1 Introduction

In this paper, we are interested on the study of the lower t-dimensional Hewitt–Stromberg measure \({\mathsf {H}}^t\) and the upper t-dimensional Hewitt–Stromberg measure \({\mathsf {P}}^t\). More precisely, we are trying to answer the question : for E and F two subsets of \(\mathbb {R}^d\) and \(\mathbb {R}^l\), does there exist a constant \(\gamma \) such that \(\Lambda ^{s+t}(E\times F) =\gamma \Lambda ^s(E) \Lambda ^t(F)\) ? where \(\Lambda \in \{{\mathsf {H}}, {\mathsf {P}}\}\). In fact, such results are far from being true. For example, in [8], it is proved that,

$$\begin{aligned} {\mathsf {b}}(E\times F) \ge {\mathsf {b}}(E) + {\mathsf {b}}(F) \end{aligned}$$
(1.1)

where \({\mathsf {b}}\) stands for the lower Hewitt–Stomberg dimension. However, it is possible to construct two subsets E and F such that \({\mathsf {b}}(E\times F) > {\mathsf {b}}(E) + {\mathsf {b}}(F)\). Similar results were proved for s-dimensional Hausdorff measure \(\mathcal {H}^s\) and s-dimensional packing measure \(\mathcal {P}^s\) [2, 15, 16, 31]. The reader can see also [14, 25, 32, 33] for various results on this problem.

The aim of this paper is to prove Theorem A and C below. This is by constructing a metric outer measure, denoted \({\mathsf {H}}^{*t}\), that is comparable to the Hewitt Stromberg measure \({\mathsf {H}}^{t}\) but significantly easier to analyse.

Theorem A

Let \(E\subseteq \mathbb {R}^d\), \(F\subseteq \mathbb {R}^l\) and \(s,t \ge 0\). There exists a positive constant c such that

$$\begin{aligned} {\mathsf {H}} ^{t+s}(E\times F)\le & {} c\; {\mathsf {H}} ^s(E) {\mathsf {P}} ^t(F) \end{aligned}$$
(1.2)
$$\begin{aligned} {\mathsf {H}} ^s(E) {\mathsf {P}} ^t(F)\le & {} c\; {\mathsf {P}} ^{t+s}(E\times F)\end{aligned}$$
(1.3)
$$\begin{aligned} {\mathsf {P}} ^{t+s}(E\times F)\le & {} c\; {\mathsf {P}} ^s(E) {\mathsf {P}} ^t(F). \end{aligned}$$
(1.4)

To prove (1.1), the authors in [8] prove that

$$\begin{aligned} {\mathsf {H}}^t(E) {\mathsf {H}}^s(F)\le c \, {\mathsf {H}}^{t+s}(E\times F). \end{aligned}$$

The key idea of the proof is to construct a new multifractal measure, on the Euclidean space \(\mathbb {R}^d\), equivalent to the lower Hewitt–Stomberg. Therefore, the authors use the class of all half-open dyadic cubes of covering sets in the definition rather than the class of all closed balls. In this paper, we will use the class of all half-open semi-dyadic cubes rather than the dyadic ones to construct two measures similar to \({\mathsf {H}}^t\) and \({\mathsf {P}}^t\). This choice is due to the fact that the semi-dyadic cube \(v_n(x)\) is less sensitive to the location of x than the corresponding dyadic cube \(u_n(x)\). Semi-dyadic cubes have been used in [15, 26, 35]. A direct consequence of Theorem A is the following result.

Corollary B

For all \(E, F\subseteq \mathbb {R}^d\) we have

$$\begin{aligned} {\mathsf {b}} (E) + {\mathsf {b}} (F) \le {\mathsf {b}} (E\times F)\le & {} {\mathsf {b}} (E) + {\mathsf {B}} (F) \end{aligned}$$
(1.5)
$$\begin{aligned}\le & {} {\mathsf {B}} (E\times F) \end{aligned}$$
(1.6)
$$\begin{aligned}\le & {} {\mathsf {B}} (E) + {\mathsf {B}} (F). \end{aligned}$$
(1.7)

where \({\mathsf {B}} \) (resp. \({\mathsf {b}} \))stands for the upper (resp. lower) Hewitt–Stomberg dimension.

The set E is said to be regular if \({\mathsf {b}}(E) = {\mathsf {B}}(E)\). Regular sets are defined with respect to the Hausdorff measure [3, 4, 17, 18], to packing measure [34, 35] or to Hewitt–Stromberg measure [20,21,22]. In particular Tricot et al. [29, 34] managed to show that a subset E of \(\mathbb {R}^d\) has an integer Hausdorff and packing dimension if it is strongly regular, i.e., \(\mathcal {H}^t(E)=\mathcal {P}^t(E)\), for \(t\ge 0\). Now, remark that, as a consequence of (1.5), we have if E or F is regular then

$$\begin{aligned} {\mathsf {b}}(E\times F) = {\mathsf {b}}(E)+ {\mathsf {b}}(F). \end{aligned}$$
(1.8)

This is the case, for example, when \(E = {\mathcal {K}}\) the triadic Cantor set. Moreover, if E and F are regular then

$$\begin{aligned} {\mathsf {b}}(E\times F) = {\mathsf {B}}(E\times F). \end{aligned}$$
(1.9)

Thus, \({\mathsf {b}}({\mathcal {K}} \times {\mathcal {K}}) = {\mathsf {B}}({\mathcal {K}} \times {\mathcal {K}}) = 2\log 2/\log 3\). In the following, we give a sufficient condition so that (1.8) and (1.9) hold. Let \(\mu \) be a positive measure on \(\mathbb {R}^d\) and \(\alpha \ge 0\), we define the lower and upper \(\alpha \)-dimensional densities at \(x\in E\) by

$$\begin{aligned} \underline{d}_\mu ^\alpha (x) =\liminf _{r\rightarrow 0}\frac{\mu (E\cap B(x, r))}{(2r)^{\alpha }} \quad \text {and}\quad \overline{d}_\mu ^\alpha (x) =\limsup _{r\rightarrow 0}\frac{\mu (E\cap B(x, r))}{(2r)^{\alpha }} \end{aligned}$$
(1.10)

respectively. We refer to the common value as the \(\alpha \)-dimensional density at x and denote it by \({d}_\mu ^\alpha (x)\). Let \(t,s \ge 0\) and assume that \(0< {\mathsf {H}}^s(E)<+\infty \) and \(0< {\mathsf {H}}^t(F)<+\infty .\) It is well known [8, Theorem 2] that (1.8) holds provided that, for all \(y\in F\), \(d^t_{{\mathsf {H}}^t} (y) >0\). In the next Theorem we will improve this result.

Theorem C

Let \(E, F\subseteq \mathbb {R}^d\), and \(\alpha , \beta >0\).

  1. (1)

    Assume that \(0<{\mathsf {H}} ^{\alpha }(E)< \infty \) and for all \(x\in E\) we have \(\underline{d}^\alpha _{{\mathsf {H}} ^{\alpha }}(x)>0\) then

    $$\begin{aligned} {\mathsf {b}} (E\times F) = {\mathsf {b}} (E)+ {\mathsf {b}} (F), \qquad \quad \forall F\subseteq \mathbb {R}^d. \end{aligned}$$
  2. (2)

    If, in addition, \(0<{\mathsf {H}} ^{\beta }(E)< \infty \) and for all \(y\in F\) we have \(\underline{d}^\beta _{{\mathsf {H}} ^{\beta }}(y)>0\) then

    $$\begin{aligned} {\mathsf {b}} (E\times F) = {\mathsf {b}} (E)+ {\mathsf {b}} (F) = {\mathsf {B}} (E\times F) = {\mathsf {B}} (E)+ {\mathsf {B}} (F). \end{aligned}$$

Remark that, if E and F are regular then \({\mathsf {B}}(E\times F) = {\mathsf {b}}(E)+{\mathsf {B}}(F)\). This is the case, for example, if we take \(E=F\) to be the triadic Cantor set. In the last section of this paper, we will construct a non-standard set F satisfying this equality with E the triadic Cantor set.

2 Construction of Multifractal Measure

Let’s recall that Hausdorff dimension, packing dimension and Hewitt–Stromberg dimension are defined as follows. For \(d\ge 1\), \(\mathbb {R}^d\) is endowed with the Euclidean distance. If \(E\subset \mathbb {R}^d\), \(t\ge 0\) and \(\delta >0\), then we write

$$\begin{aligned} \mathcal {H}_\delta ^t(E)= \inf \left\{ \sum _i \Big (|E_i|\Big )^t \;\; E\subseteq \bigcup _i E_i,\;\; |E_i|< \delta \right\} , \end{aligned}$$

where |A| is the diameter of the set A defined as \(|A| = \sup \big \{ |x - y|, \; x, y \in A \big \}\). We define the t-dimensional Hausdorff measure of E as

$$\begin{aligned} \mathcal {H}^t(E)=\sup _{\delta >0}\mathcal {H}_\delta ^t(E). \end{aligned}$$

The reader can be referred to Rogers’ classical text [30] for a systematic discussion of \(\mathcal {H}^t\).

