1 Preliminaries

Assume that \(\Sigma \) is a \(\sigma \)-algebra of subsets of a set \(X\ne \emptyset \) and let \(\mathcal J\subsetneq \Sigma \) be a \(\sigma \)-ideal. We denote \(A\sim B\) when the symmetric difference \(A\bigtriangleup B\) belongs to \(\mathcal J\). Note that \(\sim \) is an equivalence relation on \(\Sigma \). We say that a mapping \(\Phi :\Sigma \rightarrow \Sigma \) is a lower density operator with respect to \(\mathcal J\) if the following conditions are satisfied:

  1. (I)

    \(\Phi (X)=X\) and \(\Phi (\emptyset )=\emptyset \),

  2. (II)

    \(A\sim B\implies \Phi (A)=\Phi (B)\) for every \(A,B\in \Sigma \),

  3. (III)

    \(A\sim \Phi (A)\) for every \(A\in \Sigma \),

  4. (IV)

    \(\Phi (A\cap B)=\Phi (A)\cap \Phi (B)\) for every \(A,B\in \Sigma \).

A standard example of a lower density operator deals with the \(\sigma \)-algebra \({{\mathcal {L}}}\) of Lebesgue measurable subsets of \({\mathbb {R}}\) and the \(\sigma \)-ideal \({{\mathcal {N}}}\) of null sets; see [8]. The respective operator \(\Phi :{\mathcal L}\rightarrow {{\mathcal {L}}}\) assigns to \(A\in {\mathcal {L}}\) its set of density points. It is known (cf. [13]) that the value \(\Phi (A)\) is always a Borel set of type \(F_{\sigma \delta }\). The exact version of this example can also be considered for the Cantor space \(\{0,1\}^{\mathbb {N}}\) with the respective product measure (see [1] and [6, Exercise 17.9]).

New important examples of lower density operators have been obtained thanks to a reformulation of the classical notion of a Lebesgue density point proposed in [12] and [9]. Before we describe this idea, we recall the notion of convergence with respect to a \(\sigma \)-ideal, introduced in [11]. Given a pair \(\Sigma , \mathcal J\) as above, we say that a sequence \((f_n)\) of \(\Sigma \)-measurable functions \(f_n:X\rightarrow {\mathbb {R}}\), \(n\in {\mathbb {N}}\), converges to a \(\Sigma \)-measurable function \(f:X\rightarrow {\mathbb {R}}\) with respect to \(\mathcal J\) whenever every subsequence \((f_{n_k})\) of \((f_n)\) contains a subsequence \((f_{n_{k_p}})\) such that

$$\begin{aligned} \{x:f_{n_{k_p}}(x)\nrightarrow f(x)\}\in \mathcal J, \end{aligned}$$

that is \(f_{n_{k_p}}(x)\rightarrow f(x)\) holds \(\mathcal J\)-almost everywhere.

Note that, if \(X:=(-1,1)\), \(\Sigma :={\mathcal L}\upharpoonright X\) and \({\mathcal J}:={\mathcal N}\upharpoonright X\), this convergence is equivalent to convergence in measure, by the Riesz theorem. We use the fact that Lebesgue measure is finite on \((-1,1)\). Here we put \({\mathcal L}\upharpoonright X:=\{ A\in {{\mathcal {L}}}:A\subseteq X\}\) and \({{\mathcal {N}}}\upharpoonright X:=\{ A\in {{\mathcal {N}}}:A\subseteq X\}\).

For \(A\subseteq {\mathbb {R}}\) and \(\alpha , x\in {\mathbb {R}}\) we denote \(\alpha A:=\{\alpha t:t\in A\}\) and \(A\pm x:=\{t\pm x:t\in A\}\). The characteristic function of A will be written as \(\chi _A\). By \(\lambda \) we denote Lebesgue measure on \({\mathbb {R}}\). Let \(x\in {\mathbb {R}}\) and \(A\in \mathcal L\). From the original definition stating that x is a density point of A one can easily deduce the following equivalent version:

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{\lambda \left( A\cap \left( x-\frac{1}{n},x +\frac{1}{n}\right) \right) }{\frac{2}{n}}=1 \end{aligned}$$

which in turn is equivalent to each of the next two conditions:

$$\begin{aligned} \lim _{n\rightarrow \infty }\lambda (n(A-x)\cap (-1,1))=2; \end{aligned}$$
$$\begin{aligned} \text{ the } \text{ sequence } \left( \chi _{n(A-x)\cap (-1,1)}\right) _{n\in {\mathbb {N}}} \text{ converges } \text{ in } \text{ measure } \text{ to } \chi _{(-1,1)}. \end{aligned}$$
(1)

This last condition can be applied to the case of the Baire category where A is replaced by a set with the Baire property and convergence in measure is replaced by convergence with respect to the \(\sigma \)-ideal of meager sets in \({\mathbb {R}}\). This idea turned our fruitful since it initiated broad investigations of density points and density like topologies in the Baire category sense (see [3, 9, 13]).

