves and Riemann boundary value problems. J. Math. Anal. Appl., 380(1):177–187, 2011.
  • [4] R. Abreu Blaya, J. Bory Reyes, and B. A. Kats. On the solvability of the jump problem in Clifford analysis. J. Math. Sci., 189(1):1–9, 2013.
  • [5] R. Abreu Blaya, D. Peña Peña, and J. Bory Reyes. Jump problem and removable singularities for monogenic functions. J. Geom. Anal., 17(1):1–13, 2007.
  • [6] H. Begehr. Iterated integral operators in Clifford analysis. Z. Anal. Anwend, 18 (2)(2):361–377, 1999.
  • [7] S. Bernstein. Riemann-Hilbert problems in Clifford analysis. In Clifford analysis and its applications (Prague, 2000), volume 25 of NATO Sci. Ser. II Math. Phys. Chem., pages 1–8. Kluwer Acad. Publ., Dordrecht, 2001.
  • [8] F. Brackx. On (k)-monogenic functions of a quaternion variable. In R. P. Gilbert and R. J. Weinacht, editors, Function theoretic methods in differential equations, Research Notes in Mathematics 8. Pitman Publishers, 1976.
  • [9] F. Brackx, R. Delanghe, and F. Sommen. Clifford Analysis. Chapman & Hall/CRC research notes in mathematics series. Pitman Advanced Pub. Program, 1982.
  • [10] P. Cerejeiras, U. Kähler, and M. Ku. On the Riemann boundary value problem for null solutions to iterated generalized Cauchy–Riemann operator in Clifford analysis. Results Math., 63(3-4):1375–1394, 2012.
  • [11] R. Delanghe and F. Brackx. Hypercomplex function theory and Hilbert modules with reproducing kernel. Proc. London Math. Soc. (3), 37(3):545–576, 1978.
  • [12] K. Falconer. Fractal Geometry: Mathematical Foundations and Applications. Wiley, 2003.
  • [13] F. D. Gakhov. Boundary value problems. Addison-Wesley Publishing Co., 1966.
  • [14] K. Gürlebeck, K. Habetha, and W. Sprössig. Holomorphic Functions in the Plane and n-dimensional Space. Birkhäuser Basel, 2008.
  • [15] B. A. Kats. The Riemann problem on a closed Jordan curve. Izv. Vyssh. Uchebn. Zaved. Mat., 251(4):68–80, 1983.
  • [16] B. A. Kats. On solvability of the jump problem. J. Math. Anal. Appl., 356(2):577–581, 2009.
  • [17] B. A. Kats. The Riemann boundary value problem on non-rectifiable curves and related questions. Complex Var. Elliptic Equ., 59(8):1053–1069, 2014.
  • [18] B. A. Kats and D. B. Katz. Marcinkiewicz exponents and integrals over non-rectifiable paths. Math. Methods Appl. Sci., 39(12):3402–3410, 2016.
  • [19] D. B. Katz. Local and weighted Marcinkiewicz exponents with applications. J. Math. Anal. Appl., 440(1):74–85, 2016.
  • [20] D. B. Katz. New metric characteristics of nonrectifiable curves and their applications. Sib. Math. J., 57(2):285–291, 2016.
  • [21] J Le and D. Jinyuan. Riemann boundary value problems for some kk-regular functions in Clifford analysis. Acta Math. Sci. Ser. B (Engl. Ed.), 32(5):2029–2049, 2012.
  • [22] W. Liping. Some properties of a kind of generalized Teodorescu operator in Clifford analysis. J. Inequal. Appl., pages Paper No. 102, 11, 2016.
  • [23] J. K. Lu. Boundary Value Problems for Analytic Functions. Series in pure mathematics. World Scientific Publish., 1993.
  • [24] B. B. Mandelbrot. The Fractal Geometry of Nature. W. H. Freeman and Co., 1982.
  • [25] P. Mattila. Geometry of Sets and Measures in Euclidean Spaces: Fractals and Rectifiability. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1995.
  • [26] K. Min, U. Kähler, and W. Daoshun. Riemann boundary value problems on the sphere in Clifford analysis. Adv. Appl. Clifford Algebr., 22(2):365–390, 2012.
  • [27] N. I. Muskhelishvili. Singular Integral Equations. Boundary problems of function theory and their application to mathematical physics. Springer Netherlands, 1953.
  • [28] John Ryan, editor. Clifford algebras in analysis and related topics, Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1996.
  • [29] E. M. Stein. Singular Integrals and Differentiability Properties of Functions. Monographs in harmonic analysis. Princeton University Press, 1970.
  • [30] C. D. Tamayo Castro. Marcinkiewicz exponent and boundary value problems in fractal domains of n+1\mathbb{R}^{n+1}. Preprint, 2022.
  • [31] C. D. Tamayo Castro, R. Abreu Blaya, and J. Bory Reyes. Compactness of embedding of generalized higher order Lipschitz classes. Anal. Math. Phys., 9(4):1719–1727, 2018.
  • [32] B. Yude and D. Jin Yuan. The Riemann boundary value problem for kk-monogenic functions in Clifford analysis. Acta Math. Sci. Ser. A (Chinese Ed.), 29(5):1321–1330, 2009.