We also define,

$$\begin{aligned} \overline{\mathcal {P}}_\delta ^t(E)= \sup \left\{ \sum _i \Big (2r_i\Big )^t \right\} , \end{aligned}$$

where the supremum is taken over all disjoint closed balls \(\Big ( B(x_i, r_i) \Big )_i \; \text {such that}\; r_i \le \delta \) and \(x_i \in E\). The t-dimensional packing premeasure, of E is now defined by

$$\begin{aligned} \overline{\mathcal {P}}^t(E)=\sup _{\delta >0}\overline{\mathcal {P}}_\delta ^t(E), \end{aligned}$$

and the packing measure of E by

$$\begin{aligned} {\mathcal {P}}^t(E)=\inf \left\{ \sum _i\overline{\mathcal {P}}^h(E_i)\;\Big |\;\;E\subseteq \bigcup _i E_i\right\} . \end{aligned}$$

While Hausdorff and packing measures are defined using coverings and packings by families of sets with diameters less than a given positive number \(\delta \), the Hewitt–Stromberg measures are defined using covering of balls with the same diameter.

Several authors have investigated the Hewitt–Stromberg measures, that were introduced in [13, Exercise (10.51)], explicitly, for example, in Pesin’s monograph [28, 5.3] or implicitly in Mattila’s text [23]. One can also cite, [9,10,11,12, 19, 27, 36] where these measures have been used on the study of the local properties of fractals and products of fractals. In particular, in [5, pp. 32-36], Edgar provided an outstanding and fine introduction to these measures. One of the purposes of this paper is to define and study a class of natural multifractal generalizations of the Hewitt–Stromberg measures. Although packing and Hausdorff measures are defined using packings and coverings by families of sets with diameters less than a given positive number \(\varepsilon \), say, the Hewitt–Stromberg measures are defined using packings of balls with a fixed diameter \(\varepsilon \). For \(t>0\), the Hewitt–Stromberg pre-measures are defined as follows,

$$\begin{aligned} \overline{\mathsf {H}}^t_0(E)=\liminf _{r\rightarrow 0} \overline{\mathsf {H}}^t_r(E)\quad \text {with }\quad \overline{\mathsf {H}}^t_r(E) = N_r(E) \;(2r)^t \end{aligned}$$

and

$$\begin{aligned} \overline{\mathsf {P}}^t(E)=\limsup _{r\rightarrow 0} \overline{\mathsf {P}}^t_r(E)\quad \text {with }\quad \overline{\mathsf {P}}^t_r(E) =M_r(E) \;(2r)^t, \end{aligned}$$

where the covering number \(N_r(E)\) of E and the packing number \(M_r(E)\) of E are given by

$$\begin{aligned} N_r(E)=\inf \left\{ \sharp \{I\}\;\Big |\; \Big ( B(x_i, r) \Big )_{i\in I} \; \text {is a family of closed balls with}\; x_i \in E \; \text {and}\; E\subseteq \bigcup _i B(x_i, r)\right\} \end{aligned}$$

and

$$\begin{aligned} M_r(E)=\sup \left\{ \sharp \{I\}\; \Big |\; \Big ( B(x_i, r) \Big )_{i\in I} \; \text {is a family of closed balls with}\; x_i \in E \; \text {and}\; d(x_i, x_j)\ge r\; \text {for}\; i\ne j\right\} . \end{aligned}$$

It’s clear that \(\overline{\mathsf {P}}^t\) is increasing and \(\overline{\mathsf {P}}^t(\emptyset ) =0\). However it’s not \(\sigma \)-additive. Therefore, we introduce the upper t-dimensional Hewitt–Stromberg measure, which we denote by \(\mathcal {\mathsf {P}}^t\), defined by

$$\begin{aligned} {\mathsf {P}}^t(E)=\inf \left\{ \sum _i\overline{\mathsf {P}}^t(E_i)\;\Big |\;\;E\subseteq \bigcup _i E_i, \;\; E_i \;\;\text {is bounded}\right\} . \end{aligned}$$

Since \(\overline{\mathsf {H}}^t_0\) is not increasing and not countably subadditive, one needs a standard modification to get an outer measure. Hence, we modify the definition as follows

$$\begin{aligned} \overline{\mathsf {H}}^t(E)=\inf \left\{ \sum _i\overline{\mathsf {H}}^t_0(E_i)\;\Big |\;\;E\subseteq \bigcup _i E_i, \;\; E_i \;\;\text {is bounded}\right\} \end{aligned}$$

and

$$\begin{aligned} {\mathsf {H}}^t(E)=\sup _{F\subseteq E} \overline{\mathsf {H}}^t (F). \end{aligned}$$

We recall some basic inequalities satisfied by the Hewitt–Stromberg, the Hausdorff and the packing measures (see [1, 19, 27])

$$\begin{aligned} {\mathcal {H}}^t(E)\le {\mathsf {H}}^t(E)\le {\mathsf {P}}^t(E)\le {\mathcal {P}}^t(E) \le \overline{\mathcal {P}}^t(E) . \end{aligned}$$
(2.1)

The lower and upper Hewitt–Stromberg dimension \({\mathsf {b}}(E)\) and \({\mathsf {B}}(E)\) are defined by

$$\begin{aligned} {\mathsf {b}}(E)=\inf \Big \{t\ge 0\;\Big |\; {\mathsf {H}}^t(E)=0\Big \}=\sup \Big \{t\ge 0\;\Big |\; {\mathsf {H}}^t(E)=+\infty \Big \} \end{aligned}$$

and

$$\begin{aligned} {\mathsf {B}}(E)=\inf \Big \{t\ge 0\;\big |\; {\mathsf {P}}^t(E)=0\Big \}=\sup \Big \{t\ge 0\;\big |\; {\mathsf {P}}^t(E)=+\infty \Big \}. \end{aligned}$$

As a consequence of (2.1) we get

$$\begin{aligned} \dim _H(E)\le {\mathsf {b}}(E) \le {\mathsf {B}}(E) \le \dim _P(E)\le \overline{\dim }_P(E) \end{aligned}$$

where \(\dim _H\) and \(\dim _P\) denote respectively Hausdorff and packing dimensions and \(\overline{\dim }_P\) is defined as

$$\begin{aligned} \overline{\dim }_P (E)=\inf \Big \{t\ge 0\;\Big |\; \overline{\mathcal {P}}^t(E)=0\Big \}. \end{aligned}$$

Example 1

Consider for \(n\ge 1\), the set \(A_n = \{ 0\}\bigcup \{ 1/k, \;\; k\le n\}\) and

$$\begin{aligned} K= \bigcup _n A_n = \Big \{ 0 \Big \}\; \bigcup \; \Big \{ 1/n, \;\; n\in \mathbb {N}\Big \}. \end{aligned}$$

K is a countable set and then \(\dim _H (K) ={\mathsf {b}}(K) = {\mathsf {B}}(K)=\dim _p(K)=0\). Nevertheless, we have \(\overline{\dim }_p(K) =\frac{1}{2}\) ( Corollary 2.5 in [7] ). Moreover, we can construct ( [1]) a set E such that

$$\begin{aligned} \dim _H(E) = {\mathsf {b}}(E) \ne {\mathsf {B}}(E) =\dim _P(E). \end{aligned}$$

Another two widely used notions of dimensions are those of the lower and upper box dimensions, denoted by \(\underline{\dim }_{B}\) and \(\overline{\dim }_{B}\), respectively, and defined by

$$\begin{aligned} \underline{\dim }_{B}(E)=\liminf _{r\rightarrow 0}\frac{\log N_r(E)}{-\log r}=\liminf _{r\rightarrow 0}\frac{\log M_r(E)}{-\log r} \end{aligned}$$

and

$$\begin{aligned} \overline{\dim }_{B}(E)=\limsup _{r\rightarrow 0}\frac{\log M_r(E)}{-\log r}=\limsup _{r\rightarrow 0}\frac{\log N_r(E)}{-\log r} \end{aligned}$$

for any bounded subset E of \(\mathbb {R}^d\). We refer to the common value as the box dimension and denote it by \(\dim _B\). These dimensions satisfy the following inequalities,

$$\begin{aligned} \dim _H(E)\le \dim _P(E)\le \overline{\dim }_{B}(E) \end{aligned}$$

and

$$\begin{aligned} \dim _H(E)\le \underline{\dim }_{B}(E)\le \overline{\dim }_{B}(E). \end{aligned}$$