Let us describe some datails now. By \(\mathcal B\) we denote the \(\sigma \)-algebra of subsets of \({\mathbb {R}}\) with the Baire property, and by \(\mathcal M\) – the \(\sigma \)-ideal of meager subsets of \({\mathbb {R}}\). A set \(A\in \mathcal B\) can be written as the symmetric difference \(U\bigtriangleup E\) where U is open and \(E\in \mathcal M\). Moreover, U can be chosen regular open and then the above expression is unique; see [8]. The unique regular open set U in the expression \(A=U\bigtriangleup E\) will be denoted by \(\widetilde{A}\) and called the regular open part of A. If \(\Sigma :=\mathcal B\) and \(\mathcal J:=\mathcal M\), then convergence with respect to the \(\sigma \)-ideal \(\mathcal M\) will be called convergence in (the Baire) category.

We say that 0 is an \(\mathcal M\)-density point of a set \(A\in \mathcal B\) whenever the sequence of characteristic functions \((\chi _{nA\cap (-1,1)})\) converges in category to the constant function equal to 1 on \((-1,1)\). If, in this statement, the constant function equal to 1 is replaced by the constant function equal to 0, we say that number 0 is an \(\mathcal M\)-dispersion point of A. In one-sided versions of these definitions one should replace \((-1,1)\) by \((-1,0)\) and (0, 1). Given \(x\in {\mathbb {R}}\), we say that x is an \(\mathcal M\)-density (respectively, \(\mathcal M\)-dispersion) point of \(A\in \mathcal B\) whenever 0 is an \(\mathcal M\)-density (respectively, \(\mathcal M\)-dispersion) point of \(A-z\). Note that if, in the definition of \(\mathcal M\)-density point, we replace the pair \(\mathcal B,\mathcal M\) by \(\mathcal L,\mathcal N\), we obtain the usual notion of density point. Observe that x is an \(\mathcal M\)-density point of A if and only if x is an \(\mathcal M\)-dispersion point of the complementary set \(A^c\), and in the last case, \(A^c\) can be replaced by its regular open part \(\widetilde{A^c}\).

Operator \(\Phi _{\mathcal M}\) assigning to each \(A\in \mathcal B\) its set of \(\mathcal M\)-density points is a lower density operator [9]. A nontrivial question whether its values are Borel sets was settled positively in [2] where a characterization by Łazarow [7] served as a basic tool. Another result of [2] states that the operator \(\Phi _{\mathcal M}\) can be considered in the Cantor space \(\{0,1\}^{\mathbb {N}}\) and it has analogous properties.

Note that, analogously as in the measure case, the family \(\{A\in \mathcal B:A\subseteq \Phi _{\mathcal M}(A)\}\) forms a topology, called the \(\mathcal I\)-density topology (with \(\mathcal I=\mathcal M\)) which has been widely investigated by several authors (see [3, 9]).

The aim of our paper is to show some new instances of lower density or lower density like operators having Borel values. This makes a continuation of studies in [2]. In Sect. 2 we consider another Baire category notion of density point proposed in [14] and called an intensity point. It is essentially different from the notion of \(\mathcal M\)-density point [14, Thm 9]. It generates in a standard way a topology that is not homeomorphic to \(\mathcal I\)-density topology; see [14, Thm 29]. However, if one reformulates the definition of an intensity density point for the pair \(\mathcal L,\mathcal N\), this would give the usual notion of a Lebesgue density point [14, Thm 3]. Our main result of Sect. 2 states that that the intensity operator has Borel values of type \(F_{\sigma \delta \sigma \delta }\) which solves the first part of Problem 3 in [2].

In Sect. 3 we consider almost lower density operators \(\Phi _s, \Phi _c:\mathcal L\rightarrow \mathcal L\) generated by simple and complete density points introduced in [15] and [16], respectively. We show that \(\Phi _s\) and \(\Phi _c\) have Borel values of type \(G_{\delta \sigma \delta }\) and \(G_{\delta \sigma }\), respectively.

