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

Boolean polynomial expression of the discrete average on the space of rhythms

Fumio HAZAMA
Tokyo Denki University
Hatoyama, Hiki-Gun, Saitama JAPAN
e-mail address:[email protected]
Phone number: (81)49-296-5267
Abstract

An infinite family of Boolean polynomials which correspond to the discrete average maps, defined in [2], is constructed. Each member BavN0Bav_{N}^{0} of the family is found to be balanced. A recursive formula, which computes BavN+10Bav_{N+1}^{0} from BavN0Bav_{N}^{0}, is provided.

𝐊𝐞𝐲𝐰𝐨𝐫𝐝𝐬\mathbf{Keywords}: Boolean polynomial, space of rhythms, discrete average, Boolean average, balanced polynomial
MSC 2010: 05E18, 06E30, 37P25.

0 Introduction

The main purpose of this paper is to construct a Boolean counterpart of the discrete average map RavRav, investigated in [2], and to study its algebraic and combinatorial properties. The map RavRav is defined on a space 𝐑Nn\mathbf{R}_{N}^{n} of rhythms with NN beats and with nn onsets, and it is shown that an arbitrary rhythm in 𝐑Nn\mathbf{R}_{N}^{n} can be made smooth by a finite number of iterations of RavRav. An observation which motivates our study in the present paper is that a certain quotient space 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} of 𝐑N=n=0N𝐑Nn\mathbf{R}_{N}=\cup_{n=0}^{N}\mathbf{R}_{N}^{n} is found to be naturally isomorphic to the Boolean space 𝐁N=𝐅2N\mathbf{B}_{N}=\mathbf{F}_{2}^{N}, and that the discrete average descends to a self map on 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n}. It follows that there should be a Boolean counterpart Bav:𝐁N𝐁NBav:\mathbf{B}_{N}\rightarrow\mathbf{B}_{N}, which might shed light on the true nature of RavRav from the Boolean viewpoint. We do not assume the reader’s acquaintance with any theory of music. We will encounter some unexpected algebraic phenomena which might be entitled to be investigated independently of its rhythm-theoretical origin.
As a self map on 𝐁N\mathbf{B}_{N}, the Boolean average BavBav consists of NN Boolean functions. We show, however, that only one of them, called BavN0Bav_{N}^{0}, determines the other N1N-1 functions completely. Furthermore we derive a formula which expresses BavN0Bav_{N}^{0} as a sum of a simple-looking products (Theorem 5.2), and obtain a recurrence formula which computes BavN0Bav_{N}^{0} from BavN10Bav_{N-1}^{0} (Theorem 5.3). As a consequence, we see that the Boolean polynomial BavN0Bav_{N}^{0} is balanced. According to [4, Chapter 3.3], “Boolean functions in cryptographic applications almost always need to be balanced, … .” Our polynomials BavN0Bav_{N}^{0} for N3N\geq 3 provide us unexpectedly with an infinite family of balanced ones.
The plan of this paper is as follows. In Section one, we recall the definition of the space 𝐑Nn\mathbf{R}_{N}^{n} of rhythms and of the discrete average RavRav on 𝐑Nn\mathbf{R}_{N}^{n}. The space 𝐑Nn\mathbf{R}_{N}^{n} is defined to be a subset of the self-product 𝐙Nn\mathbf{Z}_{N}^{n} of 𝐙N={0,1,,N1}\mathbf{Z}_{N}=\{0,1,\cdots,N-1\} equipped with the addition and subtraction operations inherited from those on the abelian group 𝐙/N𝐙\mathbf{Z}/N\mathbf{Z}. Thus rhythms are originally inhabitants of the “modulo-NN-world.” Since our investigation in this article is based on the fact that RavRav preserves the space 𝐑Nn\mathbf{R}_{N}^{n}, we give a simplified proof of this fact here. In Section two, we construct a bridge from this modulo-NN-world to the Boolean world by introducing a certain equivalence relation on 𝐑N\mathbf{R}_{N}, which is compatible with the map RavRav. The quotient space 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} by this relation is found to be naturally isomorphic to the Boolean space 𝐁N\mathbf{B}_{N}, and hence we obtain Boolean counterpart BavNBav_{N}, called Boolean average, on 𝐁N\mathbf{B}_{N}. Section three is devoted to deduce several functorial properties of the Boolean average, which enable us to show that only one of the coordinate BavN0Bav_{N}^{0} of the self map BavNBav_{N} is enough to determine the other coordinates. In Section four we recall a standard way to translate a given propositional function to a Boolean polynomial. By this method one can express the NN coordinates of BavBav as Boolean polynomials. However several examples of the polynomial expressions of BavN0Bav_{N}^{0} for small NN are listed in Table 4.3, they do not give any clue for our attempt to express the polynomials in a unified way. This obstacle leads us to change the basic set of indices 𝐙N\mathbf{Z}_{N} of Boolean vectors to the set 𝐙N,±\mathbf{Z}_{N,\pm} of the least absolute remainders modulo NN. Thanks to the transition to 𝐙N,±\mathbf{Z}_{N,\pm} we obtain in Section five a simple formula which expresses the polynomial BavN0Bav_{N}^{0}. In order to formulate the result we introduce the notion of a ”parental pair of zero”, and that of an ”ancestor of zero” for which BavN0Bav_{N}^{0} takes value 1. By determining the structure of the set of ancestors of zero, we find an explicit form of BavN0Bav_{N}^{0} as well as some recurrence formulas. Furthermore we derive from the structure of the set of the ancestors of zero that BavN0Bav_{N}^{0} is always balanced for any N3N\geq 3.

1 The discrete average on the space of rhythms 𝐑Nn\mathbf{R}_{N}^{n}

In this section we recall some definitions in our previous paper [2], and reformulate them in more geometric language. Throughout the present article we fix an integer NN which is assumed to be greater than or equal to 3.

1.1 𝐙N\mathbf{Z}_{N}-interval and S1S^{1}-interval

Let 𝐙N={0,1,,N1}\mathbf{Z}_{N}=\{0,1,\cdots,N-1\}. We denote by nN\langle n\rangle_{N} the least nonnegative remainder of an integer nn divided by NN. We define the addition ”+N+_{N}” and the difference ”N-_{N}” on 𝐙N\mathbf{Z}_{N} by the rules

a+Nb\displaystyle a+_{N}b =\displaystyle= a+bN,\displaystyle\langle a+b\rangle_{N},
aNb\displaystyle a-_{N}b =\displaystyle= abN,\displaystyle\langle a-b\rangle_{N},

for any a,b𝐙Na,b\in\mathbf{Z}_{N}. For any pair a,b𝐙Na,b\in\mathbf{Z}_{N}, we define the 𝐙N\mathbf{Z}_{N}-interval [a,b]𝐙N[a,b]_{\mathbf{Z}_{N}} by

[a,b]𝐙N={{a,a+1,,b},if ab,{a,a+1,,N1,0,1,,b},if a>b,\displaystyle[a,b]_{\mathbf{Z}_{N}}=\left\{\begin{array}[]{ll}\{a,a+1,\cdots,b\},&\mbox{if $a\leq b$},\\ \{a,a+1,\cdots,N-1,0,1,\cdots,b\},&\mbox{if $a>b$},\end{array}\right.

and define half-closed intervals by

[a,b)𝐙N\displaystyle[a,b)_{\mathbf{Z}_{N}} =\displaystyle= [a,b]𝐙N{b},\displaystyle[a,b]_{\mathbf{Z}_{N}}\setminus\{b\},
(a,b]𝐙N\displaystyle(a,b]_{\mathbf{Z}_{N}} =\displaystyle= [a,b]𝐙N{a}.\displaystyle[a,b]_{\mathbf{Z}_{N}}\setminus\{a\}.

Therefore the number |[a,b)𝐙N||[a,b)_{\mathbf{Z}_{N}}| of the elements in [a,b)𝐙N[a,b)_{\mathbf{Z}_{N}} is given by

|[a,b)𝐙N|=bNa.\displaystyle|[a,b)_{\mathbf{Z}_{N}}|=b-_{N}a. (1.2)

We call the number bNab-_{N}a the 𝐙N\mathbf{Z}_{N}-distance from aa to bb, and denote it by d𝐙N(a,b)d_{\mathbf{Z}_{N}}(a,b). Note that, when aba\neq b, the distance satisfies the equality

d𝐙N(a,b)=Nd𝐙N(b,a),\displaystyle d_{\mathbf{Z}_{N}}(a,b)=N-d_{\mathbf{Z}_{N}}(b,a),

hence it is not necessarily symmetric. These two notions, 𝐙N\mathbf{Z}_{N}-interval and 𝐙N\mathbf{Z}_{N} distance, are translated into geometric ones through a character on the additive group 𝐑\mathbf{R} as follows. Let S1={z𝐂:|z|=1}𝐂S^{1}=\{z\in\mathbf{C}:|z|=1\}\subset\mathbf{C} be the unit circle in the complex plane, and let χN:𝐑𝐂=𝐂{0}\chi_{N}:\mathbf{R}\rightarrow\mathbf{C}^{*}=\mathbf{C}\setminus\{0\} denote the additive character defined by χN(t)=exp(2πit/n)\chi_{N}(t)=\exp(2\pi it/n) for any t𝐑t\in\mathbf{R}. Note that χN(𝐑)=S1\chi_{N}(\mathbf{R})=S^{1} and that χN(𝐙N)=μN\chi_{N}(\mathbf{Z}_{N})=\mu_{N}, where μN\mu_{N} denotes the group of the NN-th roots of unity. For any pair (z,w)S1×S1(z,w)\in S^{1}\times S^{1}, we define the S1S^{1}-interval [z,w]S1S1[z,w]_{S^{1}}\subset S^{1} to be the counterclockwise closed circular arc from zz to ww, and put

[z,w)S1\displaystyle[z,w)_{S^{1}} =\displaystyle= [z,w]S1{w},\displaystyle[z,w]_{S^{1}}\setminus\{w\},
(z,w]S1\displaystyle(z,w]_{S^{1}} =\displaystyle= [z,w]S1{z}.\displaystyle[z,w]_{S^{1}}\setminus\{z\}.

Let dS1(z,w)d_{S^{1}}(z,w) denote the length of the arc [z,w]S1[z,w]_{S^{1}}. We call dS1(z,w)d_{S^{1}}(z,w) the S1S^{1}-distance from zz to ww. Note that we have

dS1(z,w)=2πdS1(w,z),\displaystyle d_{S^{1}}(z,w)=2\pi-d_{S^{1}}(w,z),

for any distinct pair of elements z,wS1z,w\in S^{1}. The 𝐙N\mathbf{Z}_{N}-distance and the S1S^{1}-distance are related by the equation

dS1(χN(a),χN(b))=2πNd𝐙N(a,b),\displaystyle d_{S^{1}}(\chi_{N}(a),\chi_{N}(b))=\frac{2\pi}{N}d_{\mathbf{Z}_{N}}(a,b), (1.3)

for any (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N}. For any pair (z,w)μN×μN(z,w)\in\mu_{N}\times\mu_{N}, we define the μN\mu_{N}-interval [z,w]μN[z,w]_{\mu_{N}} by

[z,w]μN\displaystyle[z,w]_{\mu_{N}} =\displaystyle= [z,w]S1μN,\displaystyle[z,w]_{S^{1}}\cap\mu_{N},

and let

[z,w)μN\displaystyle[z,w)_{\mu_{N}} =\displaystyle= [z,w]μN{w},\displaystyle[z,w]_{\mu_{N}}\setminus\{w\},
(z,w]μN\displaystyle(z,w]_{\mu_{N}} =\displaystyle= [z,w]μN{z}.\displaystyle[z,w]_{\mu_{N}}\setminus\{z\}.

Note that, for any (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N}, the map χN\chi_{N} defines a bijection from the 𝐙N\mathbf{Z}_{N}-interval [a,b]𝐙N[a,b]_{\mathbf{Z}_{N}} to the μN\mu_{N}-interval [χN(a),χN(b)]μN[\chi_{N}(a),\chi_{N}(b)]_{\mu_{N}}.

1.2 𝐙N\mathbf{Z}_{N}-average and μN\mu_{N}-average

First we define a map av𝐙N:𝐙N×𝐙N𝐙Nav_{\mathbf{Z}_{N}}:\mathbf{Z}_{N}\times\mathbf{Z}_{N}\rightarrow\mathbf{Z}_{N} by the rule

av𝐙N(a,b)=a+NbNa2\displaystyle av_{\mathbf{Z}_{N}}(a,b)=a+_{N}\left\lfloor\frac{b-_{N}a}{2}\right\rfloor (1.4)

for any (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N}. However we use the notation “ravrav” in [2] to express this map av𝐙Nav_{\mathbf{Z}_{N}}, we rename it in this paper in order to contrast it with a corresponding average on μN\mu_{N}. The latter is defined as follows. For any (z,w)S1×S1(z,w)\in S^{1}\times S^{1}, we denote the midpoint of the circular arc [z,w]S1[z,w]_{S^{1}} by avS1(z,w)av_{S^{1}}(z,w), and call it S1S^{1}-average of zz and ww. Furthermore a kind of rounding-off operator μN\lfloor\cdot\rfloor_{\mu_{N}} on S1S^{1} is introduced as follows. Note that we have the equality

S1=k=0N1[ζNk,ζNk+1)S1,\displaystyle S^{1}=\bigsqcup_{k=0}^{N-1}[\zeta_{N}^{k},\zeta_{N}^{k+1})_{S^{1}},

where ζN=χN(1)\zeta_{N}=\chi_{N}(1). Hence for any zS1z\in S^{1}, there exists a unique k𝐙Nk\in\mathbf{Z}_{N} such that z[ζNk,ζNk+1)S1z\in[\zeta_{N}^{k},\zeta_{N}^{k+1})_{S^{1}}. We put zμN=ζNk\lfloor z\rfloor_{\mu_{N}}=\zeta_{N}^{k} employing this kk. The μN\mu_{N}-average avμN(z,w)av_{\mu_{N}}(z,w) of (z,w)μN×μN(z,w)\in\mu_{N}\times\mu_{N} is define by

avμN(z,w)=avS1(z,w)μN.\displaystyle av_{\mu_{N}}(z,w)=\lfloor av_{S^{1}}(z,w)\rfloor_{\mu_{N}}.

The following proposition shows that the 𝐙N\mathbf{Z}_{N}-average corresponds to the μN\mu_{N}-average through the bijection χN|𝐙N\chi_{N}|_{\mathbf{Z}_{N}}.

Proposition 1.1.

For any (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N}, we have

χN(av𝐙N(a,b))=avμN(χN(a),χN(b)),\displaystyle\chi_{N}(av_{\mathbf{Z}_{N}}(a,b))=av_{\mu_{N}}(\chi_{N}(a),\chi_{N}(b)), (1.5)

namely, the diagram

𝐙N×𝐙NχN×χNμN×μNav𝐙NavμN𝐙NχNμN\displaystyle\begin{CD}\mathbf{Z}_{N}\times\mathbf{Z}_{N}@>{\chi_{N}\times\chi_{N}}>{}>\mu_{N}\times\mu_{N}\\ @V{av_{\mathbf{Z}_{N}}}V{}V@V{av_{\mu_{N}}}V{}V\\ \mathbf{Z}_{N}@>{\chi_{N}}>{}>\mu_{N}\\ \end{CD}

commutes.

Proof.

For the proof, we employ the following lemma.

Lemma 1.1.

For any (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N}, we have

av𝐙N(a,b)={a+b2,if ab,a+b+N2N,if a>b.\displaystyle av_{\mathbf{Z}_{N}}(a,b)=\left\{\begin{array}[]{ll}\left\lfloor\frac{a+b}{2}\right\rfloor,&\mbox{if }a\leq b,\\ \langle\left\lfloor\frac{a+b+N}{2}\right\rfloor\rangle_{N},&\mbox{if }a>b.\\ \end{array}\right. (1.8)

Proof of Lemma 1.1. Suppose that aba\leq b. Then we have

χN(av𝐙N(a,b))\displaystyle\chi_{N}(av_{\mathbf{Z}_{N}}(a,b)) =\displaystyle= χN(a+NbNa2)\displaystyle\chi_{N}(a+_{N}\left\lfloor\frac{b-_{N}a}{2}\right\rfloor)
=\displaystyle= χN(a)χN(ba2)\displaystyle\chi_{N}(a)\chi_{N}(\left\lfloor\frac{b-a}{2}\right\rfloor)
=\displaystyle= χN(a+ba2)\displaystyle\chi_{N}(a+\left\lfloor\frac{b-a}{2}\right\rfloor)
=\displaystyle= χN(a+b2).\displaystyle\chi_{N}(\left\lfloor\frac{a+b}{2}\right\rfloor).

Since 0abN10\leq a\leq b\leq N-1, we have 0a+b2N10\leq\left\lfloor\frac{a+b}{2}\right\rfloor\leq N-1, and hence the assertion (1.5) holds by the bijectivity of χN|𝐙N\chi_{N}|_{\mathbf{Z}_{N}}. When a>ba>b, we have bNa=ba+Nb-_{N}a=b-a+N. Hence we have

χN(av𝐙N(a,b))\displaystyle\chi_{N}(av_{\mathbf{Z}_{N}}(a,b)) =\displaystyle= χN(a+NbNa2)\displaystyle\chi_{N}(a+_{N}\left\lfloor\frac{b-_{N}a}{2}\right\rfloor)
=\displaystyle= χN(a)χN(ba+N2)\displaystyle\chi_{N}(a)\chi_{N}(\left\lfloor\frac{b-a+N}{2}\right\rfloor)
=\displaystyle= χN(a+ba+N2)\displaystyle\chi_{N}(a+\left\lfloor\frac{b-a+N}{2}\right\rfloor)
=\displaystyle= χN(a+b+N2).\displaystyle\chi_{N}(\left\lfloor\frac{a+b+N}{2}\right\rfloor).