In addition we have, \(\underline{\dim }_{B}(E) = \underline{\dim }_{B}(\overline{E})\) and \(\overline{\dim }_{B}(E) = \overline{\dim }_{B}(\overline{E})\). An immediate consequence of this is that if E is a dense subset of an open region of \(\mathbb {R}^d\) then \(\underline{\dim }_{B}(E)=\overline{\dim }_{B}(E)=d\). In particular, countable sets can have nonzero box dimension. Thus, these notions of dimensions are not \(\sigma \)-stable in general, i.e., it is not true in general that \(\dim _B (\bigcup _i E_i) = \sup _i \dim _B (E_i)\). In fact, the lower box dimension is not even finitely stable in general. It is possible, through a slight modification of their definitions, to obtain countably stable notions of dimension called the lower and upper modified box dimensions and defined, for any \(E\subseteq \mathbb {R}^d\), by (see [6, 24])

$$\begin{aligned} \underline{\dim }_{MB}(E)= \inf \left\{ \sup _{i} \underline{\dim }_{B}(E_i)\;\Big |\;E\subseteq \bigcup _i E_i, \;\; E_i \;\;\text {is bounded}\;\; \right\} \end{aligned}$$

and

$$\begin{aligned} \overline{\dim }_{MB}(E)= \inf \left\{ \sup _{i} \overline{\dim }_{B}(E_i) \;\Big |\; E\subseteq \bigcup _i E_i,\;\; E_i \;\;\text {is bounded} \;\; \right\} . \end{aligned}$$

These dimensions satisfy the following inequalities (see for example [1, 6])

$$\begin{aligned} \dim _H(E)\le \underline{\dim }_{MB}(E)\le \overline{\dim }_{MB}(E)\le \dim _P(E), \end{aligned}$$

The reader is referred to [6] for an excellent discussion of lower and upper box dimensions.

The pre-measure \(\overline{{\mathsf {P}}}^{t}\) assigns in the usual way a multifractal dimension to each subset E of \(\mathbb {R}^d\). It is denoted by \({ \Delta }(E)\in [-\infty ,+\infty ]\) such that

$$\begin{aligned} \overline{\mathsf {P}}^{t}(E)=\left\{ \begin{matrix} \infty &{}\text {if}&{} t< { \Delta }(E),\\ \\ 0 &{} \text {if}&{} { \Delta }(E) < t.\end{matrix}\right. \end{aligned}$$

In addition, we have

$$\begin{aligned} {\mathsf {b}}(E) \le {\mathsf {B}}(E) \le { \Delta }(E). \end{aligned}$$

We note also that there exists a unique number \({\Theta }(E)\in [-\infty ,+\infty ]\) such that

$$\begin{aligned} {\overline{{\mathsf {H}}}}^{t}_{0}(E)=\left\{ \begin{matrix} \infty &{}\text {if}&{} t< {\Theta }(E),\\ \\ 0 &{} \text {if}&{} {\Theta }(E) < t.\end{matrix}\right. \end{aligned}$$

Example 2

Consider the set K defined in Example 1. We will prove that \(\Theta ( K) = 1/2\). For \(n\ge 1\) and \(r_k < \frac{1}{n+n^2}\), remark that

$$\begin{aligned} N_{r_k}(A_n) = n+1. \end{aligned}$$

It follows that

$$\begin{aligned} \overline{{\mathsf {H}}}_{r_k}^{1/2}(K)\ge \overline{{\mathsf {H}}}_{r_k}^{1/2}(A_n) = \sqrt{2} \frac{n+1}{\sqrt{n+n^2}}. \end{aligned}$$

Thereby, we have \( \overline{{\mathsf {H}}}_{r_k}^{1/2}(A_n) \ge 1\) which implies that

$$\begin{aligned} \Theta ( K) \ge \Theta (A_n) \ge 1/2. \end{aligned}$$

Let \(\Big (B(x_i , r )\Big )_{ i\in J}\) be a centred covering of K. Then, using Besicovitch’s covering theorem, we can construct \(\xi \) finite sub-families \(\Big (B(x_{1j}, r)\Big )_j\), ...,\(\Big (B(x_{\xi j}, r )\Big )_j\), such that each \( K \subseteq \displaystyle \bigcup _{i=1}^\xi \bigcup _jB(x_{ij}, r)\) and \(\Big (B(x_{ij}, r)\Big )_j\) is a packing of K. Therefore,

$$\begin{aligned} \overline{{\mathsf {H}}}_0^t(K) \le \xi \overline{{\mathsf {P}}}^t(K) \le \xi \overline{\mathcal {P}}^t(K). \end{aligned}$$

Which imply that

$$\begin{aligned} \Theta (K) \le \Delta (E) \le \overline{\dim }_p(K)=\frac{1}{2}. \end{aligned}$$

It is well known [1, Propositon 2] that, for any bounded set \(E \subseteq \mathbb {R}^d\),

$$\begin{aligned} {\Theta }(E) =\underline{\dim }_{B}(E) \qquad \text {and }\qquad {\Delta }(E) = \overline{\dim }_{B}(E). \end{aligned}$$

It is clear that \(\Delta \) and \(\Theta \) are monotonic. In addition, for any sets \(E, F\subseteq \mathbb {R}^d\), we have \(\overline{{\mathsf {P}}}^t(E\cup F)\le \overline{{\mathsf {P}}}^t(E)+\overline{{\mathsf {P}}}^t(F).\) Thus, \(\Delta \) is stable, i.e.,

$$\begin{aligned} \Delta (E\cup F) =\max \big (\Delta (E), \Delta (F)\big ). \end{aligned}$$
(2.2)

Let \(n\in \mathbb {N}\). A closed interval of the kind \([k 2^{-n}, (k+1) 2^{-n}]\), \(k\in \mathbb {Z}\), is called \(2^n\)-mesh in \(\mathbb {R}\). A \(2^n\)-mesh in \(\mathbb {R}^d\) is a closed ball, the projection of which on each axis is a \(2^n\)-mesh. We denote by \(w(2^n, E) \) the number of \(2^n\)-meshes meeting the bounded set E. There are other definitions of \(\Delta \) and \(\Theta \) in \(\mathbb {R}^d\) which are more convenient to use in this paper [6, Section 3.1]

$$\begin{aligned} {\Delta }(E)= & {} \limsup _{n\rightarrow 0} \frac{\log w(2^n, E) }{n\log 2}, \end{aligned}$$
(2.3)
$$\begin{aligned} {\Theta }(E)= & {} \liminf _{n\rightarrow 0} \frac{\log w(2^n, E) }{n\log 2}. \end{aligned}$$
(2.4)

Example 3

Let E be the set in [0, 1] defined by \(E = \displaystyle \bigcap \nolimits _n E_n\), where \(E_n\) is the union of \(2^n\) non overlapping intervals of length \(b_n\), each of them containing two intervals of \(E_{n+1}\). Assume that \(b_0=1\) and \(b_n>2b_{n+1}\). Now, if \(2r\in [b_{n+1}, b_n]\), then \(2^n\le N_{2r}(E) \le 2^{n+1}\). Hence,

$$\begin{aligned} \Theta (E) = \liminf _{n\rightarrow \infty }\frac{n\log 2}{-\log b_n} \quad \text {and}\quad \Delta (E) = \limsup _{n\rightarrow \infty }\frac{n\log 2}{-\log b_n}. \end{aligned}$$

In the special case when E is the triadic Cantor set \((b_n=3^{-n})\) then

$$\begin{aligned} \dim _H (E) ={\mathsf {b}}(E) = {\mathsf {B}}(E)=\dim _p(E)=\Theta (E) = \Delta (E)= \frac{\log 2}{\log 3}. \end{aligned}$$

3 Proofs of Main Results

For \(n\in \mathbb {N}\), we write

$$\begin{aligned} \mathcal {U}_n = \Big \{ \prod _{i=1}^d \Big [\frac{ l_i}{2^n},\frac{ l_i +1}{2^n} \Big [, \;\; l_1,\ldots , l_d\in \mathbb {Z}\Big \} \end{aligned}$$

and

$$\begin{aligned} \mathcal {V}_n = \Big \{ \prod _{i=1}^d \Big [\frac{\frac{1}{2} l_i}{2^n},\frac{\frac{1}{2} l_i +1}{2^n} \Big [, \;\; l_1,\ldots , l_d\in \mathbb {Z}\Big \}. \end{aligned}$$