2 The Intensity Operator has Borel Values

Recall the notion of an intensity point [14]. Suppose that \(A\in \mathcal B\). For \(n\in {\mathbb {N}}\) let

$$\begin{aligned} A_n:=\left( n(n+1)\left( A-\frac{1}{n+1}\right) \right) \cap (0,1). \end{aligned}$$

Define \(f_n:(0,1)\rightarrow {\mathbb {R}}\) by \(f_n:=\frac{1}{n}\sum _{i=1}^n\chi _{A_i}\) for \(n\in {\mathbb {N}}\). We say that 0 is a right-hand intensity point of A if the sequence \((f_n)\) converges in category to the constant function equal to 1 on \((-1,1)\). We say that 0 is a rarefaction point of A if the sequence \((f_n)\) converges to 0 in category. Clearly, 0 is a right-hand intensity point of A if and only if it is a rarefaction point of the complement \(A^c\) of A. The definition of a left-hand intensity and rarefaction points is self explaining (then we consider \((-1,0)\) instead of (0, 1)). We say that 0 is an intensity (rarefaction) point of A if it is both right-hand and left-hand intensity (rarefaction) point of A. At last we say that \(x\in {\mathbb {R}}\) is an intensity (rarefaction) point of A whenever 0 is an intensity (rarefaction) point of \(A-x\).

Let \(\Phi _i\) assign to every set \(A\in \mathcal B\) its set of intensity points. It is proved in [14, Thm 6] that \(\Phi _i\) is a lower density operator from \({\mathcal {B}}\) into itself. We will call it briefly the intensity operator. Observe that x is an intensity point of A if and only if it is a rarefaction point of \(\widetilde{A^c}\) (the regular open part of \(A^c\)).

We shall prove that the intensity operator is Borel-valued. We need a characterization of an intensity point which would play a similar role to that played by the result of E. Łazarow in the case of an \(\mathcal M\)-density point [7]. We shall deal with right-hand intensity and rarefaction points since the left-hand case is analogous.

Theorem 1

Let \((f_n)_{n\in {\mathbb {N}}}\) be a sequence of non-negative real-valued functions defined on (0, 1), having the Baire property. For \(h>0\) and \(n\in {\mathbb {N}}\), we put

$$\begin{aligned} f_n^h(x):=\left\{ \begin{array}{ll} h &{} \hbox { if}\,\,\, f_n(x)>h\\ 0 &{} \hbox { if}\,\,\, 0\le f_n(x)\le h. \end{array}\right. \end{aligned}$$

Denote \(f_n^h\) by \(f_n^{(k)}\) if \(h=1/k\), \(k\in {\mathbb {N}}\). Then the following conditions are equivalent:

(i):

\((f_n)_{n\in {\mathbb {N}}}\) converges to 0 in category;

(ii):

\((f_n^h)_{n\in {\mathbb {N}}}\) converges to 0 in category for each \(h>0\);

(iii):

\((f_n^{(k)})_{n\in {\mathbb {N}}}\) converges to 0 in category for each \(k\in {\mathbb {N}}\).

Proof

(i) \(\Rightarrow \) (ii) If \((f_n)_{n\in {\mathbb {N}}}\) coverges to 0 in category, then \((f_n^h)_{n\in {\mathbb {N}}}\) also converges to 0 in category because \(0\le f_n^h\le f_n\) for all \(n\in {\mathbb {N}}\) and \(h>0\).

Implication (ii) \(\Rightarrow \) (iii) is obvious.

(iii) \(\Rightarrow \) (i) From the assumption it follows that for each \(k\in {\mathbb {N}}\) and for each increasing sequence \((n_m)_{m\in {\mathbb {N}}}\) of positive integers there exists a subsequence \((n_{m_p})_{p\in {\mathbb {N}}}\) for which

$$\begin{aligned} \left\{ x\in (0,1):\limsup _{p\rightarrow \infty }f^{(k)}_{n_{m_p}}(x)\ge h\right\} \in \mathcal M. \end{aligned}$$

We will use the denotation \((n_m^{(1)}), (n_m^{(2)}),\dots \) for successive subsequences in the inductive construction described below.

For \(k=1\) pick a subsequence \((n_m^{(1)})\) such that

$$\begin{aligned} \left\{ x\in (0,1):\limsup _{m\rightarrow \infty }f^{(1)}_{n^{(1)}_m}(x)\ge 1\right\} \in \mathcal M. \end{aligned}$$

If we have constructed subsequences \((n_m^{(1)}), (n_m^{(2)}),\dots ,(n_m^{(k)})\), we pick a subsequence \((n_m^{(k+1)})\) of \((n_m^{(k)})\), for which

$$\begin{aligned} \left\{ x\in (0,1):\limsup _{m\rightarrow \infty }f^{(k+1)}_{n^{(k+1)}_m}(x)\ge \frac{1}{k+1}\right\} \in \mathcal M. \end{aligned}$$