This time we need to take the remainder modulo NN of a+b+N2\left\lfloor\frac{a+b+N}{2}\right\rfloor, which need not be smaller than NN. This finishes the proof of Lemma 1.1.
With the help of this lemma, we can prove Proposition 1.1. When aba\leq b, by the definition of avS1av_{S^{1}}, we see that avS1(χN(a),χN(b))=χN(a+b2)av_{S^{1}}(\chi_{N}(a),\chi_{N}(b))=\chi_{N}(\frac{a+b}{2}), since the usual average (a+b)/2(a+b)/2 is the midpoint of the interval [a,b]𝐑[a,b]\subset\mathbf{R}. Therefore avμN(χN(a),χN(b))=χN(a+b2)av_{\mu_{N}}(\chi_{N}(a),\chi_{N}(b))=\chi_{N}(\lfloor\frac{a+b}{2}\rfloor), which is equal to χN(av𝐙N(a,b))\chi_{N}(av_{\mathbf{Z}_{N}}(a,b)) by Lemma 1.1. The case when a>ba>b can be treated similarly also by Lemma 1.1. ∎

As a consequence, we can prove the following.

Proposition 1.2.

(1) For any (z,w)μN×μN(z,w)\in\mu_{N}\times\mu_{N}, we have

avμN(z,w)[z,w]μN.\displaystyle av_{\mu_{N}}(z,w)\in[z,w]_{\mu_{N}}.

(2) For any (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N}, we have

av𝐙N(a,b)[a,b]𝐙N.\displaystyle av_{\mathbf{Z}_{N}}(a,b)\in[a,b]_{\mathbf{Z}_{N}}.
Proof.

(1) The midpoint avS1(z,w)av_{S^{1}}(z,w) belongs to [z,w]S1[z,w]_{S^{1}} by definition, and hence avμN(z,w)=avS1(z,w)μN[z,w]μNav_{\mu_{N}}(z,w)=\lfloor av_{S^{1}}(z,w)\rfloor_{\mu_{N}}\in[z,w]_{\mu_{N}}, since the rounding-off operator cannot go beyond zμNz\in\mu_{N}.
(2) Let c=av𝐙N(a,b)c=av_{\mathbf{Z}_{N}}(a,b). Then if follows from (1.4) that

χN(c)=avμN(χN(a),χN(b)),\displaystyle\chi_{N}(c)=av_{\mu_{N}}(\chi_{N}(a),\chi_{N}(b)),

and hence

χN(c)[χN(a),χN(b)]μN\displaystyle\chi_{N}(c)\in[\chi_{N}(a),\chi_{N}(b)]_{\mu_{N}} (1.9)

by the above (1). Since χN\chi_{N} is a bijection from [a,b]𝐙N[a,b]_{\mathbf{Z}_{N}} onto [χN(a),χN(b)]μN[\chi_{N}(a),\chi_{N}(b)]_{\mu_{N}}, the statement (1.6) implies that c=av𝐙N(a,b)[a,b]𝐙Nc=av_{\mathbf{Z}_{N}}(a,b)\in[a,b]_{\mathbf{Z}_{N}}. This completes the proof. ∎

Noting that the midpoint of [z,w]S1[z,w]_{S^{1}} coincides with ww only when z=wz=w, we obtain the following.

Corollary 1.1.

(1) For any pair (z,w)(z,w) of distinct elements of μN\mu_{N}, we have

avμN(z,w)[z,w)μN.\displaystyle av_{\mu_{N}}(z,w)\in[z,w)_{\mu_{N}}.

(2) For any pair (a,b)(a,b) of distinct elements of 𝐙N\mathbf{Z}_{N}, we have

av𝐙N(a,b)[a,b)𝐙N.\displaystyle av_{\mathbf{Z}_{N}}(a,b)\in[a,b)_{\mathbf{Z}_{N}}.

1.3 The space of rhythms and the discrete average transformation

In this subsection, we recall the definition of the space of rhythms given in [2], and reformulate it by employing the S1S^{1}-average and the μN\mu_{N}-average.

Definition 1.1.

For any positive integer nn with 2nN2\leq n\leq N, an nn-tuple 𝐚=(a0,,an1)𝐙Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{Z}_{N}^{n} is said to be a rhythm modulo NN with nn onsets, if aiaja_{i}\neq a_{j} for iji\neq j, and the following condition holds:

i=0n1(ai+1Nai)=N.\displaystyle\sum_{i=0}^{n-1}(a_{i+1}-_{N}a_{i})=N. (1.10)

(Here the summation on the left hand side means the usual addition of integers, and ana_{n} is set to be a0a_{0}.) We denote by 𝐑Nn\mathbf{R}_{N}^{n} the set of rhythms modulo NN with nn onsets.

This is the original definition given in [2] of the space of rhythms. Note that the condition (1.7) is equivalent to the equality

𝐙N=i=0n1[ai,ai+n1)𝐙N.\displaystyle\mathbf{Z}_{N}=\bigsqcup_{i=0}^{n-1}[a_{i},a_{i+_{n}1})_{\mathbf{Z}_{N}}. (1.11)

Applying the map χN\chi_{N} to the both sides of (1.8) we obtain another equivalent condition

S1=i=0n1[χN(ai),χN(ai+n1))S1,\displaystyle S^{1}=\bigsqcup_{i=0}^{n-1}[\chi_{N}(a_{i}),\chi_{N}(a_{i+_{n}1}))_{S^{1}}, (1.12)

which says that the ordered set (χN(a0),,χN(an1))(\chi_{N}(a_{0}),\cdots,\chi_{N}(a_{n-1})) constitutes a simple convex nn-gon inscribed in S1S^{1}. The condition (1.9) clarifies the meaning of the algebraic condition (1.7) in more geometric terms, and will be employed frequently afterwards.
In order to make our transition to the Boolean world smooth, we need, however, to define the space 𝐑N1\mathbf{R}_{N}^{1} of rhythms with 1 onset as well as 𝐑N0\mathbf{R}_{N}^{0}. The following definition fills the gap:

Definition 1.2.

Any singleton (a)(a) of a𝐙Na\in\mathbf{Z}_{N} is regarded as a rhythm with 1 onset. The empty set ϕ\phi is considered to be the unique rhythm with 0 onset. Accordingly we define

𝐑N1\displaystyle\mathbf{R}_{N}^{1} =\displaystyle= {(a);a𝐙N},\displaystyle\{(a);a\in\mathbf{Z}_{N}\},
𝐑N0\displaystyle\mathbf{R}_{N}^{0} =\displaystyle= {ϕ}.\displaystyle\{\phi\}.

Furthermore we define the total space of rhythms 𝐑N\mathbf{R}_{N} with NN beats by

𝐑N=n=0N𝐑Nn.\displaystyle\mathbf{R}_{N}=\bigcup_{n=0}^{N}\mathbf{R}_{N}^{n}.

The discrete average transformation on 𝐑N\mathbf{R}_{N}, which is the central topic in [2], is defined as follows.

Definition 1.3.

For a rhythm 𝐚=(a0,,an1)𝐑Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n} with 2nN2\leq n\leq N, we define its discrete average transformation Rav(𝐚)Rav(\mathbf{a}) by

Rav(𝐚)=(av𝐙N(a0,a1),av𝐙N(a1,a2),,av𝐙N(an1,a0)).\displaystyle Rav(\mathbf{a})=(av_{\mathbf{Z}_{N}}(a_{0},a_{1}),av_{\mathbf{Z}_{N}}(a_{1},a_{2}),\cdots,av_{\mathbf{Z}_{N}}(a_{n-1},a_{0})). (1.13)

Since we have enlarged the space of rhythms in Definition 1.2, we need to define discrete average on 𝐑N1\mathbf{R}_{N}^{1} and 𝐑N0\mathbf{R}_{N}^{0} too. This is accomplished simply as follows:

Definition 1.4.

For any (a)𝐑N1(a)\in\mathbf{R}_{N}^{1}, we define

Rav((a))=(a),\displaystyle Rav((a))=(a), (1.14)

and on 𝐑N0\mathbf{R}_{N}^{0} we put

Rav(ϕ)=ϕ.\displaystyle Rav(\phi)=\phi. (1.15)
Remark 1.1.

It follows from the definition that av𝐙N(a,a)=0av_{\mathbf{Z}_{N}}(a,a)=0 holds for any a𝐙Na\in\mathbf{Z}_{N}. Hence our definition (1.11) is natural.

The main purpose of this subsection is to show that the discrete average transformation RavRav maps 𝐑N\mathbf{R}_{N} into itself. This fact is proved in [2] through an algebraic argument. We, however, give below a simpler and more geometric proof.

Proposition 1.3.

For any 𝐚𝐑Nn\mathbf{a}\in\mathbf{R}_{N}^{n}, we have Rav(𝐚)𝐑NnRav(\mathbf{a})\in\mathbf{R}_{N}^{n}.

Proof.

We may assume that n2n\geq 2, since the assertion follows directly from Definition 1.4 when n=0n=0 or n=1n=1. Let Rav(𝐚)=𝐛=(b0,b1,,bn1)Rav(\mathbf{a})=\mathbf{b}=(b_{0},b_{1},\cdots,b_{n-1}). It follows from the assumption that we have

S1=i=0n1[χN(ai),χN(ai+n1))S1.\displaystyle S^{1}=\bigsqcup_{i=0}^{n-1}[\chi_{N}(a_{i}),\chi_{N}(a_{i+_{n}1}))_{S^{1}}.

Furthermore it follows from Proposition 1.1 and Corollary 1.1 that

χN(bi)[χN(ai),χN(ai+n1))μN.\displaystyle\chi_{N}(b_{i})\in[\chi_{N}(a_{i}),\chi_{N}(a_{i+_{n}1}))_{\mu_{N}}.

Hence we have

S1=i=0n1[χN(bi),χN(bi+n1))S1.\displaystyle S^{1}=\bigsqcup_{i=0}^{n-1}[\chi_{N}(b_{i}),\chi_{N}(b_{i+_{n}1}))_{S^{1}}.

Therefore the nn-tuple 𝐛\mathbf{b} satisfies the condition (1.9), and hence 𝐛𝐑Nn\mathbf{b}\in\mathbf{R}_{N}^{n}. This completes the proof. ∎

2 From modulo-NN-world to the Boolean world

The purpose of this section is to construct a Boolean counterpart of the discrete average transformation RavRav on 𝐑N\mathbf{R}_{N}. We accomplish this by introducing two intermediate spaces 𝐑¯N\overline{\mathbf{R}}_{N} and 𝐈N\mathbf{I}_{N}, which lie between 𝐑N\mathbf{R}_{N} and our destination, the Boolean space 𝐁N=𝐅2N\mathbf{B}_{N}=\mathbf{F}_{2}^{N}. We define a certain average map on each of the new spaces, and build three bridges π,s,ItoB\pi,s,ItoB between each pair of the four spaces, which make the following diagram commutative.

𝐑Nπ𝐑¯Ns𝐈NItoB𝐁NRavR¯avIavBav𝐑Nπ𝐑¯Ns𝐈NItoB𝐁N\displaystyle\begin{CD}\mathbf{R}_{N}@>{\pi}>{}>\overline{\mathbf{R}}_{N}@>{s}>{}>\mathbf{I}_{N}@>{ItoB}>{}>\mathbf{B}_{N}\\ @V{Rav}V{}V@V{\overline{R}av}V{}V@V{Iav}V{}V@V{Bav}V{}V\\ \mathbf{R}_{N}@>{\pi}>{}>\overline{\mathbf{R}}_{N}@>{s}>{}>\mathbf{I}_{N}@>{ItoB}>{}>\mathbf{B}_{N}\\ \end{CD}

2.1 Quotient map 𝐑Nn𝐑¯Nn\mathbf{R}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n} and its section

A self map ”rotrot” on 𝐑Nn\mathbf{R}_{N}^{n} is introduced in [2]. Let us recall the definition and supply it with additional versions for the case n=0,1n=0,1:

Definition 2.1.

For any 𝐚=(a0,a1,,an1)𝐑Nn\mathbf{a}=(a_{0},a_{1},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n} with 2nN2\leq n\leq N, let

rot(𝐚)=(an1,a0,,an2).\displaystyle rot(\mathbf{a})=(a_{n-1},a_{0},\cdots,a_{n-2}). (2.1)

When n=0,1n=0,1, it is defined to be the identity, namely,

rot((a0))\displaystyle rot((a_{0})) =\displaystyle= (a0) for any (a0)𝐑N1,\displaystyle(a_{0})\mbox{ for any }(a_{0})\in\mathbf{R}_{N}^{1}, (2.2)
rot(ϕ)\displaystyle rot(\phi) =\displaystyle= ϕ.\displaystyle\phi. (2.3)

Note that rot(𝐚)𝐑Nnrot(\mathbf{a})\in\mathbf{R}_{N}^{n} for any 𝐚𝐑Nn\mathbf{a}\in\mathbf{R}_{N}^{n}. We define an equivalence relation on 𝐑Nn\mathbf{R}_{N}^{n} employing this map as follows:

Definition 2.2.

For any pair 𝐚,𝐛𝐑Nn\mathbf{a},\mathbf{b}\in\mathbf{R}_{N}^{n}, they are said to be equivalent if there exists an integer kk such that

rotk(𝐚)=𝐛.\displaystyle rot^{k}(\mathbf{a})=\mathbf{b}.

Since the map rotrot is bijective, this defines an equivalence relation on 𝐑Nn\mathbf{R}_{N}^{n}, which we denote by ”rot\sim_{rot}”.

Definition 2.3.

We denote by 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} the quotient of 𝐑Nn\mathbf{R}_{N}^{n} by the equivalence relation rot\sim_{rot}:

𝐑¯Nn=𝐑Nn/rot,\displaystyle\overline{\mathbf{R}}_{N}^{n}=\mathbf{R}_{N}^{n}/\sim_{rot},

and by π:𝐑Nn𝐑¯Nn\pi:\mathbf{R}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n} the natural projection.

Remark 2.1.

Our definition of ”rotrot” here is actually the inverse of the one used in [2][2]. The reason why we adopt the inverse is that it makes several formulas in this paper look more natural. The equivalence relation ”rot\sim_{rot}”, however, is actually defined through the action of the group 𝐙n\mathbf{Z}_{n}, generated by rotrot, hence the quotient space 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} is exactly the same as the one defined in [[loc.cit.]].

Remark 2.2.

When n=0,1n=0,1, the space 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} is nothing other than 𝐑Nn\mathbf{R}_{N}^{n}, since rotrot is the identity on 𝐑N1\mathbf{R}_{N}^{1} or on 𝐑N0\mathbf{R}_{N}^{0}.

It follows from Definition 1.3, 1.4, and 2.1 that the two self maps rotrot and RavRav on 𝐑Nn\mathbf{R}_{N}^{n} commute with each other for any n0n\geq 0 :

rotRav=Ravrot,\displaystyle rot\circ Rav=Rav\circ rot,

namely, we have the following commutative diagram:

𝐑Nnrot𝐑NnRavRav𝐑Nnrot𝐑Nn\displaystyle\begin{CD}\mathbf{R}_{N}^{n}@>{rot}>{}>\mathbf{R}_{N}^{n}\\ @V{Rav}V{}V@V{Rav}V{}V\\ \mathbf{R}_{N}^{n}@>{rot}>{}>\mathbf{R}_{N}^{n}\\ \end{CD}

Therefore the map RavRav descends to a self map R¯av\overline{R}av on 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n}, namely, the equality

πRav=R¯avπ.\displaystyle\pi\circ Rav=\overline{R}av\circ\pi. (2.4)

holds. Hence we have the following commutative diagram:

𝐑Nnπ𝐑¯NnRavR¯av𝐑Nnπ𝐑¯Nn\displaystyle\begin{CD}\mathbf{R}_{N}^{n}@>{\pi}>{}>\overline{\mathbf{R}}_{N}^{n}\\ @V{Rav}V{}V@V{\overline{R}av}V{}V\\ \mathbf{R}_{N}^{n}@>{\pi}>{}>\overline{\mathbf{R}}_{N}^{n}\\ \end{CD}

2.2 Construction of a section of π\pi

The purpose of this subsection is to construct a section of the quotient map π:𝐑Nn𝐑¯Nn\pi:\mathbf{R}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n}. Here the notion of jumping number, introduced in [2], plays a crucial role.

Definition 2.4.

When n2n\geq 2, for any 𝐚=(a0,a1,,an1)𝐑Nn\mathbf{a}=(a_{0},a_{1},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n}, there is a unique number j𝐙nj\in\mathbf{Z}_{n} such that ajn1>aja_{j-_{n}1}>a_{j} and that akn1<aka_{k-_{n}1}<a_{k} holds for every other k𝐙n{j}k\in\mathbf{Z}_{n}\setminus\{j\}. The number jj is called the jumping number of 𝐚\mathbf{a}. When n=1n=1, the jumping number of an arbitrary element (a0)(a_{0}) of 𝐑N1\mathbf{R}_{N}^{1} is defined to be 0, In any case it is denoted by j(𝐚)j(\mathbf{a}).

Remark 2.3.

Our definition of the jumping number increases the corresponding number defined in [2][2] by one. One of the main reasons why we adopt this version is that we want every element in the basic set 𝐈Nn\mathbf{I}_{N}^{n}, defined in Definition 2.5 later, of increasing sequences to acquire the jumping number 0. Since we derive several results in this paper by choosing from 𝐈Nn\mathbf{I}_{N}^{n} a representative of the class in 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n}, this convention makes our description simpler.

We can characterize the jumping number in geometric and intuitive terms by employing the half-closed intervals.

Proposition 2.1.