The family \(\mathcal {U}_n\) is the class of half-open dyadic cubes of order n. We denote \(u_n(x)\), for \(x\in \mathbb {R}^d\), the unique \(u\in \mathcal {U}_n\) that contains x. The family \(\mathcal {V}_n\) is the class of half-open dyadic semi-cubes of order n. We denote \(v_n(x)\), for \(x\in \mathbb {R}^d\), the unique \(v\in \mathcal {V}_n\) that contains x whose complement is at distance \(2^{-n-2}\) from \(u_{n+2}(x)\). Let \(\mathcal {K}=\{(k_1,\ldots , k_d),\; k_i=0, \frac{1}{2}\}\). For each \(\mathbf{k} =(k_1,\ldots , k_d)\in \mathcal {K}\) let

$$\begin{aligned} \mathcal {V}_\mathbf{k , n}=\Big \{ \prod _{i=1}^d \Big [\frac{ k_i+l_i}{2^n},\frac{k_i+ l_i +1}{2^n} \Big [, \;\; l_1,\ldots , l_d\in \mathbb {Z}\Big \}. \end{aligned}$$

Observe that, for \(v\ne v'\in \mathcal {V}_\mathbf{k , n}\), we have \(v\cap v'=\emptyset \). Furthermore \(\Big (\mathcal {V}_\mathbf{k , n}\Big )_\mathbf{k \in \mathcal {K}}\) is a partition of \(\mathcal {V}_n\). Moreover, if \(v, v'\in \mathcal {V}_\mathbf{k} :=\bigcup _{n\ge 0} \mathcal {V}_\mathbf{k , n}\), then either \(v\cap v'=\emptyset \) or one of them is contained in the other. Finally, for \(E\subset \mathbb {R}^d\), write

$$\begin{aligned} \mathcal {V}_n(E) = \{v_n(x),\; x\in E\}\quad \text {and} \quad \mathcal {V}_\mathbf{k , n}(E) = \mathcal {V}_n(E)\cap \mathcal {V}_\mathbf{k , n}. \end{aligned}$$

Now, we will construct new measures, on the Euclidean space \(\mathbb {R}^d\), in a similar manner to the lower and upper Hewitt–Stromberg measures using the class of all half-open dyadic semi-cubes in the definition rather than the class of all closed balls. More precisely, let \(E\subseteq \mathbb {R}^d\) and \(t\ge 0\), let

$$\begin{aligned} \overline{\mathsf {H}}^{*t}(E)=\liminf _{n\rightarrow +\infty } N_n^*(E) \; 2^{-nt}, \quad \text {and}\quad \overline{\mathsf {P}}^{*t}(E)=\limsup _{n\rightarrow +\infty } M_n^*(E) \; 2^{-nt}, \end{aligned}$$

where the numbers \(N_n^*(E)\) and \(M_n^*(E)\) of E are given by

$$\begin{aligned} N_n^{*}(E)=\inf \left\{ \sharp \{I\}\; \Big |\; \Big ( v_i \Big )_{i\in I} \; \; \text {is a family of coverings of}\;\; E\;\; \text {with } \;\; v_i\in \mathcal {V}_n(E) \right\} . \end{aligned}$$

and

$$\begin{aligned} M_n^{*}(E)=\sup \left\{ \sharp \{I\}\; \; v_i \in \mathcal {V}_n(E), \; i\in I, \quad \text {and }\quad \overline{v}_i \cap \overline{v}_j =\emptyset \quad \text {for } i\ne j \right\} . \end{aligned}$$

It’s clear that \(\overline{\mathsf {H}}^{*t}\) and \(\overline{\mathsf {P}}^{*t}\) are increasing with \(\overline{\mathsf {H}}^t(\emptyset ) = \overline{\mathsf {P}}^t(\emptyset ) =0\). However they are not \(\sigma \)-additive. Therefore, we define

$$\begin{aligned} {\mathsf {H}}^{*t}(E)= & {} \inf \left\{ \sum _i\overline{\mathsf {H}}^{*t}(E_i)\;\Big |\;\;E\subseteq \bigcup _i E_i\;\; E_i \;\;\text {is bounded} \right\} .\\ {\mathsf {P}}^{*t}(E)= & {} \inf \left\{ \sum _i\overline{\mathsf {P}}^{*t}(E_i)\;\Big |\;\;E\subseteq \bigcup _i E_i \;\; E_i \;\;\text {is bounded}\right\} . \end{aligned}$$

Lemma 1

For every set \(E\subset \mathbb {R}^d\), for any \(t\ge 0\), there exists a constant \(c>0\) such that

$$\begin{aligned} c^{-1}{{\mathsf {P}} }^t(E) \le {{\mathsf {P}} ^*}^t(E) \le c {\mathsf {P}} ^t(E) \quad \text {and}\quad c^{-1}{{\mathsf {H}} }^t(E) \le {{\mathsf {H}} ^*}^t(E) \le c {\mathsf {H}} ^t(E) \end{aligned}$$
(3.1)

Proof

It follows from the fact that \(B(x, 2^{-n-2}) \subseteq v_n(x) \subseteq B(x, \sqrt{d} 2^{-n})\) \(\square \)

In a similar manner, we can construct two measures \({\mathsf {H}}^{**t}\) and \( {\mathsf {P}}^{**t}\) using the class of all half-open dyadic cubes in the definition rather than the class of all half-open dyadic semi-cubes. Nevertheless, the new pre-measure \( \overline{\mathsf {P}}^{**t}\) is not equivalent to the pre-measure \(\overline{\mathsf {P}}^{t}\), (see [35, Example 3.5]).

Now, we state a useful proposition which will be applied in the proof of our main results.

Proposition 1

Let \(E \subseteq \mathbb {R}^d\). Then,

$$\begin{aligned} {\mathsf {b}} (E) =\sup _{F\subseteq E} \left\{ \inf \left\{ \sup _{i} {\Theta }(F_i) \;\Big |\;\displaystyle F\subseteq \bigcup _i F_i, \;\; F_i \;\;\text {is bounded in} \;\; \mathbb {R}^d \right\} \right\} \end{aligned}$$

and

$$\begin{aligned} {\mathsf {B}} (E) = \inf \left\{ \sup _{i} {\Delta } (E_i) \;\Big |\;\displaystyle E\subseteq \bigcup _i E_i, \;\; E_i \;\;\text {is bounded in} \;\; \mathbb {R}^d\right\} . \end{aligned}$$
(3.2)

Proof

Denote by

$$\begin{aligned} \beta = \inf \left\{ \sup _{i} {\Delta } (E_i) \;\Big |\;\displaystyle E\subseteq \bigcup _i E_i, \;\; E_i \;\;\text {is bounded in} \;\; \mathbb {R}^d\right\} . \end{aligned}$$

Assume that \(\beta < {\mathsf {B}}(E)\) and take \(\alpha \in (\beta , {\mathsf {B}}(E) )\). Then, there exists \(\{E_i\}\) a sequence of subsets of \(\mathbb {R}^d\) such that \(E\subseteq \cup _i E_i\), and \(\sup _i {\Delta }(E_i) < \alpha \). Now observe that \(\overline{{\mathsf {P}}}^{ \alpha }(E_i) = 0\) which implies that \({{\mathsf {P}}}^{ \alpha }(E) = 0\). It is a contradiction. Now suppose that \( {{\mathsf {B}}}(E) < \beta \), then, for \(\alpha \in ( {{\mathsf {B}}}(E), \beta ),\) we have \({{\mathsf {P}}}^{ \alpha }(E) = 0\). Thus, there exists \(\{E_i\}\) a sequence of subsets of \(\mathbb {R}^d\) such that \(E\subseteq \cup _i E_i\), and \(\sup _i \overline{{\mathsf {P}}}^{ \alpha } (E_i) < \infty \). We conclude that, \(\sup _i {\Delta } (E_i) \le \alpha \). It is also a contradiction. The proof of the second statement is identical to the first one and is therefore omitted. \(\square \)

3.1 Proof of Inequality (1.2) and (1.5)

Let \(s,t\ge 0\), \(E\subseteq \mathbb {R}^d \) and \(F\subseteq \mathbb {R}^l\). We will prove that, there exists a constant \(c>0\) such that

$$\begin{aligned} {\mathsf {H}}^{s+t}(E\times F) \le c {\mathsf {H}}^s(E) {\mathsf {P}}^t(F). \end{aligned}$$

Let \(r>0\) and \(H\subseteq E\times F\). Let \(\{B(x_i, r)\}_i\) be a centred r-covering of E. We denote n to be the integer such that \(\sqrt{l} 2^{-n} < r \le \sqrt{l} 2^{-n+1}\). For each i and \(v\in \mathcal {V}_n(F)\) with \((B(x_i,r)\times v) \cap H) \ne \emptyset \) choose a point \(y_{i,v}\in B(x_i, r)\) and a point \(y'_{i,v}\in v\) such that \((y_{i,v}, y'_{i,v}) \in (B(x_i,r)\times v) \cap H)\). Observe that

$$\begin{aligned} H\subseteq & {} \bigcup _i \Big ( \bigcup _{{\mathop {(B(x_i,r)\times v)\cap H\ne \emptyset }\limits ^{v\in \mathcal {V}_n(F)}}} B(x_i, r)\times v \Big ) \subseteq \bigcup _i \Big ( \bigcup _{{\mathop {(B(x_i,r)\times v)\cap H\ne \emptyset }\limits ^{v\in \mathcal {V}_n(F)}}} B(y_{i,v}, 2 r)\times B(y'_{i,v}, 2r) \Big ) \\\subseteq & {} \bigcup _i \Big ( \bigcup _{{\mathop {(B(x_i,r)\times v)\cap H\ne \emptyset }\limits ^{v\in \mathcal {V}_n(F)}}} B((y_{i,v},y'_{i,v}), 2r) \Big ). \end{aligned}$$