Now, consider the diagonal sequence \((n_{m_p})_{p\in {\mathbb {N}}}\) for \((n_m^{(1)}),(n_m^{(2)}),\dots \). Then

$$\begin{aligned} \left\{ x\in (0,1):\limsup _{p\rightarrow \infty }f_{n_{m_p}}(x)>0\right\} \subseteq \bigcup _{k\in {\mathbb {N}}} \left\{ x\in (0,1):\limsup _{m\rightarrow \infty } f^{(k)}_{n^{(k)}_m}(x)\ge \frac{1}{k}\right\} . \end{aligned}$$
(2)

Indeed, let \(x\in (0,1)\) and \(\limsup _{p}f_{n_{m_p}}(x)>0\). Then there exists \(h>0\) such that \(f_{n_{m_p}}(x)>h\) for infinitely many p’s. If \(k\in {\mathbb {N}}\) is such that \(1/k<h\), then \(f^{(k)}_{n^{(k)}_m}(x)\ge \frac{1}{k}\). Finally from (2) it follows that

$$\begin{aligned} \left\{ x\in (0,1):\limsup _{p\rightarrow \infty }f_{n_{m_p}}(x)>0\right\} \in \mathcal M. \end{aligned}$$

\(\square \)

We enumerate a fixed basis of the natural topology in (0, 1) as \(((a_k, b_k))_{k\in {\mathbb {N}}}\). Then we enumerate a fixed basis of the natural topology in \((a_k,b_k)\) as \(((c_{k,n}, d_{k,n}))_{n\in {\mathbb {N}}}\).

Theorem 2

Suppose that \((A_n)_{n\in {\mathbb {N}}}\) is a sequence of open subsets of (0, 1). Then the following conditions are equivalent:

(i):

the sequence \((\chi _{A_n})\) converges to 0 in category;

(ii):

for each \(k\in {\mathbb {N}}\) there exist \(\varepsilon >0\) and \(n_0\in {\mathbb {N}}\) such that for each \(n>n_0\) there exists \(m\in {\mathbb {N}}\) with \(d_{k,m}-c_{k,m}>\varepsilon \) and \(A_n\cap (c_{k,m}-d_{k,m})=\emptyset \).

Proof

(i)\(\Rightarrow \)(ii) Suppose that (ii) does not hold. Then there exists \(k_0\in {\mathbb {N}}\) such that for all \(\varepsilon >0\) and \(j\in {\mathbb {N}}\) there exists \(n=n(\varepsilon ,j)>j\) such that for each \(m\in {\mathbb {N}}\) if \(d_{k_0,m}-c_{k_0,m}>\varepsilon \), then \(A_{n(\varepsilon ,j)}\cap (c_{k_0,m}, d_{k_0,m})\ne \emptyset \). Hence for \(\varepsilon =1\) and \(j=1\) there exists \(n_1=n(1,1)>1\), for \(\varepsilon =1/2\) and \(j=n_1\) there exists \(n_2=n(1/2,n_1)>n_1\) and so on, which means that for each \(p\in {\mathbb {N}}\), if \(\varepsilon =1/p\) and \(j=n_{p-1}\), then there exists \(n_p=n(1/p,n_{p-1})\) such that, for each \(m\in {\mathbb {N}}\), \(d_{k_0,m}-c_{k_0,m}>1/p\) implies that \(A_{n_p}\cap \left( c_{k_0,m}, d_{k_0,m}\right) \ne \emptyset \). Obviously, the sequence \((n_p)\) is increasing. Observe that, if \(N_0\subseteq {\mathbb {N}}\) is an infinite set, then \(\bigcup _{p\in N_0}A_{n_p}\) is open and dense in \((a_{k_0}, b_{k_0})\). Hence the sequence \((\chi _{A_n})\) does not converge to 0 in category.

(ii)\(\Rightarrow \)(i) To simplify notation we shall choose a subsequence, convergent \(\mathcal M\)-almost everywhere to 0, from the whole sequence \((\chi _{A_n})\). First, take \(k=1\) and \((a_1,b_1)\). There exists \(\varepsilon (1)>0\) and \(n_0(1)\) such that for each \(n>n_0(1)\) there exists \(m_n\in {\mathbb {N}}\) with \((c_{1,m_n}, d_{1, m_n})\cap A_n=\emptyset \) and \(d_{1,m_n}-c_{1,m_n}>\varepsilon (1)\). It is possible to choose an increasing sequence \(S^{(1)}\) of positive integers for which there exists an interval \((u_1,v_1)\subseteq (a_1,b_1)\) such that \((u_1,v_1)\cap A_n=\emptyset \) for each n in \(S^{(1)}\). Repeating this procedure for \(k=2\), we find an interval \((u_2,v_2)\subseteq (a_2,b_2)\) and a subsequence \(S^{(2)}\) of \(S^{(1)}\) such that \((u_2,v_2)\cap A_n=\emptyset \) for each n in \(S^{(2)}\). We construct inductively a sequence \((S^{(m)})_{m\in {\mathbb {N}}}\) of sequences where \(S^{(m+1)}\) is a subsequence of \(S^{(m)}\) for every m. Consider the diagonal sequence \((n_m)\) of \((S^{(m)})\). Then \((\chi _{A_{n_m}})\) converges to 0 on the dense open set \(\bigcup _{k\in {\mathbb {N}}}(u_k,v_k)\), that is, it converges to 0, \(\mathcal M\)-almost everywhere. \(\square \)

Let us preserve our previous notation. Then we obtain the following corollary.