Assume that n2n\geq 2. For any 𝐚=(a0,a1,,an1)𝐑Nn\mathbf{a}=(a_{0},a_{1},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n}, the jumping number j(𝐚)j(\mathbf{a}) is equal to the unique number i𝐙ni\in\mathbf{Z}_{n} such that the half-open interval (χN(ain1),χN(ai)]S1(\chi_{N}(a_{i-_{n}1}),\chi_{N}(a_{i})]_{S^{1}} contains 1S11\in S^{1}.

Proof.

Replacing the left-closed and right-open intervals in the condition (1.9) by the left-open and right-closed intervals, we obtain an equivalent condition

S1=i=0n1(χN(ain1),χN(ai)]S1.\displaystyle S^{1}=\bigsqcup_{i=0}^{n-1}(\chi_{N}(a_{i-_{n}1}),\chi_{N}(a_{i})]_{S^{1}}.

Hence there exists a unique member i𝐙ni\in\mathbf{Z}_{n} such that

1(χN(ain1),χN(ai)]S1.\displaystyle 1\in(\chi_{N}(a_{i-_{n}1}),\chi_{N}(a_{i})]_{S^{1}}.

This condition is equivalent to the condition that

0(ain1,ai]𝐙N,\displaystyle 0\in(a_{i-_{n}1},a_{i}]_{\mathbf{Z}_{N}},

which holds if and only if j(𝐚)=ij(\mathbf{a})=i. This completes the proof. ∎

By the very definition of the jumping number, we have the following:

Proposition 2.2.

For any 𝐚𝐑Nn\mathbf{a}\in\mathbf{R}_{N}^{n}, we have

j(rot(𝐚))=j(𝐚)+n1.\displaystyle j(rot(\mathbf{a}))=j(\mathbf{a})+_{n}1.
Proof.

Let 𝐚=(a0,,an1)\mathbf{a}=(a_{0},\cdots,a_{n-1}) and rot(𝐚)=𝐛=(b0,,bn1)rot(\mathbf{a})=\mathbf{b}=(b_{0},\cdots,b_{n-1}) so that

bi=ain1\displaystyle b_{i}=a_{i-_{n}1} (2.5)

holds for any i𝐙ni\in\mathbf{Z}_{n}. If we put j=j(𝐚)j=j(\mathbf{a}), then we have

0aj<aj+n1<<ajn1N1.\displaystyle 0\leq a_{j}<a_{j+_{n}1}<\cdots<a_{j-_{n}1}\leq N-1.

This series of inequalities is equivalent to

0bj+n1<bj+n2<<bjN1\displaystyle 0\leq b_{j+_{n}1}<b_{j+_{n}2}<\cdots<b_{j}\leq N-1

by (2.5). Hence j(𝐛)=j+n1j(\mathbf{b})=j+_{n}1. This completes the proof. ∎

As a direct consequence, we obtain the following:

Corollary 2.1.

For any x𝐑¯Nnx\in\overline{\mathbf{R}}_{N}^{n}, the set of jumping numbers of the elements in the fiber π1(x)\pi^{-1}(x) coincides with the whole 𝐙n\mathbf{Z}_{n}.

As a result, we can choose from each fiber of π\pi a unique rhythm with its jumping number equal to zero. For this reason, we introduce the following:

Definition 2.5.

We denote by 𝐈Nn\mathbf{I}_{N}^{n} the subset of 𝐑Nn\mathbf{R}_{N}^{n} which consists of every element (a0,,an1)𝐑Nn(a_{0},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n} such that

0a0<<an1N1,\displaystyle 0\leq a_{0}<\cdots<a_{n-1}\leq N-1, (2.6)

namely, we put

𝐈Nn={𝐚𝐑Nn;j(𝐚)=0}.\displaystyle\mathbf{I}_{N}^{n}=\{\mathbf{a}\in\mathbf{R}_{N}^{n};j(\mathbf{a})=0\}.
Remark 2.4.

The condition (2.6) says that the sequence (a0,,an1)(a_{0},\cdots,a_{n-1}) is increasing. We adopt its initial as the name of the space.

Proposition 2.2 enables one to find the unique element in π1(x)𝐈Nn\pi^{-1}(x)\cap\mathbf{I}_{N}^{n} algorhythmically:

Proposition 2.3.

For any 𝐚𝐑Nn\mathbf{a}\in\mathbf{R}_{N}^{n}, let

pr𝐈(𝐚)=rotnnj(𝐚)(𝐚).\displaystyle pr_{\mathbf{I}}(\mathbf{a})=rot^{n-_{n}j(\mathbf{a})}(\mathbf{a}). (2.7)

Then it defines a contraction map pr𝐈:𝐑Nn𝐈Nnpr_{\mathbf{I}}:\mathbf{R}_{N}^{n}\rightarrow\mathbf{I}_{N}^{n} such that

pr𝐈ι𝐈=id𝐈Nn,\displaystyle pr_{\mathbf{I}}\circ\iota_{\mathbf{I}}=id_{\mathbf{I}_{N}^{n}}, (2.8)

where ι𝐈:𝐈Nn𝐑Nn\iota_{\mathbf{I}}:\mathbf{I}_{N}^{n}\rightarrow\mathbf{R}_{N}^{n} denotes the natural inclusion map.

Proof.

It follows form Proposition 2.2 and (2.7) that

j(pr𝐈(𝐚))=j(𝐚)+n(nnj(𝐚))=0.\displaystyle j(pr_{\mathbf{I}}(\mathbf{a}))=j(\mathbf{a})+_{n}(n-_{n}j(\mathbf{a}))=0.

Hence pr𝐈(𝐚)𝐈Nnpr_{\mathbf{I}}(\mathbf{a})\in\mathbf{I}_{N}^{n}. Furthermore, if 𝐚𝐈Nn\mathbf{a}\in\mathbf{I}_{N}^{n}, then j(𝐚)=0j(\mathbf{a})=0, and hence pr𝐈(𝐚)=𝐚pr_{\mathbf{I}}(\mathbf{a})=\mathbf{a}. This shows that the equality (2.8) holds true. This completes the proof. ∎

Thus we obtain a section of the quotient map π\pi:

Proposition 2.4.

For any 𝐚𝐑Nn\mathbf{a}\in\mathbf{R}_{N}^{n}, let

s(π(𝐚))=pr𝐈(𝐚).\displaystyle s(\pi(\mathbf{a}))=pr_{\mathbf{I}}(\mathbf{a}).

Then ss defines a section of the quotient map π:𝐑Nn𝐑¯Nn\pi:\mathbf{R}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n}, and its image coincides with the set 𝐈Nn\mathbf{I}_{N}^{n}. In particular 𝐈Nn\mathbf{I}_{N}^{n} and 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} are in one-to-one correspondence.

Remark 2.5.

When n=0,1n=0,1, we define s:𝐑¯Nn𝐑Nns:\overline{\mathbf{R}}_{N}^{n}\rightarrow\mathbf{R}_{N}^{n} to be the identity map. (See Remark 2.2.)

2.3 From 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} to Boolean world via 𝐈Nn\mathbf{I}_{N}^{n}

In a standard way, the set 𝐈Nn\mathbf{I}_{N}^{n} is connected directly to the Boolean world. For later use we introduce some notations and make the connection clear:

Definition 2.6.

Let 𝐁N\mathbf{B}_{N} denote the NN-dimensional vector space 𝐅2N\mathbf{F}_{2}^{N} over the prime field 𝐅2={0,1}\mathbf{F}_{2}=\{0,1\}. For any vector 𝐯𝐁N\mathbf{v}\in\mathbf{B}_{N}, we denote by supp(𝐯)supp(\mathbf{v}) the set of indices of the non-zero entries of 𝐯\mathbf{v},

supp(𝐯)={i𝐙N;vi=1}𝐙N,\displaystyle supp(\mathbf{v})=\{i\in\mathbf{Z}_{N};v_{i}=1\}\subset\mathbf{Z}_{N},

and let

𝐁Nn={𝐯𝐁N;|supp(𝐯)|=n}\displaystyle\mathbf{B}_{N}^{n}=\{\mathbf{v}\in\mathbf{B}_{N};|supp(\mathbf{v})|=n\}

for any nn with 0nN0\leq n\leq N, so that

𝐁N=n=0N𝐁Nn.\displaystyle\mathbf{B}_{N}=\bigcup_{n=0}^{N}\mathbf{B}_{N}^{n}.

First we relate every element in 𝐑Nn\mathbf{R}_{N}^{n} to 𝐁Nn\mathbf{B}_{N}^{n} by taking its characteristic function:

Definition 2.7.

With any 𝐚𝐑Nn\mathbf{a}\in\mathbf{R}_{N}^{n}, we associate a vector

RtoB(𝐚)=(v0,,vN1)𝐁N\displaystyle RtoB(\mathbf{a})=(v_{0},\cdots,v_{N-1})\in\mathbf{B}_{N}

by the rule

vi={1, if i𝐚,0, if i𝐚.\displaystyle v_{i}=\left\{\begin{array}[]{ll}1,&\mbox{ if $i\in\mathbf{a}$,}\\ 0,&\mbox{ if $i\not\in\mathbf{a}$.}\\ \end{array}\right.

This gives us the map

RtoB:𝐑Nn𝐁Nn.\displaystyle RtoB:\mathbf{R}_{N}^{n}\rightarrow\mathbf{B}_{N}^{n}.

Note that the following simple but important equality

RtoBrot=RtoB.\displaystyle RtoB\circ rot=RtoB. (2.10)

holds, since the map rotrot does not change the contents of any rhythms. It follows from (2.9) that

RtoBpr𝐈=RtoB,\displaystyle RtoB\circ pr_{\mathbf{I}}=RtoB, (2.11)

since pr𝐈pr_{\mathbf{I}} is a power of rotrot by the definition (2.7).
Next, based on the map RtoBRtoB, we relate 𝐈Nn\mathbf{I}_{N}^{n} with 𝐁Nn\mathbf{B}_{N}^{n}:

Definition 2.8.

For any nn-element subset SS of 𝐙N\mathbf{Z}_{N}, by arranging its elements in increasing order

0s0<s1<<sn1N1,\displaystyle 0\leq s_{0}<s_{1}<\cdots<s_{n-1}\leq N-1,

we define

ord(S)=(s0,,sn1).\displaystyle ord(S)=(s_{0},\cdots,s_{n-1}).

We put

ItoB\displaystyle ItoB =\displaystyle= RtoBι𝐈:𝐈Nn𝐁Nn,\displaystyle RtoB\circ\iota_{\mathbf{I}}:\mathbf{I}_{N}^{n}\rightarrow\mathbf{B}_{N}^{n}, (2.12)
BtoI\displaystyle BtoI =\displaystyle= ordsupp:𝐁Nn𝐈Nn.\displaystyle ord\circ supp:\mathbf{B}_{N}^{n}\rightarrow\mathbf{I}_{N}^{n}. (2.13)

Since one can check that

ItoBBtoI\displaystyle ItoB\circ BtoI =\displaystyle= id𝐁Nn,\displaystyle id_{\mathbf{B}_{N}^{n}}, (2.14)
BtoIItoB\displaystyle BtoI\circ ItoB =\displaystyle= id𝐈Nn,\displaystyle id_{\mathbf{I}_{N}^{n}}, (2.15)

𝐈Nn\mathbf{I}_{N}^{n} and 𝐁Nn\mathbf{B}_{N}^{n} are in one-to-one correspondence.
Finally, we relate the quotient space 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} with 𝐁Nn\mathbf{B}_{N}^{n}.

Proposition 2.5.

Let R¯toB:𝐑¯Nn𝐁Nn\overline{R}toB:\overline{\mathbf{R}}_{N}^{n}\rightarrow\mathbf{B}_{N}^{n} be the map defined by

R¯toB=ItoBs,\displaystyle\overline{R}toB=ItoB\circ s, (2.16)

and let BtoR¯:𝐁Nn𝐑¯NnBto\overline{R}:\mathbf{B}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n} be the map defined by

BtoR¯=πBtoI.\displaystyle Bto\overline{R}=\pi\circ BtoI. (2.17)

These maps are inverse to each other:

BtoR¯R¯toB\displaystyle Bto\overline{R}\circ\overline{R}toB =\displaystyle= id𝐑¯Nn,\displaystyle id_{\overline{\mathbf{R}}_{N}^{n}}, (2.18)
R¯toBBtoR¯\displaystyle\overline{R}toB\circ Bto\overline{R} =\displaystyle= id𝐁Nn.\displaystyle id_{\mathbf{B}_{N}^{n}}. (2.19)

In particular, the sets 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} and 𝐁Nn\mathbf{B}_{N}^{n} are in one-to-one correspondence.

Proof.

First we compute BtoR¯R¯toBBto\overline{R}\circ\overline{R}toB as follows:

BtoR¯R¯toB\displaystyle Bto\overline{R}\circ\overline{R}toB =\displaystyle= (πBtoI)(ItoBs)\displaystyle(\pi\circ BtoI)\circ(ItoB\circ s)
=\displaystyle= πs (by (2.14))\displaystyle\pi\circ s\hskip 42.67912pt\mbox{ (by (2.14))}
=\displaystyle= id𝐑¯Nn,\displaystyle id_{\overline{\mathbf{R}}_{N}^{n}},

hence we have (2.17). For the validity of (2.18), we note that, by Proposition 2.4, the composite map sπ:𝐑Nn𝐑Nns\circ\pi:\mathbf{R}_{N}^{n}\rightarrow\mathbf{R}_{N}^{n} becomes the identity when restricted to the subset 𝐈Nn\mathbf{I}_{N}^{n}. Hence

R¯toBBtoR¯\displaystyle\overline{R}toB\circ Bto\overline{R} =\displaystyle= (ItoBs)(πBtoI)\displaystyle(ItoB\circ s)\circ(\pi\circ BtoI)
=\displaystyle= ItoBBtoI\displaystyle ItoB\circ BtoI
=\displaystyle= id𝐁Nn,\displaystyle id_{\mathbf{B}_{N}^{n}},

the last equality being a consequence of (2.13). This completes the proof. ∎

The two maps RtoBRtoB and R¯toB\overline{R}toB are related in a natural way:

Proposition 2.6.

We have

R¯toBπ=RtoB.\displaystyle\overline{R}toB\circ\pi=RtoB.
Proof.

Combining several defining equations, we compute as follows:

R¯toBπ\displaystyle\overline{R}toB\circ\pi =\displaystyle= ItoBsπ(by (2.15))\displaystyle ItoB\circ s\circ\pi\hskip 28.45274pt(\mbox{by }(2.15))
=\displaystyle= ItoBpr𝐈(by Proposition 2.4)\displaystyle ItoB\circ pr_{\mathbf{I}}\hskip 34.1433pt(\mbox{by Proposition }2.4)
=\displaystyle= RtoBι𝐈pr𝐈(by (2.11))\displaystyle RtoB\circ\iota_{\mathbf{I}}\circ pr_{\mathbf{I}}\hskip 17.07164pt(\mbox{by }(2.11))
=\displaystyle= RtoBpr𝐈\displaystyle RtoB\circ pr_{\mathbf{I}}
=\displaystyle= RtoB.(by (2.10))\displaystyle RtoB.\hskip 56.9055pt(\mbox{by }(2.10))

This finishes the proof. ∎

2.4 Counterpart of RavRav on 𝐈Nn\mathbf{I}_{N}^{n}

The discrete average map Rav:𝐑Nn𝐑NnRav:\mathbf{R}_{N}^{n}\rightarrow\mathbf{R}_{N}^{n}, to our regret, does not keep the subspace 𝐈Nn𝐑Nn\mathbf{I}_{N}^{n}\subset\mathbf{R}_{N}^{n} invariant. For example, for 𝐚=(2,3,7)𝐈83\mathbf{a}=(2,3,7)\in\mathbf{I}_{8}^{3}, its discrete average is

Rav(2,3,7)=(2,5,0),\displaystyle Rav(2,3,7)=(2,5,0),

which does not belong to 𝐈83\mathbf{I}_{8}^{3}. In order to overcome this inconvenience and to define a self map IavIav on 𝐈Nn\mathbf{I}_{N}^{n}, we introduce the following notion:

Definition 2.9.

A rhythm 𝐚=(a0,,an1)𝐈Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{I}_{N}^{n} is said to be proper, if

Nan1>a0.\displaystyle N-a_{n-1}>a_{0}. (2.20)

If it is not proper, it is called improper.

The following proposition relates the properness of an increasing rhythm with its behavior under the discrete average map.

Proposition 2.7.

(1) For an increasing rhythm 𝐚=(a0,,an1)𝐈Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{I}_{N}^{n}, it is proper if and only if Rav(𝐚)𝐈NnRav(\mathbf{a})\in\mathbf{I}_{N}^{n}.
(2) If 𝐚=(a0,,an1)𝐈Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{I}_{N}^{n} is improper, then rot(Rav(𝐚))𝐈Nnrot(Rav(\mathbf{a}))\in\mathbf{I}_{N}^{n}.

Proof.

(1) Let Rav(𝐚)=𝐛=(b0,b1,,bn1)Rav(\mathbf{a})=\mathbf{b}=(b_{0},b_{1},\cdots,b_{n-1}). By Corollary 1.1, (2), we have a series of inequalities

0a0b0<a1b1<a2<an2bn2<an1N1.\displaystyle 0\leq a_{0}\leq b_{0}<a_{1}\leq b_{1}<a_{2}\leq\cdots<a_{n-2}\leq b_{n-2}<a_{n-1}\leq N-1. (2.21)

Therefore only the behavior of bn1=av𝐙N(an1,a0)b_{n-1}=av_{\mathbf{Z}_{N}}(a_{n-1},a_{0}) determines whether 𝐛\mathbf{b} belongs to 𝐈Nn\mathbf{I}_{N}^{n} or not. Note that the condition (2.19) is expressed as

d𝐙N(an1,0)>d𝐙N(0,a0).\displaystyle d_{\mathbf{Z}_{N}}(a_{n-1},0)>d_{\mathbf{Z}_{N}}(0,a_{0}).