Therefore, the family \(\big (B((y_{i,v},y'_{i,v}), 2r) \big )_{i\in \mathbb {N}, v\in \mathcal {V}_n(F), B(x_i,r)\times v)\cap H\ne \emptyset }\) is a centred (2r)-covering of H. Moreover, we have \(B(y'_{i,v}, \eta _r)\subseteq B(y'_{i,v}, 2^{-n-2})\) for \(\eta _r= 2^{-3}\sqrt{l} r \). Hence, for each \(\mathbf{k} \in \mathcal {K}\), the family

$$\begin{aligned} \big (B(y'_{i,v}, \eta _r \big ), \;\; i\in \mathbb {N}, v\in \mathcal {V}_\mathbf{k ,n}(F), B(x_i,r)\times v)\cap H\ne \emptyset \end{aligned}$$

is a centred \(\eta _r\)-packing of F. In follows that

$$\begin{aligned} \overline{{\mathsf {H}}}_{2r}^{s+t}(H)\le & {} \sum _i\Big ( \sum _{{\mathop {(B(x_i,r)\times v)\cap H\ne \emptyset }\limits ^{v\in \mathcal {V}_n(F)}}} (4r)^{s+t}\Big )\\\le & {} 2^{s+t} \left( \frac{8}{\sqrt{l}}\right) ^t \sum _i (2r)^s \Big ( \sum _\mathbf{k \in \mathcal {K}} \sum _{{\mathop {(B(x_i,r)\times v)\cap H\ne \emptyset }\limits ^{v\in \mathcal {V}_\mathbf{k ,n}(F)}}} (2\eta _r)^{t}\Big )\\\le & {} 2^{s+t} \left( \frac{8}{\sqrt{l}}\right) ^t \sum _i (2r)^s \Big ( \sum _\mathbf{k \in \mathcal {K}} \overline{{\mathsf {P}}}_{\eta _r}^{t}(F)\Big )\\\le & {} 2^l 2^{s+t} \overline{{\mathsf {P}}}_{\eta _r}^{t}(F) \left( \frac{8}{\sqrt{l}}\right) ^t \sum _i (2r)^s \end{aligned}$$

Then, by taking the infimum over all centred r-covering of E we obtain

$$\begin{aligned} \overline{{\mathsf {H}}}_{2r}^{s+t}(H)\le 2^l 2^{s+t} \left( \frac{8}{\sqrt{l}}\right) ^t \overline{{\mathsf {H}}}^s_{r}(E) \overline{{\mathsf {P}}}_{\eta _r}^{t}(F) . \end{aligned}$$

Therefore,

$$\begin{aligned} \overline{{\mathsf {H}}}_{0}^{s+t}(H)\le c \liminf _{r\rightarrow 0} \overline{{\mathsf {H}}}^s_{r}(E) \limsup _{r\rightarrow 0} \overline{{\mathsf {P}}}_{\eta _r}^{t}(F)= c \overline{{\mathsf {H}}}^s_{0 }(E) \overline{{\mathsf {P}}}^t (F), \end{aligned}$$

where \(c=2^l 2^{s+t} \Big (\frac{8}{\sqrt{l}}\Big )^t\). Now, assume that \(E\subseteq \bigcup _i E_i\) and \(F\subseteq \bigcup _j F_j\). Then \(H\subseteq E\times F \subseteq \bigcup _{i,j} E_i\times F_j\). Therefore

$$\begin{aligned} \overline{{\mathsf {H}}}^{s+t}(H)\le & {} \sum _{i,j} \overline{{\mathsf {H}}}_0^{s+t} (E_i\times F_j)\le c \sum _{i,j} \overline{{\mathsf {H}}}^s_{0 }(E_i) \overline{{\mathsf {P}}}^t (F_j).\\\le & {} c \Big (\sum _{i} \overline{{\mathsf {H}}}^s_{0 }(E_i)\Big )\Big (\sum _j\overline{{\mathsf {P}}}^t (F_j)\Big ). \end{aligned}$$

Since the cover \((E_i)\) of E and the cover \((F_j)\) of F were arbitrarily chosen, we deduce that

$$\begin{aligned} \overline{{\mathsf {H}}}^{s+t}(H) \le c \overline{{\mathsf {H}}}^s(E) {\mathsf {P}}^t(F) \le c {\mathsf {H}}^s(E) {\mathsf {P}}^t(F). \end{aligned}$$

This is true for all \(H\subseteq E\times F\), hence

$$\begin{aligned} {\mathsf {H}}^{s+t}(E\times F) \le c {\mathsf {H}}^s(E) {\mathsf {P}}^t(F). \end{aligned}$$

As a direct consequence, we obtain the inequality (1.5) which can be obtained directly without the use of the lower and upper Hewitt–Stomberg measures. More precisely, let \(E\subseteq \mathbb {R}^d\) and \( F\subseteq \mathbb {R}^l\) and assume, without loss of generality, that E and F are bounded. Suppose that we have :

$$\begin{aligned} {\mathsf {b}}(E\times F) \le {\mathsf {b}}(E) + \Delta (F). \end{aligned}$$
(3.3)

Therefore, if \(F=\bigcup _n F_n\), we get

$$\begin{aligned} {\mathsf {b}}(E\times F_n) \le {\mathsf {b}}(E) + \Delta (F_n) \quad \text {and then}\quad {\mathsf {b}}(E\times F) \le {\mathsf {b}}(E) + \sup _n\Delta (F_n), \end{aligned}$$

where we have used the \(\sigma \)-stability of \({\mathsf {b}}\), i.e., \(b(\cup F_n) =\sup _n b(F_n)\). Thus, it is sufficient to apply Proposition 1 to get the inequality (1.5). Now, we will prove (3.3). Let \(\alpha > {\mathsf {b}}(E)\) and \(\beta >\Delta (F)\). Let \(\widetilde{E} \subseteq E\) and let \(\widetilde{F} \subseteq F\), using (2.3), we can choose \(N\in \mathbb {N}\) such that,

$$\begin{aligned} \forall \; n\ge N, \qquad w(2^n, \widetilde{F})2^{-n\beta }< 1. \end{aligned}$$
(3.4)

In addition \(\overline{{\mathsf {H}}}^\alpha (\widetilde{E}) = 0\), then we can find a countable sequence \(\{\widetilde{E}_i\}_{i\ge 1}\) such that, for each i, we have \(\overline{{\mathsf {H}}}^\alpha _0(\widetilde{E}_i) <1/2^{i}\). Thus, there exists a decreasing sequence \(\{r_{im}\}_m\) and \(M_i \ge N\) such that

$$\begin{aligned} \forall \, m \ge M_i, \qquad N_{r_{im}}(\widetilde{E}_i) (2 r_{im})^\alpha <1/2^{i}. \end{aligned}$$

For each i and \(m\ge M_i\), we can find a family of closed balls \({\mathcal {B}}_{im}:= \{B(x_j, r_{im})\}_{j}\) with \(x_j\in \widetilde{E}_i\) and \(\widetilde{E}_i \subseteq \bigcup _{j} B(x_j, r_{im})\) such that

$$\begin{aligned} \# \{{\mathcal {B}}_{im}\} (2 r_{im})^\alpha <1/2^{i} \end{aligned}$$
(3.5)

Let \(\widetilde{B} \in {\mathcal {B}}_{im}\) and choose \(n\in \mathbb {N}\) such that \(2^{-n-1}< 2 r_{im}\le 2^{-n}\). Let’s consider the set

$$\begin{aligned} {\mathcal {A}}(\widetilde{B}) = \Big \{ \widetilde{B} \times u,\;\; u \;\;\text {is a}\;2^n\text {-mesh}, u\cap F\ne \emptyset \Big \}. \end{aligned}$$

\({\mathcal {A}}(\widetilde{B})\) is covered by \((2l)^l w(2^n, F)\) balls centred in \(\widetilde{F}\) of diameter \(2^{-n}\). Thus, using (3.4) and (3.5), we have

$$\begin{aligned} N_{2^{-n-1}}({\widetilde{E}}_{i} \times \widetilde{F}) (2^{-n})^{\alpha +\beta }\le & {} \# \{{\mathcal {B}}_{im}\} \; (2l)^l w(2^n, F) 2^{-n(\alpha +\beta )}.\\\le & {} (2l)^l 2^\alpha \; 2^{-i}. \end{aligned}$$

Therefore,

$$\begin{aligned} \overline{{\mathsf {H}}}^{\alpha +\beta }(\widetilde{E}\times \widetilde{F}) \le \sum _i \overline{{\mathsf {H}}}_0^{\alpha +\beta }(\widetilde{E}_i\times \widetilde{F}) \le (2l)^l 2^\alpha . \end{aligned}$$

Since this is true of all \(\widetilde{E}\subseteq E\) and \(\widetilde{F}\subseteq F\), we obtain :

$$\begin{aligned} {{\mathsf {H}}}^{\alpha +\beta }( E\times F) < \infty \end{aligned}$$

and then (3.3) as required.