Corollary 3

Suppose that \((f_n)\) is a sequence of non-negative lower semicontinuous functions defined on (0, 1). Then for each \(h>0\) the following conditions are equivalent:

(i):

the sequence \((f^h_n)_{n\in {\mathbb {N}}}\) converges to 0 in category;

(ii):

for each \(k\in {\mathbb {N}}\) there are \(p\in {\mathbb {N}}\) and \(n_0\in {\mathbb {N}}\) such that for each \(n>n_0\) there exists \(m\in {\mathbb {N}}\) with \(d_{k,m}-c_{k,m}>1/p\) and \(\{x\in (0,1):f_n^h(x)>0\}\cap \left( c_{k,m},d_{k,m}\right) =\emptyset \).

Proof

Fix \(h>0\). Since for \(n\in {\mathbb {N}}\) the function \(f_n\) is lower semicontinuous, the set

$$\begin{aligned} A_n:=\{x\in (0,1):f_n^h(x)>0\}=\left\{ x\in (0,1):\frac{1}{h}f_n^h(x)>0\right\} \end{aligned}$$

is open and \(\frac{1}{h}f_n^h(x)=\chi _{A_n}\). Hence it is enough to apply Theorem 2. Additionally, we observe that a number \(\varepsilon >0\) in Theorem 2(ii) can be chosen of the form 1/p for some \(p\in {\mathbb {N}}\). \(\square \)

Theorem 4

For each set \(A\in \mathcal B\), the set \(\Phi _i(A)\) of all intensity points of A is a Borel set of type \(F_{\sigma \delta \sigma \delta }\).

Proof

Fix \(A\in \mathcal B\). Then \(\Phi _i(A)= E^+\cap E^-\) where \(E^+, E^-\) denote (respectively) the sets of right-hand and left-hand intensity points of A. We will show that \(E^+\) is of type \(F_{\sigma \delta \sigma \delta }\). The proof for \(E^-\) is analogous, so it will be omitted.

For \(n\in {\mathbb {N}}\) and \(x\in (0,1)\) let \(f_{n,x}(t):=\frac{1}{n}\sum _{i=1}^n\chi _{A_i(x)}(t)\) if \(t\in (0,1)\) where

$$\begin{aligned} A_i(x):=\left( i(i+1)\left( \widetilde{A^c}-x-\frac{1}{i+1}\right) \right) \cap (0,1). \end{aligned}$$

Since sets \(A_i(x)\) for \(i\in {\mathbb {N}}\), are open, the function \(f_{n,x}\) is lower semicontinuous.

From Theorem 1 it follows that \(E^+\) consists of \(x\in (0,1)\) such that \((f^{(r)}_{n,x})_{n\in {\mathbb {N}}}\) converges to 0 in category for every \(r\in {\mathbb {N}}\) (one should replace \(f^{(k)}_{n}\) by \(f^{(k)}_{n,x}\) in condition (ii) of Theorem 1). Then observe that the statement “\((f^{(r)}_{n,x})_{n\in {\mathbb {N}}}\) converges to 0 in category” can be formulated in an equivalent form by the use of Corollary 3(ii). Consider the sequence \(((c_{k,m}, d_{k,m}))_{m\in {\mathbb {N}}}\) introduced prior to Theorem 2. Given \(p\in {\mathbb {N}}\), we enumerate as \(((c^{p}_{k,m}, d^{p}_{k,m}))_{m\in {\mathbb {N}}}\) all intervals in \(((c_{k,m}, d_{k,m}))_{m\in {\mathbb {N}}}\) for which \(d_{k,m}-c_{k,m}>1/p\). Now, by the previous observation and Corollary 3(ii), we infer that x belongs to \(E^+\) if and only if,

for all \(r, k\in {\mathbb {N}}\) there exist \(p, n_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \forall n>n_0\;\exists m\in {\mathbb {N}}\;\;\{t\in (0,1):f^{(r)}_{n,x}(t)>0\}\cap (c^p_{k,m},d^p_{k,m})=\emptyset . \end{aligned}$$
(3)

Our purpose is to show that the set

$$\begin{aligned} \{x\in (0,1):\{t\in (0,1):f^{(r)}_{n,x}(t)>0\}\cap (c^p_{k,m},d^p_{k,m})=\emptyset \} \end{aligned}$$
(4)

is closed since, by (3), this implies that \(E^+\) is of type \(F_{\sigma \delta \sigma \delta }\).