Since the two distances d𝐙Nd_{\mathbf{Z}_{N}} and dS1d_{S^{1}} are in proportion through the map χN\chi_{N} by (1.2), this inequality is equivalent to

dS1(χN(an1),1)>dS1(1,χN(a0)).\displaystyle d_{S^{1}}(\chi_{N}(a_{n-1}),1)>d_{S^{1}}(1,\chi_{N}(a_{0})).

This condition holds if and only if the midpoint avS1(χN(an1),χN(a0))av_{S^{1}}(\chi_{N}(a_{n-1}),\chi_{N}(a_{0})) lies on the S1S^{1}-interval [χN(an1),1)S1[\chi_{N}(a_{n-1}),1)_{S^{1}}. Moreover the last condition is equivalent to avμN(χN(an1),χN(a0))[χN(an1),1)S1av_{\mu_{N}}(\chi_{N}(a_{n-1}),\chi_{N}(a_{0}))\in[\chi_{N}(a_{n-1}),1)_{S^{1}}. It follows from Proposition 1.1 that this is equivalent to the condition that av𝐙N(an1,a0)[an1,0)𝐙Nav_{\mathbf{Z}_{N}}(a_{n-1},a_{0})\in[a_{n-1},0)_{\mathbf{Z}_{N}}, which together with (2.20) says that Rav(𝐚)𝐈NnRav(\mathbf{a})\in\mathbf{I}_{N}^{n}.
(2) If 𝐚=(a0,,an1)𝐈Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{I}_{N}^{n} is improper, then it follows from the proof for the point (1) that

0bn1<a0b0<a1b1<a2bn2<an1N1.\displaystyle 0\leq b_{n-1}<a_{0}\leq b_{0}<a_{1}\leq b_{1}<a_{2}\leq\cdots\leq b_{n-2}<a_{n-1}\leq N-1. (2.22)

Therefore rot(𝐛)=(bn1,b0,b1,,bn2)𝐈Nnrot(\mathbf{b})=(b_{n-1},b_{0},b_{1},\cdots,b_{n-2})\in\mathbf{I}_{N}^{n}. This completes the proof. ∎

The following lemma is a rephrasing of Proposition 2.7.

Corollary 2.2.

For any increasing rhythm 𝐚𝐈Nn\mathbf{a}\in\mathbf{I}_{N}^{n}, we introduce the number p(𝐚){0,1}p(\mathbf{a})\in\{0,1\} by the rule

p(𝐚)={0, if 𝐚 is proper,1, if 𝐚 is improper.\displaystyle p(\mathbf{a})=\left\{\begin{array}[]{ll}0,&\mbox{ if }\mathbf{a}\mbox{ is proper,}\\ 1,&\mbox{ if }\mathbf{a}\mbox{ is improper.}\\ \end{array}\right.

Then we have

rotp(𝐚)(Rav(𝐚))𝐈Nn.\displaystyle rot^{p(\mathbf{a})}(Rav(\mathbf{a}))\in\mathbf{I}_{N}^{n}.

Now we can define 𝐈\mathbf{I}-version of the discrete average transformation as follows:

Definition 2.10.

For any 𝐚𝐈Nn\mathbf{a}\in\mathbf{I}_{N}^{n}, let

Iav(𝐚)=rotp(𝐚)(Rav(𝐚)).\displaystyle Iav(\mathbf{a})=rot^{p(\mathbf{a})}(Rav(\mathbf{a})).

It follows from Corollary 2.2 that IavIav defines a self map on 𝐈Nn\mathbf{I}_{N}^{n}:

Iav:𝐈Nn𝐈Nn.\displaystyle Iav:\mathbf{I}_{N}^{n}\rightarrow\mathbf{I}_{N}^{n}.

The following proposition shows that IavIav is compatible with R¯av\overline{R}av:

Proposition 2.8.

We have the following equality

π|𝐈NnIav=R¯avπ|𝐈Nn,\displaystyle\pi|_{\mathbf{I}_{N}^{n}}\circ Iav=\overline{R}av\circ\pi|_{\mathbf{I}_{N}^{n}}, (2.24)

namely, the diagram

𝐈Nnπ|𝐈Nn𝐑¯NnIavR¯av𝐈Nnπ|𝐈Nn𝐑¯Nn\displaystyle\begin{CD}\mathbf{I}_{N}^{n}@>{\pi|_{\mathbf{I}_{N}^{n}}}>{}>\overline{\mathbf{R}}_{N}^{n}\\ @V{Iav}V{}V@V{\overline{R}av}V{}V\\ \mathbf{I}_{N}^{n}@>{\pi|_{\mathbf{I}_{N}^{n}}}>{}>\overline{\mathbf{R}}_{N}^{n}\\ \end{CD}

commutes.

Proof.

Let 𝐚\mathbf{a} be an arbitrary element of 𝐈Nn\mathbf{I}_{N}^{n}. When 𝐚\mathbf{a} is proper, it follows from Definition 2.10 and Corollary 2.2 that Iav(𝐚)=Rav(𝐚)Iav(\mathbf{a})=Rav(\mathbf{a}). Hence the equality (2.22) follows from (2.4). When 𝐚\mathbf{a} is improper, it follows from Definition 2.10 that Iav(𝐚)=rot(Rav(𝐚))Iav(\mathbf{a})=rot(Rav(\mathbf{a})). Hence we have

(π|𝐈NnIav)(𝐚)\displaystyle(\pi|_{\mathbf{I}_{N}^{n}}\circ Iav)(\mathbf{a}) =\displaystyle= π|𝐈Nn(rot(Rav(𝐚)))\displaystyle\pi|_{\mathbf{I}_{N}^{n}}(rot(Rav(\mathbf{a})))
=\displaystyle= π(rot(Rav(𝐚)))\displaystyle\pi(rot(Rav(\mathbf{a})))
=\displaystyle= π(Rav(𝐚))(by the definition of the quotient map π)\displaystyle\pi(Rav(\mathbf{a}))\hskip 14.22636pt(\mbox{by the definition of the quotient map }\pi)
=\displaystyle= (R¯avπ)(𝐚)(by (2.4))\displaystyle(\overline{R}av\circ\pi)(\mathbf{a})\hskip 8.53581pt(\mbox{by }(2.4))
=\displaystyle= (R¯avπ|𝐈Nn)(𝐚).\displaystyle(\overline{R}av\circ\pi|_{\mathbf{I}_{N}^{n}})(\mathbf{a}).

This completes the proof. ∎

Since π|𝐈Nn:𝐈Nn𝐑¯Nn\pi|_{\mathbf{I}_{N}^{n}}:\mathbf{I}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n} and s:𝐑¯Nn𝐈Nns:\overline{\mathbf{R}}_{N}^{n}\rightarrow\mathbf{I}_{N}^{n} are inverse to each other, this proposition implies the following:

Corollary 2.3.

We have the following equality

Iavs=sR¯av,\displaystyle Iav\circ s=s\circ\overline{R}av, (2.25)

namely, the diagram

𝐈Nns𝐑¯NnIavR¯av𝐈Nns𝐑¯Nn\displaystyle\begin{CD}\mathbf{I}_{N}^{n}@<{}<{s}<\overline{\mathbf{R}}_{N}^{n}\\ @V{Iav}V{}V@V{\overline{R}av}V{}V\\ \mathbf{I}_{N}^{n}@<{}<{s}<\overline{\mathbf{R}}_{N}^{n}\\ \end{CD}

commutes.

This corollary in turn implies the following:

Corollary 2.4.

We have the following equality

Iavpr𝐈=pr𝐈Rav,\displaystyle Iav\circ pr_{\mathbf{I}}=pr_{\mathbf{I}}\circ Rav, (2.26)

namely, the diagram

𝐈Nnpr𝐈𝐑NnIavRav𝐈Nnpr𝐈𝐑Nn\displaystyle\begin{CD}\mathbf{I}_{N}^{n}@<{}<{pr_{\mathbf{I}}}<\mathbf{R}_{N}^{n}\\ @V{Iav}V{}V@V{Rav}V{}V\\ \mathbf{I}_{N}^{n}@<{}<{pr_{\mathbf{I}}}<\mathbf{R}_{N}^{n}\\ \end{CD}

commutes.

Proof.

Since pr𝐈=sπpr_{\mathbf{I}}=s\circ\pi, we can compute as follows:

Iavpr𝐈\displaystyle Iav\circ pr_{\mathbf{I}} =\displaystyle= Iavsπ\displaystyle Iav\circ s\circ\pi
=\displaystyle= sR¯avπ(by (2.23))\displaystyle s\circ\overline{R}av\circ\pi\hskip 14.22636pt(\mbox{by }(2.23))
=\displaystyle= sπRav(by (2.4))\displaystyle s\circ\pi\circ Rav\hskip 14.22636pt(\mbox{by }(2.4))
=\displaystyle= pr𝐈Rav(by (2.4)).\displaystyle pr_{\mathbf{I}}\circ Rav\hskip 14.22636pt(\mbox{by }(2.4)).

This completes the proof. ∎

2.5 Counterpart of RavRav on 𝐁Nn\mathbf{B}_{N}^{n}

We transport the discrete average on 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n} to the corresponding map on 𝐁Nn\mathbf{B}_{N}^{n}:

Definition 2.11.

The Boolean average Bav:𝐁Nn𝐁NnBav:\mathbf{B}_{N}^{n}\rightarrow\mathbf{B}_{N}^{n} is defined by

Bav=R¯toBR¯avBtoR¯,\displaystyle Bav=\overline{R}toB\circ\overline{R}av\circ Bto\overline{R}, (2.27)

which makes the following diagram commute:

𝐑¯NnR¯toB𝐁NnR¯avBav𝐑¯NnR¯toB𝐁Nn\displaystyle\begin{CD}\overline{\mathbf{R}}_{N}^{n}@>{\overline{R}toB}>{}>\mathbf{B}_{N}^{n}\\ @V{\overline{R}av}V{}V@V{Bav}V{}V\\ \overline{\mathbf{R}}_{N}^{n}@>{\overline{R}toB}>{}>\mathbf{B}_{N}^{n}\\ \end{CD}

Namely the equality

BavR¯toB=R¯toBR¯av\displaystyle Bav\circ\overline{R}toB=\overline{R}toB\circ\overline{R}av (2.28)

holds.

The following lemma relates BavBav with IavIav:

Lemma 2.1.
Bav=ItoBIavBtoI.\displaystyle Bav=ItoB\circ Iav\circ BtoI. (2.29)
Proof.

This is proved by inserting the defining equations (2.15), (2.16) into (2.25) as follows:

Bav\displaystyle Bav =\displaystyle= R¯toBR¯avBtoR¯\displaystyle\overline{R}toB\circ\overline{R}av\circ Bto\overline{R}
=\displaystyle= ItoBsR¯avπBtoI(by (2.15),(2.16))\displaystyle ItoB\circ s\circ\overline{R}av\circ\pi\circ BtoI\hskip 28.45274pt(\mbox{by }(2.15),(2.16))
=\displaystyle= ItoBsπIavBtoI(by (2.22))\displaystyle ItoB\circ s\circ\pi\circ Iav\circ BtoI\hskip 28.45274pt(\mbox{by }(2.22))
=\displaystyle= ItoBIavBtoI.(by Proposition 2.4)\displaystyle ItoB\circ Iav\circ BtoI.\hskip 56.9055pt(\mbox{by Proposition }2.4)

This finishes the proof. ∎

Since ItoBItoB is the inverse of BtoIBtoI, we have the following:

Proposition 2.9.

We have the equality

BavItoB=ItoBIav,\displaystyle Bav\circ ItoB=ItoB\circ Iav, (2.30)

namely, the diagram

𝐈NnItoB𝐁NnIavBav𝐈NnItoB𝐁Nn\displaystyle\begin{CD}\mathbf{I}_{N}^{n}@>{ItoB}>{}>\mathbf{B}_{N}^{n}\\ @V{Iav}V{}V@V{Bav}V{}V\\ \mathbf{I}_{N}^{n}@>{ItoB}>{}>\mathbf{B}_{N}^{n}\\ \end{CD}

commutes.

Thus we have accomplised the purpose of this section, namely our construction of three maps R¯av,Iav,Bav\overline{R}av,Iav,Bav, each of which corresponds to the discrete average map RavRav on 𝐑N\mathbf{R}_{N}.
The following example indicates how we can compute IavIav and BavBav:
Example 2.1. Let N=8N=8 and n=3n=3. For the vector 𝐯=(0,0,1,1,0,0,0,1)𝐁83\mathbf{v}=(0,0,1,1,0,0,0,1)\in\mathbf{B}_{8}^{3}, we can compute Bav(𝐯)Bav(\mathbf{v}) as follows:

Bav(𝐯)\displaystyle Bav(\mathbf{v}) =\displaystyle= (ItoBIavBtoI)(0,0,1,1,0,0,0,1)(by Lemma 2.1)\displaystyle(ItoB\circ Iav\circ BtoI)(0,0,1,1,0,0,0,1)\hskip 14.22636pt(\mbox{by Lemma 2.1)}
=\displaystyle= (ItoB(Iav(2,3,7))(by (2.12))\displaystyle(ItoB(Iav(2,3,7))\hskip 99.58464pt(\mbox{by (2.12)})
=\displaystyle= ItoB(rot(Rav(2,3,7)))(since (2,3,7) is improper)\displaystyle ItoB(rot(Rav(2,3,7)))\hskip 42.67912pt(\mbox{since }(2,3,7)\mbox{ is improper})
=\displaystyle= ItoB(rot(2,5,0))\displaystyle ItoB(rot(2,5,0))
=\displaystyle= ItoB(0,2,5)\displaystyle ItoB(0,2,5)
=\displaystyle= (1,0,1,0,0,1,0,0),\displaystyle(1,0,1,0,0,1,0,0),

the last quality coming from (2.11) and Definition 2.7.

3 Cyclicity of the components of the Boolean average

Our main theme in this paper is to investigate the properties of the Boolean average BavBav. As a map from 𝐅2N\mathbf{F}_{2}^{N} to itself, BavBav has NN components each of which is a Boolean function on 𝐅2N\mathbf{F}_{2}^{N}. In this section we show that only one of these NN Boolean functions determines the shapes of the other N1N-1 functions. For this purpose we need to check the functorialities of several maps.

3.1 Boolean counterpart of the translation map trtr on 𝐑Nn\mathbf{R}_{N}^{n}

In [2], we introduce a self map tr:𝐑Nn𝐑Nntr:\mathbf{R}_{N}^{n}\rightarrow\mathbf{R}_{N}^{n} when n2n\geq 2, which is defined by

tr(a0,,an1)=(a0+N1,,an1+N1)\displaystyle tr(a_{0},\cdots,a_{n-1})=(a_{0}+_{N}1,\cdots,a_{n-1}+_{N}1)

for any (a0,,an1)𝐑Nn(a_{0},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n}. When n=1n=1 or n=0n=0, we supply this with the following natural ones:

tr(a0)\displaystyle tr(a_{0}) =\displaystyle= (a0+N1),\displaystyle(a_{0}+_{N}1),
tr(ϕ)\displaystyle tr(\phi) =\displaystyle= ϕ.\displaystyle\phi.

Since this map trtr commutes with the map rotrot for any n0n\geq 0, It descends to a self map on 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n}, which we denote by R¯tr\overline{R}tr:

R¯tr:𝐑¯Nn𝐑¯Nn.\displaystyle\overline{R}tr:\overline{\mathbf{R}}_{N}^{n}\rightarrow\overline{\mathbf{R}}_{N}^{n}.

Therefore we have

R¯trπ\displaystyle\overline{R}tr\circ\pi =\displaystyle= πtr,\displaystyle\pi\circ tr, (3.1)

namely, we have the following commutative diagram:

𝐑Nnπ𝐑¯NntrR¯tr𝐑Nnπ𝐑¯Nn\displaystyle\begin{CD}\mathbf{R}_{N}^{n}@>{\pi}>{}>\overline{\mathbf{R}}_{N}^{n}\\ @V{tr}V{}V@V{\overline{R}tr}V{}V\\ \mathbf{R}_{N}^{n}@>{\pi}>{}>\overline{\mathbf{R}}_{N}^{n}\\ \end{CD}

Next we show that the map trtr commutes with RavRav:

Proposition 3.1.

The two self maps trtr and RavRav on 𝐑Nn\mathbf{R}_{N}^{n} commute with each other:

trRav=Ravtr,\displaystyle tr\circ Rav=Rav\circ tr,

namely, we have the following commutative diagram:

𝐑Nntr𝐑NnRavRav𝐑Nntr𝐑Nn\displaystyle\begin{CD}\mathbf{R}_{N}^{n}@>{tr}>{}>\mathbf{R}_{N}^{n}\\ @V{Rav}V{}V@V{Rav}V{}V\\ \mathbf{R}_{N}^{n}@>{tr}>{}>\mathbf{R}_{N}^{n}\\ \end{CD}
Proof.