Corollary 1

Assume that \({\mathsf {b}} (F) ={\mathsf {B}} (F)\) then

$$\begin{aligned} {\mathsf {b}} (E\times F) = {\mathsf {b}} (E)+{\mathsf {b}} (F). \end{aligned}$$

Example 4

In this example, we will consider the Cantor target set H and we will calculate \({\mathsf {b}}(H).\) Let \(t>0\). \(S: \mathbb {R}^d\rightarrow \mathbb {R}^l\) is a Lipschitz function if

$$\begin{aligned} |S(x) - S(y) | \le c |x - y|, \qquad \forall x, y\in \mathbb {R}^d, \end{aligned}$$

for some \(c>0\) and S is a bi-Lipschitz function if

$$\begin{aligned} c_1 |x - y| \le |S(x) - S(y) | \le c_2 |x - y|, \qquad \forall x, y\in \mathbb {R}^d, \end{aligned}$$

for some \(c_1, c_2 >0\). Now, we will prove the following elementary lemma.

Lemma 2

Let \(t>0\) and S be a Lipschitz function, Then, we have

$$\begin{aligned} {\mathsf {H}} ^t(S(A)) \le c^t \, {\mathsf {H}} ^t(A), \quad \forall A\subseteq \mathbb {R}^d, \end{aligned}$$
(3.6)

and then \({\mathsf {b}} (S(A)) \le {\mathsf {b}} (A)\). Moreover, if S is a bi-Lipschitz transformation then \({\mathsf {b}} (S(A))={\mathsf {b}} (A)\).

Proof

Let \(({\mathcal {B}}_i)_{i\in I}\) be a (r/c)-cover of a bounded set \(E\subset \mathbb {R}^d\) by centred closed balls, then there exists \(({\mathcal {B}'}_i)_{i\in I}\) an r-cover of S(E) by centred closed balls. Therefore,

$$\begin{aligned} N_{r}(S(E)) (2r)^t \le c^t \; N_{r/c} (E) (2r/c)^t. \end{aligned}$$

Thus, letting r tend to 0, we get

$$\begin{aligned} {\overline{{\mathsf {H}}}}^t_0(S(E)) \le c^t\; {\overline{{\mathsf {H}}}}^t_0(E). \end{aligned}$$

Therefore, for any set \(F\subseteq A \) such that \(F \subseteq \bigcup _i E_i\) with \(E_i\) are bounded, we get

$$\begin{aligned} \overline{\mathsf {H}}^t(S(F)) \le \sum _i\overline{\mathsf {H}}^t_0 (S(E_i)) \le c^t \sum _i\overline{\mathsf {H}}^t_0 (E_i). \end{aligned}$$

Since \(\bigcup _i E_i\) is an arbitrary cover of F we obtain

$$\begin{aligned} \overline{\mathsf {H}}^t(S(F)) \le c^t\; \overline{\mathsf {H}}^t(F) \le c^t\; {\mathsf {H}}^t(A). \end{aligned}$$
(3.7)

Since this is true for any subset \(F\subseteq A\), we get the desired result. \(\square \)

Let E be the uniform Cantor set. Then, from Example 3, we have \({\mathsf {b}}(E) ={\mathsf {B}}(E) =\frac{\log 2}{\log 3}\). Therefore for any subset F of \(\mathbb {R}\) we have

$$\begin{aligned} {\mathsf {b}}(E\times F) = {\mathsf {b}}(E) + {\mathsf {b}}(F). \end{aligned}$$
(3.8)

Now, we consider the plane set given in polar coordinates by

$$\begin{aligned} H = \Big \{ (r, \theta ), \; r\in E, \; \theta \in [0, 2\pi ]\, \Big \} \end{aligned}$$

We will prove that \({\mathsf {b}}(H) = 1+ \frac{\log 2}{\log 3}\). We consider the function \(S : \mathbb {R}^2\rightarrow \mathbb {R}^2\) defined as \(S((x, y)) = (x\cos y, x\sin y)\). It is clear that S is a Lipschitz mapping and \(H= S(E\times [0, 2\pi ])\). It follows, using Lemma 2 and (3.8), that

$$\begin{aligned} {\mathsf {b}}(H)= & {} {\mathsf {b}}(S(E\times [0, 2\pi ])) \le {\mathsf {b}}(E\times [0, 2\pi ])\\= & {} {\mathsf {b}}(E) + {\mathsf {b}}([0, 2\pi ])= 1+ \frac{\log 2}{\log 3}. \end{aligned}$$

On the other hand, if we restrict S to \([\frac{2}{3}, 1]\times [0 , \pi ]\) then S is bi-Lipschitz function on this domain. Since\(S\big ( (E\cap [\frac{2}{3}, 1]) \times [0, \pi ]\big ) \subset H\), we deduce that

$$\begin{aligned} {\mathsf {b}}(H)\ge & {} {\mathsf {b}}\Big ( S\big ( (E\cap [\frac{2}{3}, 1]) \times [0, \pi ]\big )\Big )={\mathsf {b}}\Big ((E\cap [\frac{2}{3}, 1]) \times [0 , \pi ]\Big )\\= & {} {\mathsf {b}}\big ( E\cap [\frac{2}{3}, 1] \big ) + {\mathsf {b}}\big ( [0 , \pi ]\big ) = 1+ \frac{\log 2}{\log 3}. \end{aligned}$$

Corollary 2

If \(E_1, E_2, \ldots , E_n\) are subsets of \(\mathbb {R}\) then

$$\begin{aligned} {\mathsf {b}}(E_1\times \cdots \times E_n)\le n-1 +\min _{1\le i\le n} {\mathsf {b}}(E_i). \end{aligned}$$

3.2 Proof of Inequalities (1.3) and (1.6)

Let \(s,t\ge 0\), \(E\subseteq \mathbb {R}^d \) and \(F\subseteq \mathbb {R}^l\). We will prove that, there exists a constant \(c>0\), such that

$$\begin{aligned} {\mathsf {P}}^{s+t}(E\times F) \ge c {\mathsf {H}}^s(E) {\mathsf {P}}^t(F). \end{aligned}$$

For simplicity, we restrict the result to subsets of the plane, though the work extends to higher dimensions without difficulty. Let \(\mathcal {Q}\) be any packing of F containing semi-dyadic intervals and \(\mathcal {R}\) be any covering of E containing semi-dyadic intervals. We denote

$$\begin{aligned} \mathcal {R}_1= & {} \Big \{u_i \in \mathcal {R}, u_i \; \text {is dyadic and}\;\; \overline{u}_i \cap \overline{u}_j = \emptyset \;\; \text {for}\;\; i\ne j \;\; \Big \}\\ \mathcal {R}_2= & {} \Big \{u_i \in \mathcal {R}, u_i \; \text {is not dyadic and}\;\; \overline{u}_i \cap \overline{u}_j = \emptyset \;\; \text {for}\;\; i\ne j \;\; \Big \}\\ \mathcal {R}_3= & {} \Big \{u_i \in \mathcal {R}, u_i \; \text {is dyadic} \;\; \Big \}\bigcap \mathcal {R}\backslash \mathcal {R}_1\;\;\text {and}\\&\mathcal {R}_4= \Big \{u_i \in \mathcal {R}, u_i \; \text {is not dyadic} \;\; \Big \}\bigcap \mathcal {R}\backslash \mathcal {R}_2. \end{aligned}$$

Clearly, we have each of \(\mathcal {R}_i\) is a packing of E and \(\mathcal {R}_i\times \mathcal {Q}\) is a packing of \(E\times F\). Therefore,

$$\begin{aligned} 4 M_n^{*} (E\times F) 2^{-n(s+t)}\ge \sum _{\mathcal {Q}} 2^{-nt} \Big ( \sum _{\mathcal {R}_1} 2^{-ns} + \sum _{\mathcal {R}_2} 2^{-ns}+\sum _{\mathcal {R}_3} 2^{-ns}+\sum _{\mathcal {R}_4} 2^{-ns}\Big ). \end{aligned}$$