Claim 1. For a fixed \(t\in (0,1)\), condition \(f^{(r)}_{n,x}(t)>0\) is equivalent to \(f^{(r)}_{n,x}(t)>1/r\). This in turn is equivalent to the following condition:

t

here exist an integer \(j(r)\le n\), with \(j(r)>\lfloor \frac{n}{r}\rfloor +1\), and a sequence \(i_1<i_2<\dots <i_{j(r)}\) of integers in \(\{1,\dots , n\}\) such that \(t\in \bigcap _{j=1}^{j(r)}A_{i_j}(x)\).

Let us prove this last equivalence. Implication “\(\Rightarrow \)” follows from the fact that \(f^{(r)}_{n,x}(t)\) is the arithmetic mean with the components equal to zeros or ones, so the number j(r) of ones should be such that \(\frac{j(r)}{n}>\frac{1}{r}\), hence \(j(r)>\lfloor \frac{n}{r}\rfloor +1\). The proof of the reverse implication is based on a similar reasoning.

From Claim 1 it follows that \(\{t\in (0,1):f^{(r)}_{n,x}(t)>0\}\) is a finite union of sets of the form \(\bigcap _{j=1}^{j(r)}A_{i_j}(x)\). The next claim is a simple observation.

Claim 2. The set \(\{t\in (0,1):f^{(r)}_{n,x}(t)>0\}\) is disjoint from \((c^p_{k,m},d^p_{k,m})\) if and only if all components of the above-mentioned finite union are disjoint from \((c^p_{k,m},d^p_{k,m})\).

Now, observe that, to show the closedness of the set in (4), it suffices to prove that every set of the form \(\{x\in (0,1):\bigcap _{j=1}^{j(r)}A_{i_j}(x)\cap (c^p_{k,m},d^p_{k,m})=\emptyset \}\) is closed. This is true since the complementary set

$$\begin{aligned} \left\{ x\in (0,1):\bigcap _{j=1}^{j(r)}\left( i_j(i_j+1)\left( \widetilde{A^c} -\frac{1}{i_j+1}\right) -i_j(i_j+1)x\right) \cap (c^p_{k,m},d^p_{k,m})\ne \emptyset \right\} \end{aligned}$$

is open. Indeed, the above formula says that the intersection of two open sets is nonempty. So, if \(x_0\) belongs to the above set, then every x sufficiently close to \(x_0\) is also in this set. \(\square \)

It would be interesting to establish whether the class \(F_{\sigma \delta \sigma \delta }\) in Theorem 4 can be replaced by a lower class in the Borel hierarchy.

The definition of an intensity (rarefaction) point of a set \(A\subseteq {\mathbb {R}}\) with the Baire property, introduced in [14], was motivated by the fact that, if we use a similar idea in the measure case, we obtain the classical notion of a density (dispersion) point. In fact in the measure case, the analogue for 0 to be a right-hand rarefaction point of a measurable set \(A\subseteq {\mathbb {R}}\) means that

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{1}{n}\sum _{k=1}^n b_k=0 \end{aligned}$$
(5)

where \(b_k:=k(k+1)\lambda \left( A\cap \left[ \frac{1}{k+1},\frac{1}{k}\right] \right) \) is the average density of A on \(\left[ \frac{1}{k+1},\frac{1}{k}\right] \). By [14, Thm 3] we know that condition (5) is equivalent to the fact that 0 is the right-hand dispersion point of A.

Unfortunately, for the Cantor space \(\{0,1\}^{\mathbb {N}}\) and the respective product probability measure, we are not able to find a reasonable analogue of condition (5). Therefore, we cannot find a counterpart of the notion of a rarefaction point of a set \(A\subseteq \{0,1\}^{\mathbb {N}}\) with the Baire property. Also, the characterization of intensity points contained in the proof of Theorem 4 seems useless for this purpose. Hence the second part of Problem 3 in [2] remains open and we do not see a reasonable solution.

3 The Simple Density Operator and the Complete Density Operator have Borel Values

In [15] and [16], two versions of modified Lebesgue density points were introduced. A main idea was to replace convergence in measure stated in condition (1) by stronger kinds of convergence.