It follows from the definition of the 𝐙N\mathbf{Z}_{N}-average that

av𝐙N(a+N1,b+N1)\displaystyle av_{\mathbf{Z}_{N}}(a+_{N}1,b+_{N}1) =\displaystyle= av𝐙N(a,b)+N1.\displaystyle av_{\mathbf{Z}_{N}}(a,b)+_{N}1. (3.2)

Therefore for any 𝐚=(a0,,an1)𝐑Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n}, we have

(Ravtr)(𝐚)\displaystyle(Rav\circ tr)(\mathbf{a}) =\displaystyle= Rav(tr(a0,,an1))\displaystyle Rav(tr(a_{0},\cdots,a_{n-1}))
=\displaystyle= Rav(a0+N1,,an1+N1)\displaystyle Rav(a_{0}+_{N}1,\cdots,a_{n-1}+_{N}1)
=\displaystyle= (av𝐙N(a0+N1,a1+N1),,av𝐙N(an1+N1,a0+N1))\displaystyle(av_{\mathbf{Z}_{N}}(a_{0}+_{N}1,a_{1}+_{N}1),\cdots,av_{\mathbf{Z}_{N}}(a_{n-1}+_{N}1,a_{0}+_{N}1))
=\displaystyle= (av𝐙N(a0,a1)+N1,,av𝐙N(an1,a0)+N1)(by (3.2))\displaystyle(av_{\mathbf{Z}_{N}}(a_{0},a_{1})+_{N}1,\cdots,av_{\mathbf{Z}_{N}}(a_{n-1},a_{0})+_{N}1)\hskip 14.22636pt(\mbox{by }(3.2))
=\displaystyle= (trRav)(𝐚).\displaystyle(tr\circ Rav)(\mathbf{a}).

This completes the proof. ∎

Since we have seen that both maps trtr and RavRav on 𝐑Nn\mathbf{R}_{N}^{n} descend to the corresponding maps on 𝐑¯Nn\overline{\mathbf{R}}_{N}^{n}, this proposition implies the following:

Corollary 3.1.

For the two self maps R¯av\overline{R}av and R¯tr\overline{R}tr, the equality

R¯avR¯tr=R¯trR¯av,\displaystyle\overline{R}av\circ\overline{R}tr=\overline{R}tr\circ\overline{R}av, (3.3)

holds, namely, the following diagram commutes:

𝐑¯NnR¯tr𝐑¯NnR¯avR¯av𝐑¯NnR¯tr𝐑¯Nn\displaystyle\begin{CD}\overline{\mathbf{R}}_{N}^{n}@>{\overline{R}tr}>{}>\overline{\mathbf{R}}_{N}^{n}\\ @V{\overline{R}av}V{}V@V{\overline{R}av}V{}V\\ \overline{\mathbf{R}}_{N}^{n}@>{\overline{R}tr}>{}>\overline{\mathbf{R}}_{N}^{n}\\ \end{CD}

Next we define a Boolean counterpart BtrBtr of trtr:

Definition 3.1.

Let Btr:𝐁N𝐁NBtr:\mathbf{B}_{N}\rightarrow\mathbf{B}_{N} denote the map defined by

Btr(v0,v1,v2,,vN1)=(vN1,v0,v1,,vN2)\displaystyle Btr(v_{0},v_{1},v_{2},\cdots,v_{N-1})=(v_{N-1},v_{0},v_{1},\cdots,v_{N-2})

for any 𝐯=(v0,v1,v2,,vN1)𝐁N\mathbf{v}=(v_{0},v_{1},v_{2},\cdots,v_{N-1})\in\mathbf{B}_{N}.

The following proposition shows the compatibility of BtrBtr and trtr:

Proposition 3.2.

We have the equality of maps:

BtrRtoB=RtoBtr,\displaystyle Btr\circ RtoB=RtoB\circ tr, (3.4)

namely, the following diagram commutes:

𝐑NnRtoB𝐁NntrBtr𝐑NnRtoB𝐁Nn\displaystyle\begin{CD}\mathbf{R}_{N}^{n}@>{RtoB}>{}>\mathbf{B}_{N}^{n}\\ @V{tr}V{}V@V{Btr}V{}V\\ \mathbf{R}_{N}^{n}@>{RtoB}>{}>\mathbf{B}_{N}^{n}\\ \end{CD}
Proof.

For any 𝐚=(a0,a1,,an1)𝐑Nn\mathbf{a}=(a_{0},a_{1},\cdots,a_{n-1})\in\mathbf{R}_{N}^{n}, Definition 2.7 implies that

RtoB(𝐚)=𝐯=(v0,v1,,vN1)𝐁Nn,\displaystyle RtoB(\mathbf{a})=\mathbf{v}=(v_{0},v_{1},\cdots,v_{N-1})\in\mathbf{B}_{N}^{n},

where

vi={1, if i𝐚,0, if i𝐚,\displaystyle v_{i}=\left\{\begin{array}[]{ll}1,&\mbox{ if $i\in\mathbf{a}$},\\ 0,&\mbox{ if $i\not\in\mathbf{a}$},\\ \end{array}\right.

Therefore the left hand side of (3.4) maps 𝐚\mathbf{a} to

Btr(RtoB(𝐚))\displaystyle Btr(RtoB(\mathbf{a})) =\displaystyle= Btr(v0,v1,,vN1)\displaystyle Btr(v_{0},v_{1},\cdots,v_{N-1})
=\displaystyle= (vN1,v0,,vN2).\displaystyle(v_{N-1},v_{0},\cdots,v_{N-2}).

Let wi=viN1w_{i}=v_{i-_{N}1} for any i𝐙Ni\in\mathbf{Z}_{N}. Then we have

𝐰=(w0,w1,,wN1)=(vN1,v0,,vN2),\displaystyle\mathbf{w}=(w_{0},w_{1},\cdots,w_{N-1})=(v_{N-1},v_{0},\cdots,v_{N-2}),

namely we have

𝐰=Btr(RtoB(𝐚)).\displaystyle\mathbf{w}=Btr(RtoB(\mathbf{a})).

Notice that we have the following series of equivalences:

wi=1\displaystyle w_{i}=1 \displaystyle\Leftrightarrow viN1=1\displaystyle v_{i-_{N}1}=1
\displaystyle\Leftrightarrow iN1𝐚\displaystyle i-_{N}1\in\mathbf{a}
\displaystyle\Leftrightarrow itr(𝐚).\displaystyle i\in tr(\mathbf{a}).

It follows that

𝐰=RtoB(tr(𝐚)).\displaystyle\mathbf{w}=RtoB(tr(\mathbf{a})).

Hence we have

BtrRtoB=RtoBtr,\displaystyle Btr\circ RtoB=RtoB\circ tr,

This completes the proof. ∎

As an immediate consequence, we have the following:

Corollary 3.2.

The equality of composite maps

BtrR¯toB=R¯toBR¯tr\displaystyle Btr\circ\overline{R}toB=\overline{R}toB\circ\overline{R}tr (3.6)

holds, namely, the following diagram commutes:

𝐑¯NnR¯toB𝐁NnR¯trBtr𝐑¯NnR¯toB𝐁Nn\displaystyle\begin{CD}\overline{\mathbf{R}}_{N}^{n}@>{\overline{R}toB}>{}>\mathbf{B}_{N}^{n}\\ @V{\overline{R}tr}V{}V@V{Btr}V{}V\\ \overline{\mathbf{R}}_{N}^{n}@>{\overline{R}toB}>{}>\mathbf{B}_{N}^{n}\\ \end{CD}
Proof.

For any x𝐑¯Nnx\in\overline{\mathbf{R}}_{N}^{n}, let 𝐚\mathbf{a} be an arbitrary representative of the equivalence class xx. Then we have

(R¯toBR¯tr)(x)\displaystyle(\overline{R}toB\circ\overline{R}tr)(x) =\displaystyle= (R¯toBR¯tr)(π(𝐚))\displaystyle(\overline{R}toB\circ\overline{R}tr)(\pi(\mathbf{a}))
=\displaystyle= (R¯toBR¯trπ)(𝐚)\displaystyle(\overline{R}toB\circ\overline{R}tr\circ\pi)(\mathbf{a})
=\displaystyle= (R¯toBπtr)(𝐚)(by (3.1))\displaystyle(\overline{R}toB\circ\pi\circ tr)(\mathbf{a})\hskip 28.45274pt(\mbox{by }(3.1))
=\displaystyle= (RtoBtr)(𝐚)(by Proposition 2.6)\displaystyle(RtoB\circ tr)(\mathbf{a})\hskip 45.5244pt(\mbox{by Proposition }2.6)
=\displaystyle= (BtrRtoB)(𝐚)(by (3.4))\displaystyle(Btr\circ RtoB)(\mathbf{a})\hskip 36.98857pt(\mbox{by }(3.4))
=\displaystyle= (BtrR¯toBπ)(𝐚)(by Propotision 2.6)\displaystyle(Btr\circ\overline{R}toB\circ\pi)(\mathbf{a})\hskip 22.76219pt(\mbox{by Propotision }2.6)
=\displaystyle= (BtrR¯toB)(x).\displaystyle(Btr\circ\overline{R}toB)(x).

This completes the proof of Corollary 3.2. ∎

3.2 Cyclicity of the Boolean average

Now we can prove the main result of this section, which asserts that only one of the components of BavBav determines the others completely:

Theorem 3.1.

(1)(1) The two self maps Bav,BtrBav,Btr of the space 𝐁Nn\mathbf{B}_{N}^{n} commute with each other:

BtrBav=BavBtr,\displaystyle Btr\circ Bav=Bav\circ Btr, (3.7)

namely, the following diagram commutes:

𝐁NnBtr𝐁NnBavBav𝐁NnBtr𝐁Nn\displaystyle\begin{CD}\mathbf{B}_{N}^{n}@>{Btr}>{}>\mathbf{B}_{N}^{n}\\ @V{Bav}V{}V@V{Bav}V{}V\\ \mathbf{B}_{N}^{n}@>{Btr}>{}>\mathbf{B}_{N}^{n}\\ \end{CD}

(2)(2) For any i𝐙Ni\in\mathbf{Z}_{N}, we denote by BavNiBav_{N}^{i} the composite priBav:𝐁Nn𝐅2pr_{i}\circ Bav:\mathbf{B}_{N}^{n}\rightarrow\mathbf{F}_{2}, where pri:𝐅2N𝐅2pr_{i}:\mathbf{F}_{2}^{N}\rightarrow\mathbf{F}_{2} is the projection onto the ii-th factor. Then we have

Btr(BavNi)=BavNiN1.\displaystyle Btr^{*}(Bav_{N}^{i})=Bav_{N}^{i-_{N}1}. (3.8)

Here the left hand side means the pull-back of the function BavNiBav_{N}^{i} through the self map BtrBtr on 𝐁Nn\mathbf{B}_{N}^{n}, namely the composite BavNiBtrBav_{N}^{i}\circ Btr.

Proof.

(1) Composing the left hand side of (3.6) with the bijection R¯toB\overline{R}toB, we compute as follows:

BtrBavR¯toB\displaystyle Btr\circ Bav\circ\overline{R}toB =\displaystyle= BtrR¯toBR¯av(by (2.26))\displaystyle Btr\circ\overline{R}toB\circ\overline{R}av\hskip 28.45274pt(\mbox{by }(2.26))
=\displaystyle= R¯toBR¯trR¯av(by (3.5))\displaystyle\overline{R}toB\circ\overline{R}tr\circ\overline{R}av\hskip 28.45274pt(\mbox{by }(3.5))
=\displaystyle= R¯toBR¯avR¯tr(by (3.3))\displaystyle\overline{R}toB\circ\overline{R}av\circ\overline{R}tr\hskip 28.45274pt(\mbox{by }(3.3))
=\displaystyle= BavR¯toBR¯tr(by (2.26))\displaystyle Bav\circ\overline{R}toB\circ\overline{R}tr\hskip 28.45274pt(\mbox{by }(2.26))
=\displaystyle= BavBtrR¯toB(by (3.5))\displaystyle Bav\circ Btr\circ\overline{R}toB\hskip 28.45274pt(\mbox{by }(3.5))

This shows that the equality (3.6) holds.
(2) Note that we have

priBtr=priN1,\displaystyle pr_{i}\circ Btr=pr_{i-_{N}1}, (3.9)

since for any 𝐯=(v0,v1,,vN1)𝐁Nn\mathbf{v}=(v_{0},v_{1},\cdots,v_{N-1})\in\mathbf{B}_{N}^{n}, we have

(priBtr)(𝐯)\displaystyle(pr_{i}\circ Btr)(\mathbf{v}) =\displaystyle= pri(vN1,v0,v1,,vN2)\displaystyle pr_{i}(v_{N-1},v_{0},v_{1},\cdots,v_{N-2})
=\displaystyle= viN1\displaystyle v_{i-_{N}1}
=\displaystyle= priN1(𝐯).\displaystyle pr_{i-_{N}1}(\mathbf{v}).

Hence we can compute as follows:

Btr(BavNi)\displaystyle Btr^{*}(Bav_{N}^{i}) =\displaystyle= BavNiBtr\displaystyle Bav_{N}^{i}\circ Btr
=\displaystyle= priBavBtr\displaystyle pr_{i}\circ Bav\circ Btr
=\displaystyle= priBtrBav( by (3.6))\displaystyle pr_{i}\circ Btr\circ Bav\hskip 34.1433pt(\Leftarrow\mbox{ by }(3.6))
=\displaystyle= priN1Bav( by (3.8))\displaystyle pr_{i-_{N}1}\circ Bav\hskip 42.67912pt(\Leftarrow\mbox{ by }(3.8))
=\displaystyle= BavNiN1.\displaystyle Bav_{N}^{i-_{N}1}.

This completes the proof. ∎

Remark 3.1.

In concrete terms, the equality (3.7) means that every BavNiBav_{N}^{i} (i𝐙N)(i\in\mathbf{Z}_{N}) can be obtained by a cyclic coordinate change from BavN0Bav_{N}^{0}.

4 Polynomial expression of the Boolean average

In this section we investigate how the Boolean average BavBav is expressed as a polynomial map on 𝐁N\mathbf{B}_{N}. First we recall some standard method which translates any propositional functions into polynomial expressions.

4.1 Algebraic standard form of propositional functions

In this subsection we recall some standard facts about propositional functions. See [3,4] for details.
Let the truth-values 0 and 11 stand for ”false” and ”true”, respectively. Then any proposition can be regarded as an 𝐅2\mathbf{F}_{2}-valued function on 𝐅2N\mathbf{F}_{2}^{N} for an appropriate NN. For example, the logical connective ”and” corresponds to the multiplication operator ”×\times”, as is observed from the truth table of ”and”. On the other hand, the addition operator ”++” corresponds to the connective ”xor (exclusive or)”. Since any proposition can be expressed as a compound of these two connectives, it corresponds to a Boolean function of a certain number of variables. In view of the fact that we express a polynomial in several variables usually as a sum of monomials, the disjunctive normal form of logical formula plays a fundamental role for our purpose. We recall its construction by a typical example.
Example 4.1. The disjunctive normal form of P=(xy)(yz)P=(x\vee y)\wedge(y\vee z). The truth table of PP is given by

numberxyzxyyzP(1)000000(2)001010(3)010111(4)011111(5)100100(6)101111(7)110111(8)111111\displaystyle\begin{array}[]{lccccccccc}\mbox{number}&\vline&x&y&z&\vline&x\vee y&y\vee z&\vline&P\\ \hline\cr(1)&\vline&0&0&0&\vline&0&0&\vline&0\\ (2)&\vline&0&0&1&\vline&0&1&\vline&0\\ (3)&\vline&0&1&0&\vline&1&1&\vline&1\\ (4)&\vline&0&1&1&\vline&1&1&\vline&1\\ (5)&\vline&1&0&0&\vline&1&0&\vline&0\\ (6)&\vline&1&0&1&\vline&1&1&\vline&1\\ (7)&\vline&1&1&0&\vline&1&1&\vline&1\\ (8)&\vline&1&1&1&\vline&1&1&\vline&1\\ \end{array}

Table 4.1. Truth table for P=(xy)(yz)P=(x\vee y)\wedge(y\vee z)

There are five rows, named (3), (4), (6), (7), (8), for which the value of PP equals 1. We associate a term with each of these five rows. For example, the third row gives PP the value 1. We make a constituent of the corresponding term by the following rule. Here ”ww” stands for any one of the variables x,y,zx,y,z:

If the value of ww is 1, then use ”ww”,
otherwise use ”w¯\overline{w}” as its constituent.

Therefore the third row provides us with the term x¯yz¯\overline{x}y\overline{z}, since (x,y,z)=(0,1,0)(x,y,z)=(0,1,0). Connecting the term x¯yz¯\overline{x}y\overline{z} with the other four terms obtained from the remaining four rows, we obtain the following disjunctive normal form of the proposition PP:

P=x¯yz¯x¯yzxy¯zxyz¯xyz.\displaystyle P=\overline{x}y\overline{z}\vee\overline{x}yz\vee x\overline{y}z\vee xy\overline{z}\vee xyz.

Note that we can replace all ”\vee” by ”¯\underline{\vee}”, since each term on the right hand side contradicts to each other. Therefore we can rewrite this as

P=x¯yz¯¯x¯yz¯xy¯z¯xyz¯¯xyz.\displaystyle P=\overline{x}y\overline{z}\hskip 2.84526pt\underline{\vee}\hskip 2.84526pt\overline{x}yz\hskip 2.84526pt\underline{\vee}\hskip 2.84526ptx\overline{y}z\hskip 2.84526pt\underline{\vee}\hskip 2.84526ptxy\overline{z}\hskip 2.84526pt\underline{\vee}\hskip 2.84526ptxyz.

By letting w¯=1+w\overline{w}=1+w for each variable w{x,y,z}w\in\{x,y,z\}, and by reducing the result modulo 2, we arrive at the following expression BPB_{P} of PP as a Boolean polynomial:

BP\displaystyle B_{P} =\displaystyle= (1+x)y(1+z)+(1+x)yz+x(1+y)z+xy(1+z)+xyz\displaystyle(1+x)y(1+z)+(1+x)yz+x(1+y)z+xy(1+z)+xyz
=\displaystyle= 5xyz+2xy+2yz+xz+y\displaystyle 5xyz+2xy+2yz+xz+y
=\displaystyle= xyz+xz+y.\displaystyle xyz+xz+y.

One can check that the rightmost side xyz+xz+yxyz+xz+y coincides with PP as a function on 𝐅23\mathbf{F}_{2}^{3}.