Since \(\mathcal {Q}\) is an arbitrary packing of F and \(\mathcal {R}=\bigcup _i \mathcal {R}_i\), we obtain

$$\begin{aligned} 4 M_n^{*} (E\times F) 2^{-n(s+t)}\ge M_n^*(F) 2^{-nt} \sum _{\mathcal {R}} 2^{-ns}\ge M_n^*(F) 2^{-nt} \mathbb {N}_n^*(E) 2^{-ns}. \end{aligned}$$

Thus,

$$\begin{aligned} \overline{{\mathsf {P}}}^{*t+s}(E\times F) \ge \frac{1}{4} \overline{{\mathsf {P}}}^{*t}(F) \overline{{\mathsf {H}}}^{*t}(E) \ge \frac{1}{4} {{\mathsf {P}}}^{*t}(F) {{\mathsf {H}}}^{*t}(E). \end{aligned}$$

Finally, using (3.1), we get the desired result. As a direct consequence, we obtain the inequality (1.6) which can be obtained directly using the definitions. More precisely, remark that for each \(n\in \mathbb {N}\), we have

$$\begin{aligned} w(2^n, E\times F) = w(2^n, E) \; w(2^n, F). \end{aligned}$$
(3.9)

Therefore, using Proposition 1, we get

$$\begin{aligned} \Delta (E\times F) \ge \Theta (E) +\Delta (F) \ge {\mathsf {b}}(E) +\Delta (F). \end{aligned}$$
(3.10)

In addition, by Proposition 1 and (2.2), we may assume, without loss of generality, that for \(\epsilon >0\), there exists \((E_n) \nearrow E\) and \((F_n)\nearrow F\) such that

$$\begin{aligned} \sup \Delta (E_n\times F_n) \le {\mathsf {B}}(E\times F) + \epsilon . \end{aligned}$$
(3.11)

Thus, by (3.10), we have \(\Delta (E_n\times F_n) \ge {\mathsf {b}}(E_n) +\Delta (F_n)\) and using (3.11), we get

$$\begin{aligned} {\mathsf {B}}(E\times F) \ge \sup _n\Big ( {\mathsf {b}}(E_n) +\Delta (F_n)\Big )-\epsilon \ge {\mathsf {b}}(E) +{\mathsf {B}}(F) -\epsilon \end{aligned}$$

as required.

3.3 Proof of Inequalities (1.4) and (1.7)

Let \(s,t\ge 0\), \(E\subseteq \mathbb {R}^d \) and \(F\subseteq \mathbb {R}^l\). We will prove that, there exists a constant \(c>0\), such that

$$\begin{aligned} {\mathsf {P}}^{s+t}(E\times F) \le c {\mathsf {P}}^s(E) {\mathsf {P}}^t(F). \end{aligned}$$

For simplicity, we restrict the result to subsets of the plane, though the work extends to higher dimensions without difficulty. Let \(\mathcal {B}\) be any packing of \(E\times F\) containing semi-dyadic squares, each of which is the product of two semi-dyadic intervals. We denote

$$\begin{aligned} \mathcal {R}=\Big \{u_n(x) : \exists v_n(y) \;\; \text {such that }\;\; w_n(x,y)=u_n(x)\times v_n(y) \in {\mathcal {B}}, \;\; x\in E, y\in F\Big \} \end{aligned}$$

and

$$\begin{aligned} \mathcal {Q}=\Big \{v_n(x) : \exists u_n(y) \;\; \text {such that }\;\; w_n(x,y)=u_n(x)\times v_n(y) \in {\mathcal {B}}, \;\; x\in E, y\in F\Big \}. \end{aligned}$$

Now, we consider the subclasses

$$\begin{aligned} \mathcal {R}_1= & {} \Big \{u_n(x) \in \mathcal {R},\;\; u_n(x) \;\; \text {is dyadic} \Big \} \quad \mathcal {Q}_1 =\Big \{v_n(x) \in \mathcal {Q},\;\; v_n(x) \;\; \text {is dyadic} \Big \}\\ \mathcal {R}_2= & {} \Big \{u_n(x) \in \mathcal {R},\;\; u_n(x) \;\; \text {is not dyadic} \Big \} \quad \mathcal {Q}_2 =\Big \{v_n(x) \in \mathcal {Q},\;\; v_n(x) \;\; \text {is not dyadic} \Big \}. \end{aligned}$$

Clearly, we have each of \(\mathcal {R}_1, \mathcal {R}_2\) is a packing of E and each of \(\mathcal {Q}_1\), \(\mathcal {Q}_2\) is a packing of F. Moreover, each square of the packing \({\mathcal {B}}\) is in the collection \(\mathcal {R}_i\times \mathcal {Q}_j\), \(i, j\in \{1, 2\}\). Therefore,

$$\begin{aligned} \sum _{\mathcal {B}} 2^{-n(t+s)}\le & {} \Big [ \sum _{\mathcal {R}_1} 2^{-nt} + \sum _{\mathcal {R}_2} 2^{-nt}\Big ]\Big [ \sum _{\mathcal {Q}_1} 2^{-ns} + \sum _{\mathcal {Q}_2} 2^{-ns}\Big ]\\\le & {} 4 M_n^*(E) 2^{-nt} M_n^*(F) 2^{-ns}. \end{aligned}$$

Since, this is true for any packing of \(E\times F\), we obtain \(M_n^*(E\times F) 2^{-n(t+s)} \le 4 M_n^*(E) 2^{-nt} M_n^*(F) 2^{-ns}\) and then

$$\begin{aligned} \overline{{\mathsf {P}}}^{*t+s} (E\times F) \le 4 \overline{{\mathsf {P}}}^{*t}(E)\; \overline{{\mathsf {P}}}^{*s}(F). \end{aligned}$$

Now, for \(E\subseteq \bigcup _i E_i\), and \( F\subseteq \bigcup _j F_j\), we have :

$$\begin{aligned} {\mathsf {P}}^{*t+s} (E\times F)\le & {} \sum _{i,j} \overline{{\mathsf {P}}}^{*t+s}(E_i\times F_j)\le 4 \sum _{i,j} \overline{{\mathsf {P}}}^{*t}(E_i)\; \overline{{\mathsf {P}}}^{*s}(F_j).\\\le & {} 4 \Big (\sum _i \overline{{\mathsf {P}}}^{*t}(E_i)\Big ) \; \Big ( \sum _j \overline{{\mathsf {P}}}^{*s}(F_j)\Big ). \end{aligned}$$

Since \((E_i)\) is an arbitrary covering of E and \((F_j)\) is an arbirary covering of F, we get \({\mathsf {P}}^{*t+s} (E\times F) \le 4 {\mathsf {P}}^{*t} (E){\mathsf {P}}\) \(^{*s} ( F)\). Finally, using (3.1), we obtain the desired result. As a direct consequence, we obtain the inequality (1.7) which can be obtained directly using the definitions. Indeed, using again (3.9), we obtain

$$\begin{aligned} \Delta (E\times F) \le \Delta (E) +\Delta (F). \end{aligned}$$

Hence, by Proposition 1, we get the inequality (1.7).

Corollary 3

If \(E_1, E_2, \ldots , E_n\) are subsets of \(\mathbb {R}\) then

$$\begin{aligned} {\mathsf {B}}(E_1\times \cdots \times E_n)\le n-1 +\min _{1\le i\le n} {\mathsf {B}}(E_i). \end{aligned}$$

3.4 Proof of Theorem C

To prove the first assertion, we only have to prove that E is regular. Assume, without loss of generality, that the set E is bounded and consider, for \(k, n\ge 1\) the set

$$\begin{aligned} E(n, k) = \Big \{ x\in E; \; r< \frac{1}{k} \implies \frac{(2r)^\alpha }{n} < {\mathsf {H}}^\alpha (E\cap B(x,r)\Big \}. \end{aligned}$$

Therefore, under our assumption, we have \(E= \bigcup _{n, k} E(n,k)\). Let \(n, k\ge 1\) such that \(E(n,k)\ne \emptyset \). Then, for each ball B(xr) such that \(x\in E\) and \(r< \frac{1}{k}\) we have \( (2r)^\alpha < n \; {\mathsf {H}}^\alpha (E\cap B(x,r)) \) and then

$$\begin{aligned} M_r(E(n,k))(2r)^\alpha\le & {} \sum _B (2r)^\alpha \le n \sum _B \; {\mathsf {H}}^\alpha (E\cap B(x,r)) \\\le & {} n \; {\mathsf {H}}^\alpha (E). \end{aligned}$$

It follows that \(M_r(E(n,k)) \le n (2r)^{-\alpha } {\mathsf {H}}^\alpha (E)\) and then \(\Delta (E(n,k)) \le \alpha \). Now, remark that our assumptions imply that \({\mathsf {b}}(E) =\alpha \), on the other hand, we have

$$\begin{aligned} {\mathsf {B}}(E) = \sup _{n, k} {\mathsf {B}}(E(n,k)) \le \sup _{n, k} \Delta (E(n,k))\le \alpha , \end{aligned}$$

as required. Now, similarly, we have F is regular and the second assertion follows.