If we consider convergence almost everywhere in (1), we obtain the notion of a simple density point from [15]. So, a number \(x\in {\mathbb {R}}\) is called a simple density point of a set \(A\in \mathcal L\) if the sequence \(\left( \chi _{n(A-x)\cap (-1,1)}\right) _{n\in {\mathbb {N}}}\) converges to \(\chi _{(-1,1)}\) almost everywhere. Denote by \(\Phi _s(A)\) the set of all simple density points of A. If \(\Phi (A)\) is the set of usual (Lebesgue) density points of A, we have \(\Phi _s(A)\subseteq \Phi (A)\) which is an easy consequence of the characterization (1) of \(x\in \Phi (A)\).

From [15] we know that \(\Phi _s:\mathcal L\rightarrow \mathcal L\) and \(\Phi _s\) satisfies all axioms (I)–(IV) of a lower density operator except for (III) which is only partially fulfilled: we have \(\Phi _s(A)\setminus A\in \mathcal N\) (which follows from \(\Phi _s(A)\subseteq \Phi (A)\) and \(\Phi (A){\setminus } A\in \mathcal N\)) but \(A\setminus \Phi _s(A)\) need not hold (see [15, Thm 1]). However, \(\Phi _s\) is useful enough since the family \(\tau _s:=\{A\in \mathcal L:A\subseteq \Phi _s(A)\}\) is a topology stronger than the natural topology and weaker than the usual density topology \(\tau _d\) [15, Thm 2].

In general, a mapping \(\Psi :\Sigma \rightarrow \Sigma \) fulfilling conditions (I), (II), (IV) and the condition \(\Psi (A){\setminus } A\in \mathcal J\) (weaker than (III)) is called an almost lower density operator. See, for instance, [4, 10].

For a set \(C\subseteq {\mathbb {R}}^2\) and \(x\in {\mathbb {R}}\) we define the x-section of C by \(C(x):=\{y:(x,y)\in C\}\).

Theorem 5

For each set \(A\in \mathcal L\), the set \(\Phi _s(A)\) of all simple density points of A is a Borel set of type \(G_{\delta \sigma \delta }\).

Proof

We modify the proof of [15, Proposition 2]. Let \(A\in \mathcal L\). We will check the Borelness of \(\Phi _s(A)\). We may replace A by a \(G_\delta \) set \(A_0\) such that \(A\sim A_0\) since then \(\Phi _s(A)=\Phi _s(A_0)\).

Define \(F:{\mathbb {R}}^2\rightarrow {\mathbb {R}}^2\) by \(F(x,y):=(x,y-x)\) and let \(B:=F({\mathbb {R}}\times A_0)\). Since F is a homeomorphism, B is of type \(G_\delta \). Note that \(B=\{(x,y):y\in A_0-x\}\). For \(n\in {\mathbb {N}}\), let

$$\begin{aligned} B_n=\{(x,ny):(x,y)\in B\}=\{(x,y):x\in {\mathbb {R}},\; y\in n(A_0-x)\}. \end{aligned}$$

Then \(B_n\) is of type \(G_\delta \). For \(x\in {\mathbb {R}}\), condition \(x\in \Phi _s(A_0)\) means that \(\left( \chi _{n(A_0-x)\cap (-1,1)}\right) _{n\in {\mathbb {N}}}\) converges to \(\chi _{(-1,1)}\) almost everywhere which is equivalent to \(\lambda (D(x))=2\) where

$$\begin{aligned} D:=(\liminf _{n\in {\mathbb {N}}} B_n)\cap ({\mathbb {R}}\times (-1,1))=\left( \bigcup _{n\in {\mathbb {N}}}\bigcap _{k>n}B_k\right) \cap ({\mathbb {R}}\times (-1,1)). \end{aligned}$$

Then D is of type \(G_{\delta \sigma }\). Observe that \(\lambda (D(x))=2\) is equivalent to \(\lambda (D(x))>2-\frac{1}{n}\) for every n. Finally, recall that, since D is of type \(G_{\delta \sigma } =\Sigma ^0_3\) (in modern notation), the set

$$\begin{aligned} D_n:=\left\{ x:\lambda (D(x))>2-\frac{1}{n}\right\} \end{aligned}$$

is of the same Borel class (see [6, 17.25, 22.25]) for each \(n\in {\mathbb {N}}\). Hence \(\Phi _s(A_0) =\bigcap _{n\in {\mathbb {N}}} D_n\) is of type \(G_{\delta \sigma \delta }\) as desired. \(\square \)

In [16], another almost lower density operator \(\Phi _c:\mathcal L\rightarrow \mathcal L\) was introduced. This time, in condition (1), complete convergence was considered instead of convergence in measure. We say (see [5]) that a sequence \((f_n)\) of real-valued measurable functions converges completely to a measurable function f on \((-1,1)\) whenever, for every \(\varepsilon >0\),

$$\begin{aligned} \lim _{n\rightarrow \infty }\sum _{j=n}^\infty \lambda (\{y\in (-1,1):\vert f_j (y)-f(y)\vert >\varepsilon \})=0. \end{aligned}$$