Remark 4.1.

In general, any Boolean polynomial f(x1,,xn)f(x_{1},\cdots,x_{n}) should be regarded as an object in 𝐅2[x1,,xn]/I\mathbf{F}_{2}[x_{1},\cdots,x_{n}]/I, where II denotes the ideal

I=(x12x1,,xn2xn).\displaystyle I=(x_{1}^{2}-x_{1},\cdots,x_{n}^{2}-x_{n}).

In the example above, however, one does not need to reduce BPB_{P} modulo the ideal II, since it has no squared variables.

4.2 Rhythm polynomials

In this subsection we investigate the problem how the self map BavBav on 𝐁Nn\mathbf{B}_{N}^{n} can be expressed as a polynomial map. Since 𝐁N=n=0N𝐁Nn\mathbf{B}_{N}=\bigcup_{n=0}^{N}\mathbf{B}_{N}^{n}, we can extend the domain of the definition of the map BavBav to the whole 𝐁N\mathbf{B}_{N} naturally. By abuse of language we also denote the extended map by BavBav. Furthermore the composite

BavNi=priBav:𝐅2N𝐅2\displaystyle Bav_{N}^{i}=pr_{i}\circ Bav:\mathbf{F}_{2}^{N}\rightarrow\mathbf{F}_{2}

is called the ii-th rhythm polynomial modulo NN. By Theorem 3.1, (2), every rhythm polynomial is obtained from the 0-th rhythm polynomial BavN0Bav_{N}^{0} through a cyclic change of variables viv_{i}, i𝐙Ni\in\mathbf{Z}_{N}. Therefore we can focus our attention exclusively on BavN0Bav_{N}^{0}.
First we examine the problem for the case N=3N=3. The following table displays the Boolean averages of vectors in 𝐁3\mathbf{B}_{3}. Recall that Bav=ItoBIavBtoIBav=ItoB\circ Iav\circ BtoI by (2.27):

𝐯BtoI(𝐯)Iav(BtoI(𝐯))Bav(𝐯)(0,0,0)ϕϕ(0,0,0)(0,0,1)(2)(2)(0,0,1)(0,1,0)(1)(1)(0,1,0)(0,1,1)(1,2)(0,1)(1,1,0)(1,0,0)(0)(0)(1,0,0)(1,0,1)(0,2)(1,2)(0,1,1)(1,1,0)(0,1)(0,2)(1,0,1)(1,1,1)(0,1,2)(0,1,2)(1,1,1)\displaystyle\begin{array}[]{ccccccc}\mathbf{v}&\vline&BtoI(\mathbf{v})&\vline&Iav(BtoI(\mathbf{v}))&\vline&Bav(\mathbf{v})\\ \hline\cr(0,0,0)&\vline&\phi&\vline&\phi&\vline&(0,0,0)\\ (0,0,1)&\vline&(2)&\vline&(2)&\vline&(0,0,1)\\ (0,1,0)&\vline&(1)&\vline&(1)&\vline&(0,1,0)\\ (0,1,1)&\vline&(1,2)&\vline&(0,1)&\vline&(1,1,0)\\ (1,0,0)&\vline&(0)&\vline&(0)&\vline&(1,0,0)\\ (1,0,1)&\vline&(0,2)&\vline&(1,2)&\vline&(0,1,1)\\ (1,1,0)&\vline&(0,1)&\vline&(0,2)&\vline&(1,0,1)\\ (1,1,1)&\vline&(0,1,2)&\vline&(0,1,2)&\vline&(1,1,1)\\ \end{array}

Table 4.2. Boolean averages of three-dimensional vectors

Employing the algorithm explained in the previous subsection, we have

Bav30(v0,v1,v2)\displaystyle Bav_{3}^{0}(v_{0},v_{1},v_{2}) =\displaystyle= v0¯v1v2+v0v1¯v2¯+v0v1v2¯+v0v1v2\displaystyle\overline{v_{0}}v_{1}v_{2}+v_{0}\overline{v_{1}}\overline{v_{2}}+v_{0}v_{1}\overline{v_{2}}+v_{0}v_{1}v_{2}
=\displaystyle= v0+v0v2+v1v2,\displaystyle v_{0}+v_{0}v_{2}+v_{1}v_{2},
Bav31(v0,v1,v2)\displaystyle Bav_{3}^{1}(v_{0},v_{1},v_{2}) =\displaystyle= v0¯v1v2¯+v0¯v1v2+v0v1¯v2+v0v1v2\displaystyle\overline{v_{0}}v_{1}\overline{v_{2}}+\overline{v_{0}}v_{1}v_{2}+v_{0}\overline{v_{1}}v_{2}+v_{0}v_{1}v_{2}
=\displaystyle= v1+v0v1+v0v2,\displaystyle v_{1}+v_{0}v_{1}+v_{0}v_{2},
Bav32(v0,v1,v2)\displaystyle Bav_{3}^{2}(v_{0},v_{1},v_{2}) =\displaystyle= v0¯v1¯v2+v0v1¯v2+v0v1v2¯+v0v1v2\displaystyle\overline{v_{0}}\overline{v_{1}}v_{2}+v_{0}\overline{v_{1}}v_{2}+v_{0}v_{1}\overline{v_{2}}+v_{0}v_{1}v_{2}
=\displaystyle= v2+v0v1+v1v2.\displaystyle v_{2}+v_{0}v_{1}+v_{1}v_{2}.

This computation confirms the cyclicity stated in Theorem 3.1, (2).
In a similar way we can compute the rhythm polynomial BavN0Bav_{N}^{0} for small NN. We list the results as follows:

NBavN03v0+v0v2+v1v24v0+v0v2+v0v3+v1v3+v2v3+v0v1v2+v1v2v35v0+v0v2+v0v3+v0v4+v1v4+v2v3+v2v4+v0v1v2+v0v1v3+v0v3v4+v1v2v3+v1v2v4+v2v3v4+v0v1v3v4+v1v2v3v4\displaystyle\begin{array}[]{cl}N&Bav_{N}^{0}\\ \hline\cr 3&v_{0}+v_{0}v_{2}+v_{1}v_{2}\\ 4&v_{0}+v_{0}v_{2}+v_{0}v_{3}+v_{1}v_{3}+v_{2}v_{3}+v_{0}v_{1}v_{2}+v_{1}v_{2}v_{3}\\ 5&v_{0}+v_{0}v_{2}+v_{0}v_{3}+v_{0}v_{4}+v_{1}v_{4}+v_{2}v_{3}+v_{2}v_{4}+v_{0}v_{1}v_{2}+v_{0}v_{1}v_{3}\\ &+v_{0}v_{3}v_{4}+v_{1}v_{2}v_{3}+v_{1}v_{2}v_{4}+v_{2}v_{3}v_{4}+v_{0}v_{1}v_{3}v_{4}+v_{1}v_{2}v_{3}v_{4}\\ \end{array}

Table 4.3. BavN0Bav_{N}^{0} for N=3,4,5N=3,4,5 in variables viv_{i}

We observe here that for each N{3,4,5}N\in\{3,4,5\} the number of terms in BavN0Bav_{N}^{0} is equal to 2N112^{N-1}-1. If we change, however, the variables viv_{i} to wi+1w_{i}+1 for i[0,N1]i\in[0,N-1], the number of terms diminishes to 2(N1)2(N-1), as is observed in the following list:

NBavN031+w1+w0w2+w1w241+w1+w0w1+w0w3+w0w1w2+w1w2w351+w1+w0w1+w0w4+w0w1w2+w0w1w4+w0w1w3w4+w1w2w3w461+w1+w0w1+w0w5+w0w1w2+w0w1w5+w0w1w2w5+w0w1w4w5+w0w1w2w3w5+w1w2w3w4w5\displaystyle\begin{array}[]{cl}N&Bav_{N}^{0}\\ \hline\cr 3&1+w_{1}+w_{0}w_{2}+w_{1}w_{2}\\ 4&1+w_{1}+w_{0}w_{1}+w_{0}w_{3}+w_{0}w_{1}w_{2}+w_{1}w_{2}w_{3}\\ 5&1+w_{1}+w_{0}w_{1}+w_{0}w_{4}+w_{0}w_{1}w_{2}+w_{0}w_{1}w_{4}\\ &+w_{0}w_{1}w_{3}w_{4}+w_{1}w_{2}w_{3}w_{4}\\ 6&1+w_{1}+w_{0}w_{1}+w_{0}w_{5}+w_{0}w_{1}w_{2}+w_{0}w_{1}w_{5}\\ &+w_{0}w_{1}w_{2}w_{5}+w_{0}w_{1}w_{4}w_{5}+w_{0}w_{1}w_{2}w_{3}w_{5}+w_{1}w_{2}w_{3}w_{4}w_{5}\\ \end{array}

Table 4.4. BavN0Bav_{N}^{0} for N=3,4,5,6N=3,4,5,6 in variables wiw_{i}

At this point, we can imagine neither the general shape of BavN0Bav_{N}^{0}, nor a possible relation between BavN0Bav_{N}^{0} and BavN+10Bav_{N+1}^{0}. We will show that, if we change the index set 𝐙N\mathbf{Z}_{N} to 𝐙N,±\mathbf{Z}_{N,\pm}, the set of the least absolute remainders modulo NN, as a new index set, then we can detect the structure of BavN0Bav_{N}^{0} more easily. When NN is even, we choose N2\frac{N}{2} from the two possible representatives {N2,N2}\{-\frac{N}{2},\frac{N}{2}\} as the remainder, and hence the set 𝐙N,±\mathbf{Z}_{N,\pm} is specified as

𝐙N,±={{m,m+1,,1,0,1,,m1,m}, if N=2m+1,{m+1,,1,0,1,,m1,m}, if N=2m.\displaystyle\mathbf{Z}_{N,\pm}=\left\{\begin{array}[]{ll}\{-m,-m+1,\cdots,-1,0,1,\cdots,m-1,m\},&\mbox{ if $N=2m+1$},\\ \{-m+1,\cdots,-1,0,1,\cdots,m-1,m\},&\mbox{ if $N=2m$}.\\ \end{array}\right.

By employing the floor symbol, we can express the set 𝐙N,±\mathbf{Z}_{N,\pm} regardless of the parity of NN as follows:

𝐙N,±=[N12,N2].\displaystyle\mathbf{Z}_{N,\pm}=[-\left\lfloor\frac{N-1}{2}\right\rfloor,\left\lfloor\frac{N}{2}\right\rfloor]. (4.6)

For later use, we divide the set 𝐙N,±\mathbf{Z}_{N,\pm} into the following three parts:

𝐙N,±=𝐙N,{0}𝐙N,+,\displaystyle\mathbf{Z}_{N,\pm}=\mathbf{Z}_{N,-}\sqcup\{0\}\sqcup\mathbf{Z}_{N,+},

where

𝐙N,\displaystyle\mathbf{Z}_{N,-} =\displaystyle= {k𝐙N,±;k<0},\displaystyle\{k\in\mathbf{Z}_{N,\pm};k<0\},
𝐙N,+\displaystyle\mathbf{Z}_{N,+} =\displaystyle= {k𝐙N,±;k>0},\displaystyle\{k\in\mathbf{Z}_{N,\pm};k>0\},

and put

𝐙N,0\displaystyle\mathbf{Z}_{N,\leq 0} =\displaystyle= 𝐙N,{0}.\displaystyle\mathbf{Z}_{N,-}\cup\{0\}.

We construct a bridge between 𝐙N\mathbf{Z}_{N} and 𝐙N,±\mathbf{Z}_{N,\pm} employing a simple bijection defined as follows:

Definition 4.1.

Let rem±:𝐙𝐙N,±rem_{\pm}:\mathbf{Z}\rightarrow\mathbf{Z}_{N,\pm} denote the map whose value at k𝐙k\in\mathbf{Z} is the least absolute remainder of kk modulo NN. Let φN:𝐙N𝐙N,±\varphi_{N}:\mathbf{Z}_{N}\rightarrow\mathbf{Z}_{N,\pm} denote the composite rem±ιrem_{\pm}\circ\iota, where ι:𝐙N𝐙\iota:\mathbf{Z}_{N}\rightarrow\mathbf{Z} is the natural inclusion map, namely, φN\varphi_{N} is a bijection between 𝐙N\mathbf{Z}_{N} and 𝐙N,±\mathbf{Z}_{N,\pm} defined by

φN(k)={k, if 0kN2,kN, if N2<kN1.\displaystyle\varphi_{N}(k)=\left\{\begin{array}[]{ll}k,&\mbox{ if }0\leq k\leq\left\lfloor\frac{N}{2}\right\rfloor,\\ k-N,&\mbox{ if }\left\lfloor\frac{N}{2}\right\rfloor<k\leq N-1.\\ \end{array}\right.
Remark 4.2.

The map φN\varphi_{N} satisfies the condition

φN(a)a(modN)\displaystyle\varphi_{N}(a)\equiv a\pmod{N} (4.8)

for any a𝐙Na\in\mathbf{Z}_{N}. Therefore φN\varphi_{N} exchanges the two representatives in 𝐙N\mathbf{Z}_{N} and in 𝐙N,±\mathbf{Z}_{N,\pm} of a class in the quotient 𝐙/N𝐙\mathbf{Z}/N\mathbf{Z}.

We introduce new variables yjy_{j}, j𝐙N,±j\in\mathbf{Z}_{N,\pm}, by the rule

wi=yϕN(i)\displaystyle w_{i}=y_{\phi_{N}(i)}

for any i𝐙Ni\in\mathbf{Z}_{N}. Then Table 4.4 transforms to the following:

NBavN031+y1+y0y1+y1y141+y1+y0y1+y0y1+y0y1y2+y1y2y151+y1+y0y1+y0y1+y0y1y2+y0y1y1+y0y1y2y1+y1y2y2y161+y1+y0y1+y0y1+y0y1y2+y0y1y1+y0y1y2y1+y0y1y1y1+y0y1y2y3y1+y1y2y3y2y1\displaystyle\begin{array}[]{cl}N&Bav_{N}^{0}\\ \hline\cr 3&1+y_{1}+y_{0}y_{-1}+y_{1}y_{-1}\\ 4&1+y_{1}+y_{0}y_{1}+y_{0}y_{-1}+y_{0}y_{1}y_{2}+y_{1}y_{2}y_{-1}\\ 5&1+y_{1}+y_{0}y_{1}+y_{0}y_{-1}+y_{0}y_{1}y_{2}+y_{0}y_{1}y_{-1}\\ &+y_{0}y_{1}y_{-2}y_{-1}+y_{1}y_{2}y_{-2}y_{-1}\\ 6&1+y_{1}+y_{0}y_{1}+y_{0}y_{-1}+y_{0}y_{1}y_{2}+y_{0}y_{1}y_{-1}\\ &+y_{0}y_{1}y_{2}y_{-1}+y_{0}y_{1}y_{-1}y_{-1}\\ &+y_{0}y_{1}y_{2}y_{3}y_{-1}+y_{1}y_{2}y_{3}y_{-2}y_{-1}\\ \end{array}

Table 4.5. BavN0Bav_{N}^{0} for N=3,4,5,6N=3,4,5,6 in variables yiy_{i}, i𝐙N,±i\in\mathbf{Z}_{N,\pm}

This table suggests us the structure of the rhythm polynomial BavN0Bav_{N}^{0} more clearly than Table 4.4. We explore in the next section the true nature of these polynomials.

5 Determination of rhythm polynomials

In this section we determine the Boolean polynomial BavN0Bav_{N}^{0} completely for any NN, and derive a recurrence formula which connects BavN0Bav_{N}^{0} and BavN10Bav_{N-1}^{0}. Furthermore we prove that BavN0Bav_{N}^{0} is a balanced polynomial. (See Definition 5.10.)

5.1 Change of the set of indices

As is motivated in Subsection 4.2, we need to use two index sets 𝐙N\mathbf{Z}_{N} and 𝐙N,±\mathbf{Z}_{N,\pm} properly to understand the rhythm polynomials. Fortunately enough, even if we employ 𝐙N,±\mathbf{Z}_{N,\pm} as the index set, several corresponding objects can be defined in almost the same way.
By abuse of language, we denote also by φN\varphi_{N} the bijective map 𝐙Nn𝐙N,±n\mathbf{Z}_{N}^{n}\rightarrow\mathbf{Z}_{N,\pm}^{n} induced by φN\varphi_{N} between their self-products, and put

𝐑N,±n=φN(𝐑Nn).\displaystyle\mathbf{R}_{N,\pm}^{n}=\varphi_{N}(\mathbf{R}_{N}^{n}).
Definition 5.1.

Let 𝐈N,±n\mathbf{I}_{N,\pm}^{n} denote the subset of 𝐑N,±n\mathbf{R}_{N,\pm}^{n} which consists of the monotone increasing sequence:

𝐈N,±n={𝐚=(a0,,an1)𝐑N,±n;a0<a1<<an1}.\displaystyle\mathbf{I}_{N,\pm}^{n}=\{\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{R}_{N,\pm}^{n};a_{0}<a_{1}<\cdots<a_{n-1}\}.
Definition 5.2.

Let 𝐁N,±\mathbf{B}_{N,\pm} denote the set of nn-dimensional Boolean vectors with index set 𝐙N,±\mathbf{Z}_{N,\pm}, namely,

𝐁N,±={(v,,vg);vi𝐅2 for any i[,g]},\displaystyle\mathbf{B}_{N,\pm}=\{(v_{\ell},\cdots,v_{g});v_{i}\in\mathbf{F}_{2}\mbox{ for any }i\in[\ell,g]\},

where we introduce the two symbols ,g\ell,g by

=N12,g=N2,\displaystyle\ell=-\left\lfloor\frac{N-1}{2}\right\rfloor,\hskip 5.69054ptg=\left\lfloor\frac{N}{2}\right\rfloor,

which are the least and the greatest elements in 𝐙N,±\mathbf{Z}_{N,\pm}.