4 Example

Let E be the triadic Cantor set and then \( {\mathsf {b}}(E) =\Theta (E) =\frac{\log 2}{\log 3}<1.\) In this example, we will construct, for any \(\gamma \) such that \( \Theta (E)<\gamma <1\), a set \(F\subset \mathbb {R}\) such that

$$\begin{aligned} {\mathsf {B}}(F)\ge 1-\gamma \end{aligned}$$
(4.1)

and

$$\begin{aligned} {\mathsf {B}}(E\times F)\ge 1 \end{aligned}$$
(4.2)

Thus, since for any \(\epsilon >0\), there exists \(\gamma \in ({\mathsf {b}}(E), {\mathsf {b}}(E)+\epsilon )\) and

$$\begin{aligned} \gamma -\epsilon + 1-\gamma \le {\mathsf {b}}(E) + {\mathsf {B}}(F) \le {\mathsf {B}}(E\times F)\le 1 \end{aligned}$$

and then by arbitrariness of \(\epsilon \) we get \({\mathsf {b}}(E) + {\mathsf {B}}(F) = {\mathsf {B}}(E\times F).\)

First, we will give a sufficient condition to get \(\Delta ={\mathsf {B}}\) on a compact set. Let E be a compact set on \(\mathbb {R}^d\) and let

$$\begin{aligned} \Delta (E, x) =\lim _{r\rightarrow 0} \Delta (E\cap B_x( r)), \end{aligned}$$

where \(B_x( r)\) is the open ball with center x and radius r. \(\Delta \) is said to be uniform on E if \(\Delta (E, x)\) is constant on E. This constant will be denoted by \(\Delta ^*(E).\)

Theorem 1

Let E be a compact set on \(\mathbb {R}^d\). If \(\Delta \) is uniform on E then

$$\begin{aligned} \Delta (E) = {\mathsf {B}}(E). \end{aligned}$$

Proof

We will prove that \(\Delta (E)= \Delta ^*(E). \) Let \(x\in E\) and \(\epsilon >0\). Since \(\Delta \) is uniform on E, there exists \(r_x>0\) such that \(\Delta (E\cap B_x( r_x)) \le \Delta ^*(E)+\epsilon \). Now E is compact then we can extract from \(\{B_x(r_x)\}_{x\in E}\) a finite subfamily \(\{B_i\}_{i=1}^n\) that cover E. Therefore

$$\begin{aligned} \Delta (E) \le \Delta \Big (\bigcup _{i=1}^nB_i\Big )=\sup _{i} \Delta (B_i)\le \Delta ^*(E) +\epsilon . \end{aligned}$$

Since \(\epsilon \) is arbitrarily chosen, we get \(\Delta (E)\le \Delta ^*(E)\) and hence

$$\begin{aligned} \Delta (E)= \Delta ^*(E) \end{aligned}$$

since the converse inequality is trivial. This equality implies, for each \(x\in E\) and \(r>0\), that

$$\begin{aligned} \Delta (E\cap B_x(r))=\Delta (E). \end{aligned}$$

Recall (3.2) and assume that \({\mathsf {B}}(E)< \Delta (E)\), then there exists a sequence \((E_n)\) such that \(E = \bigcup E_n\) and \(\Delta (E_n)< \Delta (E)\) for all \(n\ge 1\). Let \(x\in E_n\) and \(r>0\), then

$$\begin{aligned} \Delta (E_n)=\Delta (E_n \cap B_x(r)) < \Delta (E)= \Delta (E\cap B_x(r)). \end{aligned}$$

Using the fact that \(\Delta (F) =\Delta (\overline{F})\) for every set \(F\subseteq \mathbb {R}^d\), we conclude that \(E_n\) is nowhere dense (that is contained in a closed set with empty interior) in E and then E is meager (or of first category). Therefore E is not a closed set by Baire’s theorem. That is contradiction, then \({\mathsf {B}}(E) =\Delta (E)\). \(\square \)

4.1 Construction of the Set F

Since \( \Theta (E)<\gamma ,\) then for n big enough and for any decreasing sequence \((\delta _n)_{n\ge _0}\) which converges to 0 we have

$$\begin{aligned} N_{\delta _n}(E) < \delta _n^{-\gamma }. \end{aligned}$$
(4.3)

In addition, we may choose \((\delta _n)\) such that

$$\begin{aligned} \delta _0=1 \quad \text {and} \quad \delta _n = a_n^{1/(\gamma -1)} \delta _{n-1} \end{aligned}$$

with \((a_n)\) is a positive sequence of integers such that \(a_n\rightarrow \infty \). Hence, for n big enough, we have

$$\begin{aligned} \Big (\frac{ \delta _{n}}{\delta _{n-1}}\Big )^\gamma < \frac{1}{2}. \end{aligned}$$
(4.4)

Now, we will construct by inductively a sequence of sets \((F_n)_{n\ge 1}\). For \(n = 1\), let

$$\begin{aligned} b_1= \Big ( \frac{\delta _{0}}{\delta _1}\Big )^{1-\gamma } \end{aligned}$$

and let

$$\begin{aligned} F_1= \bigcup _{j=1}^{b_1} J_{i_1}, \end{aligned}$$

where \(J_{i_1}\), \(i_1=1, 2,\ldots , b_1\), are closed subintervals of [0, 1] of length \(\delta _1\). We arrange those intervals such that they are equally spaced with gaps \(\delta _1\) and they are contained in an interval of length at most \(2 b_1 \delta _1 = 2 \delta _1^\gamma <1\) by (4.4).

Let \(n\ge 2 \) and assume that \(F_{n-1}\) has been constructed as the union of \( \frac{1}{\delta _{n-1}^{1-\gamma }} \) disjoint closed intervals \(J_{i_1,\ldots , i_{n-1}}\) each of them of length \(\delta _{n-1}\). Let

$$\begin{aligned} b_n= \Big (\frac{\delta _{n-1}}{\delta _n}\Big )^{1-\gamma } , \end{aligned}$$

in each interval of \(F_{n-1}\) we construct \(b_n\) closed intervals \(J_{i_1,\ldots , i_{n}}\) of length \(\delta _n\) in such a way that these intervals are equall with gaps \(\delta _n\) and they are contained in an interval of length at most \(2 b_n \delta _n = 2 \delta _{n-1}^{1-\gamma } \delta _n^\gamma < \delta _{n-1}\) by (4.4).

Therefore, \(F_n \) is the union of \(b_1\times \cdots \times b_n \) intervals \(J_{i_1,\ldots , i_{n}}\) and then

$$\begin{aligned} F_n= \bigcup _{i_1=1}^{b_1} \cdots \bigcup _{i_n=1}^{b_n} J_{i_1,\ldots , i_{n}}, \end{aligned}$$

where \( J_{i_1,\ldots , i_{n}}\), are closed subintervals of [0, 1] of length \(\delta _n\). Finally let

$$\begin{aligned}F= \bigcap _{n} F_n.\end{aligned}$$

4.2 Proofs of (4.1) and (4.2)

It is clear that, for \(r = \delta _n/2\), we have

$$\begin{aligned} N_r(F) = \frac{1}{\delta _{n}^{1-\gamma }} \end{aligned}$$

and then, by definition of \(\Delta \) and Theorem 1, we obtain

$$\begin{aligned} {\mathsf {B}}(F) =\Delta (F) \ge 1-\gamma . \end{aligned}$$

Now, we will prove (4.2). Let \(0<r<2^{-1/\gamma }\) and choose n such that \(\delta _n \le r< \delta _{n-1}\). If   \(2 \delta _{n-1}^{1-\gamma } \delta _n^{\gamma }\le r < \delta _{n-1}\), then, using (4.3), there exists a positive constant C such that

$$\begin{aligned} N_r(E\times F ) \le N_{\delta _{n-1}}(E) C \frac{\delta _{n-1}}{r} \frac{1}{\delta _{n}^{1-\gamma }}\le \frac{C}{r}. \end{aligned}$$

Now, we consider the case where \(\delta _n \le r< 2 \delta _{n-1}^{1-\gamma } \delta _n^{\gamma },\) then, using again (4.3) we obtain

$$\begin{aligned} N_r(E\times F)\le N_{\delta _n}(E) \frac{1}{\delta _{n}^{1-\gamma }} C \frac{2 \delta _{n-1}^{1-\gamma } \delta _n^{\gamma }}{r}\le \frac{C}{r}. \end{aligned}$$

Therefore, we have

$$\begin{aligned} \Delta (E\times F) = {\mathsf {B}}(E\times F) \le 1. \end{aligned}$$