Note that this condition implies that \((f_n)\) converges to f almost everywhere. Also, it can be written in an equivalent form as follows

$$\begin{aligned} \sum _{j=1}^\infty \lambda (\{y\in (-1,1):\vert f_j (y)-f(y)\vert >\varepsilon \})<\infty . \end{aligned}$$

Now, a number \(x\in {\mathbb {R}}\) is called a complete density point of a set \(A\in \mathcal L\) if the sequence \(\left( \chi _{n(A-x)\cap (-1,1)}\right) _{n\in {\mathbb {N}}}\) converges completely to \(\chi _{(-1,1)}\). Observe that this is equivalent to the following condition

$$\begin{aligned} \sum _{j=1}^\infty \lambda (j(A^c-x)\cap (-1,1))<\infty . \end{aligned}$$
(6)

Denote by \(\Phi _c(A)\) the set of all complete density points of A. From [16] we know that \(\Phi _c:\mathcal L\rightarrow \mathcal L\) and \(\Phi _c\) is an almost lower density operator such that \(\Phi _c(A)\subseteq \Phi _s(A)\) for all \(A\in \mathcal L\) (moreover, this inclusion can be proper). Also, the family \(\tau _c:=\{A\in \mathcal L:A\subseteq \Phi _c(A)\}\) is a topology stronger than the natural topology and weaker than the topology \(\tau _s\) described above.

The following lemma seems to be known.

Lemma 6

For a measurable set \(B\subset {\mathbb {R}}\), define \(f_B(x):=\lambda ((B+x)\cap (-1,1))\) where \(x\in {\mathbb {R}}\). Then the function \(f_B\) is continuous.

Proof

First we show that \(f_B\) is continuous at 0. Let \(\vert x\vert <1\) and assume, for instance, that \(x>0\). Clearly,

$$\begin{aligned} \lambda ((B+x)\cap (-1,1))= \lambda ((B\cap (-1-x,1-x)) \end{aligned}$$

and this value can be expressed as the sum

$$\begin{aligned} \lambda (B\cap (-1,1))-\lambda ((B\cap [1-x,1))+\lambda ((B\cap (-1-x,-1]). \end{aligned}$$

Hence \(\vert f_B(x)-f_B(0)\vert =\vert -\lambda ((B\cap [1-x,1))+\lambda ((B\cap (-1-x,-1])\vert \le x\). The case \(x<0\) is similar. Thus \(\vert f_B(x)-f_B(0)\vert \le \vert x\vert \) which yields the desired continuity. Now, let \(x_0\ne 0\). Observe that \(f_B(x_0)=f_{B+x_0}(0)\), so the continuity of \(f_B\) at \(x_0\) follows from the previous part of our proof. \(\square \)

Theorem 7

For each set \(A\in \mathcal L\), the set \(\Phi _c(A)\) of all complete density points of A is a Borel set of type \(F_{\sigma }\).

Proof

Let \(A\in \mathcal L\). Then, by (6), a point \(x\in {\mathbb {R}}\) is in \(\Phi _c(A)\) if and only if

$$\begin{aligned} \exists k\in {\mathbb {N}}\;\; \forall n\in {\mathbb {N}}\;\;\,\sum _{j=1}^n \lambda (j(A^c-x)\cap (-1,1)) \le k. \end{aligned}$$

Hence

$$\begin{aligned} \Phi _c(A)=\bigcup _{k\in {\mathbb {N}}}\bigcap _{n\in {\mathbb {N}}}\left\{ x\in {\mathbb {R}}:\sum _{j=1}^n \lambda (j(A^c-x)\cap (-1,1)) \le k\right\} . \end{aligned}$$
(7)

It follows from Lemma 6 that every function \(x\mapsto \lambda (j(A^c-x)\cap (-1,1))\) (for \(j\in \{1,\dots , n\}\)) is continuous, hence so is the function

$$\begin{aligned} x\mapsto \sum _{j=1}^n \lambda (j(A^c-x)\cap (-1,1)). \end{aligned}$$

Consequently, by (7), the set \(\Phi _c(A)\) is of type \(F_{\sigma }\). \(\square \)

We can ask whether evaluation of a Borel class in Theorems 5 and 7 can be improved. If we expect the negative answer for the case of \(\Phi _c\), we should find a set \(A\in \mathcal L\) such that \(\Phi _c(A)\) is of class \(F_{\sigma }{\setminus } G_{\delta }\).