We use the same symbol ItoBItoB and BtoIBtoI as before to denote the pair of bijections between 𝐈N,±n\mathbf{I}_{N,\pm}^{n} and 𝐁N,±n\mathbf{B}_{N,\pm}^{n}, which are defined as follows:

Definition 5.3.

(1)(1) For any 𝐚𝐈N,±n\mathbf{a}\in\mathbf{I}_{N,\pm}^{n}, the Boolean vector 𝐯=(v,,vg)=ItoB(𝐚)𝐁N,±\mathbf{v}=(v_{\ell},\cdots,v_{g})=ItoB(\mathbf{a})\in\mathbf{B}_{N,\pm} is defined by

vi={1, if i𝐚,0, if i𝐚.\displaystyle v_{i}=\left\{\begin{array}[]{ll}1,&\mbox{ if }i\in\mathbf{a},\\ 0,&\mbox{ if }i\not\in\mathbf{a}.\\ \end{array}\right. (5.3)

(2)(2) For any 𝐯𝐁N,±n\mathbf{v}\in\mathbf{B}_{N,\pm}^{n}, the nn-tuple BtoI(𝐯)𝐈N,±nBtoI(\mathbf{v})\in\mathbf{I}_{N,\pm}^{n} is defined to be the unique arrangement of the set {i𝐙N,±;vi=1}\{i\in\mathbf{Z}_{N,\pm};v_{i}=1\} in increasing order.

5.2 Parental pair and ancestor of zero in 𝐙N\mathbf{Z}_{N}

It is important to keep in mind the following fact:

Lemma 5.1.

The 0-th rhythm polynomial BavN0Bav_{N}^{0} takes value 11 at 𝐯𝐁N\mathbf{v}\in\mathbf{B}_{N} if and only if

0Rav(BtoI(𝐯)).\displaystyle 0\in Rav(BtoI(\mathbf{v})).
Proof.

Let 𝐚=(a0,,an1)=BtoI(𝐯)𝐈Nn\mathbf{a}=(a_{0},\cdots,a_{n-1})=BtoI(\mathbf{v})\in\mathbf{I}_{N}^{n}. Since it is in 𝐈Nn\mathbf{I}_{N}^{n}, we have

Iav(𝐚)=Rav(𝐚).\displaystyle Iav(\mathbf{a})=Rav(\mathbf{a}).

Recall that the two maps IavIav and BavBav are compatible through the map ItoBItoB by Proposition 2.9. Hence we have the following series of equivalences:

0Rav(𝐚)\displaystyle 0\in Rav(\mathbf{a}) \displaystyle\Leftrightarrow 0Iav(𝐚)\displaystyle 0\in Iav(\mathbf{a})
\displaystyle\Leftrightarrow 0(BtoIBavItoB)(𝐚)(by (2.27))\displaystyle 0\in(BtoI\circ Bav\circ ItoB)(\mathbf{a})\hskip 28.45274pt(\mbox{by }(2.27))
\displaystyle\Leftrightarrow 0(BtoIBav)(𝐯)\displaystyle 0\in(BtoI\circ Bav)(\mathbf{v})
\displaystyle\Leftrightarrow 0supp(Bav(𝐯))(by (2.12))\displaystyle 0\in supp(Bav(\mathbf{v}))\hskip 71.13188pt(\mbox{by }(2.12))
\displaystyle\Leftrightarrow pr0(Bav(𝐯))=1\displaystyle pr_{0}(Bav(\mathbf{v}))=1
\displaystyle\Leftrightarrow BavN0(𝐯)=1.\displaystyle Bav_{N}^{0}(\mathbf{v})=1.

Thus we finish the proof. ∎

By this lemma, in order to find the form of BavN0Bav_{N}^{0}, it is essential to determine the set of pairs (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N} such that av𝐙N(a,b)=0av_{\mathbf{Z}_{N}}(a,b)=0. This leads us to the following:

Definition 5.4.

A pair (a,b)𝐙N×𝐙N(a,b)\in\mathbf{Z}_{N}\times\mathbf{Z}_{N} is said to be a parental pair of zero if av𝐙N(a,b)=0av_{\mathbf{Z}_{N}}(a,b)=0. A rhythm 𝐚𝐑N\mathbf{a}\in\mathbf{R}_{N} is called an ancestor of zero if 0Rav(𝐚)0\in Rav(\mathbf{a}).

The following proposition provides us with a necessary and sufficient condition for a pair to be a parental pair of zero. It is expressed as a congruence modulo NN. This will facilitate our transition from 𝐙N\mathbf{Z}_{N} to 𝐙N,±\mathbf{Z}_{N,\pm} in the next subsection.

Proposition 5.1.

Let a,ba,b be an arbitrary pair of elements in 𝐙N\mathbf{Z}_{N}.
(0)(0) When a=0a=0, the following conditions are equivalent:
(0.A)av𝐙N(a,b)=0(0.{\rm A})\hskip 2.84526ptav_{\mathbf{Z}_{N}}(a,b)=0,
(0.B)b=0,1(0.{\rm B})\hskip 2.84526ptb=0,1.
(1)(1) When a=1a=1, these exists no b𝐙Nb\in\mathbf{Z}_{N} such that av𝐙N(a,b)=0av_{\mathbf{Z}_{N}}(a,b)=0.
(2)(2) When a2a\geq 2, the following conditions are equivalent:
(2.A)av𝐙N(a,b)=0(2.{\rm A})\hskip 2.84526ptav_{\mathbf{Z}_{N}}(a,b)=0,
(2.B)a>b and a+b0,1(modN)(2.{\rm B})\hskip 2.84526pta>b\mbox{ and }a+b\equiv 0,1\pmod{N}.

Remark 5.1.

When n2n\geq 2, an element in 𝐑Nn\mathbf{R}_{N}^{n} consists of nn distinct elements of 𝐙N\mathbf{Z}_{N} by the definition. Hence the first alternative in (0.B)(0.{\rm B}), namely the case when (a,b)=(0,0)(a,b)=(0,0) occurs only if n=1n=1. In that case, the discrete average Rav((0))Rav((0)) of the one-element rhythm (0)(0) is (0)(0).

Proof.

(0) When a=0a=0, we have

av𝐙N(0,b)=b2,\displaystyle av_{\mathbf{Z}_{N}}(0,b)=\left\lfloor\frac{b}{2}\right\rfloor,

and hence it is equal to 0 if and only if b=0,1b=0,1. Therefore (0.A) and (0.B) are equivalent.
(1) When a=1a=1, if av𝐙N(1,b)=0av_{\mathbf{Z}_{N}}(1,b)=0, then 0[1,b]N0\in[1,b]_{N} by Proposition 1.2, (2). This occurs only if b=0b=0. When NN is odd and N=2m+1N=2m+1, we have

av𝐙N(1,0)\displaystyle av_{\mathbf{Z}_{N}}(1,0) =\displaystyle= 1+N0N12\displaystyle 1+_{N}\left\lfloor\frac{0-_{N}1}{2}\right\rfloor
=\displaystyle= 1+NN12\displaystyle 1+_{N}\left\lfloor\frac{N-1}{2}\right\rfloor
=\displaystyle= 1+Nm.\displaystyle 1+_{N}m.

Note that 1+Nm=1+m<1+2m=N1+_{N}m=1+m<1+2m=N, and hence 1+Nm1+_{N}m cannot be equal to 0. When NN is even and N=2mN=2m, it follows by a similar computation that av𝐙N(1,0)=1+N(m1)av_{\mathbf{Z}_{N}}(1,0)=1+_{N}(m-1), which cannot be zero. This completes the proof of the assertion (1).
(2) (2.A)\Rightarrow(2.B): Since av𝐙N(a,b)[a,b]Nav_{\mathbf{Z}_{N}}(a,b)\in[a,b]_{N} by Proposition 1.2, (2), the condition av𝐙N(a,b)=0av_{\mathbf{Z}_{N}}(a,b)=0 implies that 0[a,b]0\in[a,b], and hence a>ba>b. Furthermore the equality av𝐙N(a,b)=a+NbNa2=0av_{\mathbf{Z}_{N}}(a,b)=a+_{N}\left\lfloor\frac{b-_{N}a}{2}\right\rfloor=0 holds if and only if

bNa2=Na,\displaystyle\left\lfloor\frac{b-_{N}a}{2}\right\rfloor=N-a, (5.4)

since a0a\neq 0. Note that

2x2={x, if x0(mod2),x1, if x1(mod2),\displaystyle 2\left\lfloor\frac{x}{2}\right\rfloor=\left\{\begin{array}[]{ll}x,&\mbox{ if }x\equiv 0\pmod{2},\\ x-1,&\mbox{ if }x\equiv 1\pmod{2},\\ \end{array}\right.

for any x𝐙x\in\mathbf{Z}. When bNa0(mod2)b-_{N}a\equiv 0\pmod{2}, doubling the both sides of (5.2), we have

bNa=2N2a,\displaystyle b-_{N}a=2N-2a,

which gives us the congruence

a+b0(modN).\displaystyle a+b\equiv 0\pmod{N}.

When bNa1(mod2)b-_{N}a\equiv 1\pmod{2}, the equality (5.2) implies similarly that

bNa1=2N2a,\displaystyle b-_{N}a-1=2N-2a,

which gives us the congruence

a+b1(modN).\displaystyle a+b\equiv 1\pmod{N}.

Thus (2.A) implies (2.B).
(2.B)\Rightarrow(2.A): In case a+b0(modN)a+b\equiv 0\pmod{N}, we have b=Nab=N-a since a>0a>0. Furthermore, since a>ba>b, we have a>Naa>N-a, therefore 0>N2a>N0>N-2a>-N. It follows that

bNa\displaystyle b-_{N}a =\displaystyle= (Na)Na\displaystyle(N-a)-_{N}a
=\displaystyle= (Na)aN\displaystyle\langle(N-a)-a\rangle_{N}
=\displaystyle= N2aN\displaystyle\langle N-2a\rangle_{N}
=\displaystyle= 2N2a.\displaystyle 2N-2a.

Hence we have

av𝐙N(a,b)\displaystyle av_{\mathbf{Z}_{N}}(a,b) =\displaystyle= a+NbNa2\displaystyle a+_{N}\left\lfloor\frac{b-_{N}a}{2}\right\rfloor
=\displaystyle= a+N2N2a2\displaystyle a+_{N}\left\lfloor\frac{2N-2a}{2}\right\rfloor
=\displaystyle= a+N(Na)\displaystyle a+_{N}(N-a)
=\displaystyle= 0.\displaystyle 0.

In case a+b1(modN)a+b\equiv 1\pmod{N}, we have b=N+1ab=N+1-a, since we have assumed that a2a\geq 2. Since a>ba>b, we have a>N+1aa>N+1-a, therefore 0>N+12a>N0>N+1-2a>-N. It follows that

bNa\displaystyle b-_{N}a =\displaystyle= (N+1a)Na\displaystyle(N+1-a)-_{N}a
=\displaystyle= (N+1a)aN\displaystyle\langle(N+1-a)-a\rangle_{N}
=\displaystyle= N+12aN\displaystyle\langle N+1-2a\rangle_{N}
=\displaystyle= 2N+12a.\displaystyle 2N+1-2a.

Hence we have

av𝐙N(a,b)\displaystyle av_{\mathbf{Z}_{N}}(a,b) =\displaystyle= a+NbNa2\displaystyle a+_{N}\left\lfloor\frac{b-_{N}a}{2}\right\rfloor
=\displaystyle= a+N2N+12a2\displaystyle a+_{N}\left\lfloor\frac{2N+1-2a}{2}\right\rfloor
=\displaystyle= a+N(Na)\displaystyle a+_{N}(N-a)
=\displaystyle= 0.\displaystyle 0.

This completes the proof. ∎

Corollary 5.1.

When a[1,N/2]a\in[1,N/2], there exists no b𝐙Nb\in\mathbf{Z}_{N} such that av𝐙N(a,b)av_{\mathbf{Z}_{N}}(a,b)
=0=0.

Proof.

By Proposition 5.1, (1), we may assume that 2aN/22\leq a\leq N/2. Suppose that av𝐙N(a,b)=0av_{\mathbf{Z}_{N}}(a,b)=0 holds for some b𝐙Nb\in\mathbf{Z}_{N}. Then it follows from Proposition 5.1, (2) that a+b0,1(modN)a+b\equiv 0,1\pmod{N}, which implies b=Nab=N-a or b=N+1ab=N+1-a. Since a>ba>b by (2.B), we have 2a>N2a>N or 2a>N+12a>N+1, and hence a>N/2a>N/2 in both cases. This contradiction finishes the proof. ∎

5.3 Parental pair and ancestor of zero in 𝐙N,±\mathbf{Z}_{N,\pm}

In this subsection, we show that, if we employ the index set 𝐙N,±\mathbf{Z}_{N,\pm} instead of 𝐙N\mathbf{Z}_{N}, then the assertions of Proposition 5.1 and Corollary 5.1 are simplified considerably.

Definition 5.5.

For any (a,b)𝐙N,±×𝐙N,±(a,b)\in\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm}, 𝐙N,±\mathbf{Z}_{N,\pm}-average av𝐙N,±(a,b)av_{\mathbf{Z}_{N},\pm}(a,b) is defined by

av𝐙N,±(a,b)=φN(av𝐙N(φN1(a),φN1(b))).\displaystyle av_{\mathbf{Z}_{N},\pm}(a,b)=\varphi_{N}(av_{\mathbf{Z}_{N}}(\varphi_{N}^{-1}(a),\varphi_{N}^{-1}(b))).

Furthermore, on 𝐑N,±n=φN(𝐑Nn)\mathbf{R}_{N,\pm}^{n}=\varphi_{N}(\mathbf{R}_{N}^{n}), we define ±\pm-discrete average transformation Rav±:𝐑N,±n𝐑N,±nRav_{\pm}:\mathbf{R}_{N,\pm}^{n}\rightarrow\mathbf{R}_{N,\pm}^{n} by

Rav±(𝐚)=φN(Rav(φN1(𝐚)))\displaystyle Rav_{\pm}(\mathbf{a})=\varphi_{N}(Rav(\varphi_{N}^{-1}(\mathbf{a})))

for any 𝐚𝐑N,±n\mathbf{a}\in\mathbf{R}_{N,\pm}^{n}.

The two notions, parental pairs and ancestors of zero, introduced in the previous subsection, are translated as follows:

Definition 5.6.

A pair (a,b)𝐙N,±×𝐙N,±(a,b)\in\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm} is said to be a parental pair of zero if rav±(a,b)=0rav_{\pm}(a,b)=0. We denote the set of parental pairs of zero in 𝐙N,±×𝐙N,±\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm} by ParNPar_{N}:

ParN={(a,b)𝐙N,±×𝐙N,±;rav±(a,b)=0}.\displaystyle Par_{N}=\{(a,b)\in\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm};rav_{\pm}(a,b)=0\}.

A rhythm 𝐚𝐑N,±\mathbf{a}\in\mathbf{R}_{N,\pm} is called an ancestor of zero if 0Rav±(𝐚)0\in Rav_{\pm}(\mathbf{a}).

We can determine the set ParNPar_{N} completely in the following way:

Proposition 5.2.

(0)(0) If (a,b)ParN(a,b)\in Par_{N}, then a𝐙N,0=𝐙N,{0}a\in\mathbf{Z}_{N,\leq 0}=\mathbf{Z}_{N,-}\cup\{0\}.
(1)(1) We have

ParN=ParN0ParN1,\displaystyle Par_{N}=Par_{N}^{0}\sqcup Par_{N}^{1},

where

ParN0\displaystyle Par_{N}^{0} =\displaystyle= {(k,k);0kN12}𝐙N,±×𝐙N,±,\displaystyle\{(-k,k);0\leq k\leq\lfloor\frac{N-1}{2}\rfloor\}\subset\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm},
ParN1\displaystyle Par_{N}^{1} =\displaystyle= {(k,k+1);0kN22}𝐙N,±×𝐙N,±.\displaystyle\{(-k,k+1);0\leq k\leq\lfloor\frac{N-2}{2}\rfloor\}\subset\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm}.

In particular there are NN parental pairs of zero for any NN.

Proof.

(0) By Definition 5.5, if rav±(a,b)=0rav_{\pm}(a,b)=0, then

rav(φN1(a),φN1(b))=φN1(0)=0.\displaystyle rav(\varphi_{N}^{-1}(a),\varphi_{N}^{-1}(b))=\varphi_{N}^{-1}(0)=0.

This implies by Corollary 5.1 that φN1(a)=0\varphi_{N}^{-1}(a)=0 or N2<φN1(a)N1\frac{N}{2}<\varphi_{N}^{-1}(a)\leq N-1, which is equivalent to the condition that a𝐙N,0a\in\mathbf{Z}_{N,\leq 0}.

(1) Suppose that (a,b)ParN(a,b)\in Par_{N}. Since we have (0,0),(0,1)ParN(0,0),(0,1)\in Par_{N} by Proposition 5.1, (0), we may assume that a0a\neq 0, and hence a𝐙N,a\in\mathbf{Z}_{N,-}. The latter implies in particular that φN1(a)2\varphi_{N}^{-1}(a)\geq 2. Therefore it follows from Proposition 5.3, (2) that

φN1(a)+φN1(b)0,1(modN).\displaystyle\varphi_{N}^{-1}(a)+\varphi_{N}^{-1}(b)\equiv 0,1\pmod{N}.

This is equivalent to

a+b0,1(modN)\displaystyle a+b\equiv 0,1\pmod{N} (5.6)

by Remark 4.2. Since we can put a=ka=-k with k[1,N12]k\in[1,\lfloor\frac{N-1}{2}\rfloor], the condition (5.3) implies that b=kb=k or b=k+1b=k+1. The latter alternative is excluded only when NN is odd and k=N12k=\frac{N-1}{2} by the shape of 𝐙N,±\mathbf{Z}_{N,\pm}. The last assertion follows from the equality

|ParN0|+|ParN1|=(N12+1)+(N22+1)=N.\displaystyle|Par_{N}^{0}|+|Par_{N}^{1}|=(\lfloor\frac{N-1}{2}\rfloor+1)+(\lfloor\frac{N-2}{2}\rfloor+1)=N.

This completes the proof. ∎

For small NN, the set ParNPar_{N} is given by the following:

NParN3(0,0),(0,1),(1,1)4(0,0),(0,1),(1,1),(1,2)5(0,0),(0,1),(1,1),(1,2),(2,2)6(0,0),(0,1),(1,1),(1,2),(2,2),(2,3)\displaystyle\begin{array}[]{lcl}N&\vline&Par_{N}\\ \hline\cr 3&\vline&(0,0),(0,1),(-1,1)\\ 4&\vline&(0,0),(0,1),(-1,1),(-1,2)\\ 5&\vline&(0,0),(0,1),(-1,1),(-1,2),(-2,2)\\ 6&\vline&(0,0),(0,1),(-1,1),(-1,2),(-2,2),(-2,3)\\ \end{array}

Table 5.1. ParNPar_{N} for N=3,4,5,6N=3,4,5,6

Remark 5.2.

When n=1n=1, there is only one ancestor of zero in 𝐑N,±1\mathbf{R}_{N,\pm}^{1}, namely, the singleton (0)(0). This is because the discrete average map Rav±Rav_{\pm} is defined to be the identity map on 𝐑N,±1\mathbf{R}_{N,\pm}^{1}. For this reason we assume n2n\geq 2 from now on.

Based on Proposition 5.2, we can understand completely the structure of the set of ancestors of zero, and hence the shape of the rhythm polynomial BavN0Bav_{N}^{0}. The following proposition is a rephrasing of Definition 5.4:

Proposition 5.3.

When n2n\geq 2, an element 𝐚=(a0,,an1)𝐑N,±n\mathbf{a}=(a_{0},\cdots,a_{n-1})\in\mathbf{R}_{N,\pm}^{n} is an ancestor of zero if and only if there exists an index i𝐙ni\in\mathbf{Z}_{n} such that (ai,ai+n1)ParN(a_{i},a_{i+_{n}1})\in Par_{N}.

In view of this, we introduce the following:

Definition 5.7.

For any (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\} and for any n[2,N]n\in[2,N], we put

Anc(a,b)n={𝐚𝐈N,±n;(ai,ai+n1)=(a,b) for some i𝐙n},\displaystyle Anc_{(a,b)}^{n}=\{\mathbf{a}\in\mathbf{I}_{N,\pm}^{n};(a_{i},a_{i+_{n}1})=(a,b)\mbox{ for some }i\in\mathbf{Z}_{n}\},

and

Anc(a,b)=n=2NAnc(a,b)n.\displaystyle Anc_{(a,b)}=\bigcup_{n=2}^{N}Anc_{(a,b)}^{n}. (5.8)

When (a,b)=(0,0)(a,b)=(0,0), we put

Anc(0,0)={(0)}.\displaystyle Anc_{(0,0)}=\{(0)\}.

Collecting these families of ancestors, we put

AncN=(a,b)ParNAnc(a,b),\displaystyle Anc_{N}=\bigcup_{(a,b)\in Par_{N}}Anc_{(a,b)},

and call it the set of ancestors of zero.

In order to formulate a Boolean counterpart of the notion of ancestor of zero, we put

[a,b]N,±\displaystyle[a,b]_{N,\pm} =\displaystyle= φN([φN1(a),φN1(b)]N),\displaystyle\varphi_{N}([\varphi_{N}^{-1}(a),\varphi_{N}^{-1}(b)]_{N}), (5.9)
[a,b)N,±\displaystyle\mbox{$[$}a,b)_{N,\pm} =\displaystyle= [a,b]N,±{b}\displaystyle[a,b]_{N,\pm}\setminus\{b\} (5.10)

for any (a,b)𝐙N,±×𝐙N,±(a,b)\in\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm}. Then we have the following:

Proposition 5.4.

Fix an arbitrary (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\}. For any 𝐚𝐈N,±n\mathbf{a}\in\mathbf{I}_{N,\pm}^{n} with n2n\geq 2, let 𝐯=ItoB(𝐚)𝐁N,±\mathbf{v}=ItoB(\mathbf{a})\in\mathbf{B}_{N,\pm}. Then the following three conditions are equivalent:
(1)(1) 𝐚Anc(a,b)\mathbf{a}\in Anc_{(a,b)}.
(2)(2) 𝐚[a,b]N,±={a,b}\mathbf{a}\cap[a,b]_{N,\pm}=\{a,b\}.
(3)(3) va=1,va+N1=0,,vbN1=0,vb=1v_{a}=1,v_{a+_{N}1}=0,\cdots,v_{b-_{N}1}=0,v_{b}=1.

Remark 5.3.

Here in the item (3), we denote the addition on 𝐙N,±\mathbf{Z}_{N,\pm} also by ”+N+_{N}”, which is used originally for the addition on 𝐙N\mathbf{Z}_{N}. Strictly speaking, for any (a,b)𝐙N,±×𝐙N,±(a,b)\in\mathbf{Z}_{N,\pm}\times\mathbf{Z}_{N,\pm}, we should define a new addition ”a+N,±ba+_{N,\pm}b” by the rule

a+N,±b=φN(φN1(a)+NφN1(b)).\displaystyle a+_{N,\pm}b=\varphi_{N}(\varphi_{N}^{-1}(a)+_{N}\varphi_{N}^{-1}(b)).

We, however, abuse the notation, since the context will tell us which meaning we are adopting.

Proof.

(1)(2)(1)\Rightarrow(2): Let 𝐚=(a0,,an1)\mathbf{a}=(a_{0},\cdots,a_{n-1}). The statement (1) implies by Definition 5.7 that there exists an index i𝐙ni\in\mathbf{Z}_{n} such that (ai,ai+n1)=(a,b)(a_{i},a_{i+_{n}1})=(a,b). On the other hand, by (1.8), we have

𝐙N,±=j𝐙n[aj,aj+n1)N,±.\displaystyle\mathbf{Z}_{N,\pm}=\bigsqcup_{j\in\mathbf{Z}_{n}}[a_{j},a_{j+_{n}1})_{N,\pm}.

Since the right hand side is a disjoint union, we have 𝐚[a,b]={a,b}\mathbf{a}\cap[a,b]=\{a,b\}.

(2)(3)(2)\Rightarrow(3): By the definition of the map ItoBItoB, the ii-th coordinate viv_{i}, i𝐙N,±i\in\mathbf{Z}_{N,\pm}, of 𝐯=ItoB(𝐚)\mathbf{v}=ItoB(\mathbf{a}) is given by

vi={1,i𝐚,0,i𝐚.\displaystyle v_{i}=\left\{\begin{array}[]{ll}1,&i\in\mathbf{a},\\ 0,&i\not\in\mathbf{a}.\\ \end{array}\right.

Therefore the equality 𝐚[a,b]N,±={a,b}\mathbf{a}\cap[a,b]_{N,\pm}=\{a,b\} in (2) implies the equalities va=1,va+N1=0,,vbN1=0,vb=1v_{a}=1,v_{a+_{N}1}=0,\cdots,v_{b-_{N}1}=0,v_{b}=1 in the statement (3).

(3)(1)(3)\Rightarrow(1): It follows from (3) and the definition of ItoBItoB that there exists an i𝐙N,±i\in\mathbf{Z}_{N,\pm} such that ai=a,ai+n1=ba_{i}=a,a_{i+_{n}1}=b. Hence it follows from Definition 5.7 that the statement (1) holds true. This completes the proof. ∎

In view of this proposition, we introduce the notion of 𝐁\mathbf{B}-ancestor of zero as follows:

Definition 5.8.

Every element in ItoB(AncN)𝐁N,±ItoB(Anc_{N})\subset\mathbf{B}_{N,\pm} are called a 𝐁\mathbf{B}-ancestor of zero. We denote by 𝐁\mathbf{B}-AncNAnc_{N} the set of 𝐁\mathbf{B}-ancestors of zero. Furthermore for each (a,b)ParN(a,b)\in Par_{N}, we put

𝐁-Anc(a,b)=ItoB(Anc(a,b))\displaystyle\mathbf{B}\mbox{-}Anc_{(a,b)}=ItoB(Anc_{(a,b)})

so that we have

𝐁-AncN=(a,b)ParN𝐁-Anc(a,b)\displaystyle\mathbf{B}\mbox{-}Anc_{N}=\bigcup_{(a,b)\in Par_{N}}\mathbf{B}\mbox{-}Anc_{(a,b)}

The following notation will simplify our description below:

Definition 5.9.

For any (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\}, we denote by 𝐞[a,b]𝐅2ba+1\mathbf{e}_{[a,b]}\in\mathbf{F}_{2}^{b-a+1} the Boolean vector whose ii-th coordinate 𝐞[a,b]i\mathbf{e}_{[a,b]}^{i} (1iba+1)(1\leq i\leq b-a+1) is defined by

𝐞[a,b]i={1, if i=1 or i=ba+1,0, if 1<i<ba+1.\displaystyle\mathbf{e}_{[a,b]}^{i}=\left\{\begin{array}[]{ll}1,&\mbox{ if }i=1\mbox{ or }i=b-a+1,\\ 0,&\mbox{ if }1<i<b-a+1.\\ \end{array}\right.

In other words, the vector 𝐞[a,b]\mathbf{e}_{[a,b]} is a Boolean vector of length |[a,b]||[a,b]| such that the first and the last coordinates are 11 and the others are 0.

The equivalence of (1) and (3) in Proposition 5.4 implies the following:

Corollary 5.2.

For any (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\}, let pr[a,b]:𝐁N,±𝐅2ba+1pr_{[a,b]}:\mathbf{B}_{N,\pm}\rightarrow\mathbf{F}_{2}^{b-a+1} denote the projection defined by

pr[a,b](v,,vg)=(va,,vb).\displaystyle pr_{[a,b]}(v_{\ell},\cdots,v_{g})=(v_{a},\cdots,v_{b}).

Then we have

𝐁-Anc(a,b)=pr[a,b]1(𝐞[a,b]).\displaystyle\mathbf{B}\mbox{-}Anc_{(a,b)}=pr_{[a,b]}^{-1}(\mathbf{e}_{[a,b]}).

This implies further the following:

Corollary 5.3.

For any (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\}, the number of elements in Anc(a,b)Anc_{(a,b)} is given by

|Anc(a,b)|=2N1(ba).\displaystyle|Anc_{(a,b)}|=2^{N-1-(b-a)}.
Proof.

By the bijectivity of the map ItoBItoB, we have only to count the number of elements in 𝐁\mathbf{B}-Anc[a,b]Anc_{[a,b]}, which coincides with pr[a,b]1(𝐞[a,b])pr_{[a,b]}^{-1}(\mathbf{e}_{[a,b]}) by Corollary 5.2. Hence we can compute as follows:

|Anc(a,b)|\displaystyle|Anc_{(a,b)}| =\displaystyle= |𝐁-Anc(a,b)|\displaystyle|\mathbf{B}\mbox{-}Anc_{(a,b)}|
=\displaystyle= |pr[a,b]1(𝐞[a,b])|\displaystyle|pr_{[a,b]}^{-1}(\mathbf{e}_{[a,b]})|
=\displaystyle= 2N(ba+1).\displaystyle 2^{N-(b-a+1)}.

This finishes the proof. ∎

We can deduce from this corollary that the polynomial BavN0Bav_{N}^{0} is balanced.

Definition 5.10.

A Boolean function P:𝐅2n𝐅2P:\mathbf{F}_{2}^{n}\rightarrow\mathbf{F}_{2} is said to be balanced if

|P1({0})|=|P1({1})|=2n1.\displaystyle|P^{-1}(\{0\})|=|P^{-1}(\{1\})|=2^{n-1}.

For the importance of the balanced polynomials in cryptology, we refer the reader to [4, Chapter 3].

Theorem 5.1.

The 0-th rhythm polynomial BavN0Bav_{N}^{0} is balanced.

Proof.

The set of ancestors of zero is the disjoint union of Anc(a,b)Anc_{(a,b)} ((a,b)ParN{(0,0)})((a,b)\in Par_{N}\setminus\{(0,0)\}) and Anc(0,0)Anc_{(0,0)}. Hence the total is computed through Corollary 5.3 as follows:

((a,b)ParN{(0,0)}|Anc(a,b)|)+|Anc(0,0)|\displaystyle\left(\sum_{(a,b)\in Par_{N}\setminus\{(0,0)\}}|Anc_{(a,b)}|\right)+|Anc_{(0,0)}|
=((a,b)ParN{(0,0)}2N1(ba))+1\displaystyle=\left(\sum_{(a,b)\in Par_{N}\setminus\{(0,0)\}}2^{N-1-(b-a)}\right)+1
=((a,b)ParN0{(0,0)}2N1(ba)+(a,b)ParN12N1(ba))+1\displaystyle=\left(\sum_{(a,b)\in Par_{N}^{0}\setminus\{(0,0)\}}2^{N-1-(b-a)}+\sum_{(a,b)\in Par_{N}^{1}}2^{N-1-(b-a)}\right)+1
=(1kN122N1(k(k))+0kN222N1((k+1)(k)))+1\displaystyle=\left(\sum_{1\leq k\leq\lfloor\frac{N-1}{2}\rfloor}2^{N-1-(k-(-k))}+\sum_{0\leq k\leq\lfloor\frac{N-2}{2}\rfloor}2^{N-1-((k+1)-(-k))}\right)+1
(by Proposition 5.2, (1))\displaystyle\hskip 199.16928pt(\mbox{by Proposition 5.2, (1))}
=(1kN122N12k+0kN222N1(2k+1))+1\displaystyle=\left(\sum_{1\leq k\leq\lfloor\frac{N-1}{2}\rfloor}2^{N-1-2k}+\sum_{0\leq k\leq\lfloor\frac{N-2}{2}\rfloor}2^{N-1-(2k+1)}\right)+1 (5.13)

Here we notice that the set {2k;1kN12}\{2k;1\leq k\leq\lfloor\frac{N-1}{2}\rfloor\} coincides with the set of positive even numbers in the interval [0,N1][0,N-1], and that the set {2k+1;0kN22}\{2k+1;0\leq k\leq\lfloor\frac{N-2}{2}\rfloor\} coincides with the set of odd numbers in the interval [0,N1][0,N-1]. Therefore we have

{2k;1kN12}{2k+1;0kN22}=[1,N1]\displaystyle\{2k;1\leq k\leq\lfloor\frac{N-1}{2}\rfloor\}\cup\{2k+1;0\leq k\leq\lfloor\frac{N-2}{2}\rfloor\}=[1,N-1]

Hence the rightmost side of (5.7) is equal to

(1jN12N1j)+1=(0jN22j)+1=2N1.\displaystyle\left(\sum_{1\leq j\leq N-1}2^{N-1-j}\right)+1=\left(\sum_{0\leq j\leq N-2}2^{j}\right)+1=2^{N-1}.

This completes the proof. ∎

5.4 Main theorem on the structure of the rhythm polynomial BavN0Bav_{N}^{0}

In order to state our main theorem on BavN0Bav_{N}^{0}, we introduce some notations:

Definition 5.11.

For any (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\}, we denote by f[a,b]f_{[a,b]} the Boolean polynomial given by

f[a,b]=(ya+1)(k=a+1b1yk)(yb+1).\displaystyle f_{[a,b]}=(y_{a}+1)\left(\prod_{k=a+1}^{b-1}y_{k}\right)(y_{b}+1).

Furthermore we set

f[0,0]N=(y0+1)k𝐙N,±{0}yk.\displaystyle f_{[0,0]}^{N}=(y_{0}+1)\prod_{k\in\mathbf{Z}_{N,\pm}\setminus\{0\}}y_{k}.
Remark 5.4.

Notice that the definition of f[a,b]f_{[a,b]} for any [a,b]ParN{(0,0)}[a,b]\in Par_{N}\setminus\{(0,0)\} does not depend on NN, but only on the values of a,ba,b. On the other hand the definition of f[0,0]Nf_{[0,0]}^{N} depends on NN. This is why we put ”NN” on the superscript.

The general form of the Boolean polynomial BavN0Bav_{N}^{0} is given by the following:

Theorem 5.2.

For any N3N\geq 3, we have

BavN0=(a,b)ParN{(0,0)}f[a,b]+f[0,0]N.\displaystyle Bav_{N}^{0}=\sum_{(a,b)\in Par_{N}\setminus\{(0,0)\}}f_{[a,b]}+f_{[0,0]}^{N}. (5.14)

Our proof of this theorem will be given later in the subsection 5.6, after we prepare several results concerning the algebraic standard forms of special Boolean functions.

5.5 Preliminaries for the proof of Theorem 5.2

For any S={s1,,sk}𝐙NS=\{s_{1},\cdots,s_{k}\}\subset\mathbf{Z}_{N}, we denote by prS:𝐅2N𝐅2|S|pr_{S}:\mathbf{F}_{2}^{N}\rightarrow\mathbf{F}_{2}^{|S|} the projection defined by

prS(x0,,xN1)=(xs1,,xsk)\displaystyle pr_{S}(x_{0},\cdots,x_{N-1})=(x_{s_{1}},\cdots,x_{s_{k}})

for any (x0,,xN1)𝐅2N(x_{0},\cdots,x_{N-1})\in\mathbf{F}_{2}^{N}.

Proposition 5.5.

Let P=P(x0,,xN1):𝐅2N𝐅2P=P(x_{0},\cdots,x_{N-1}):\mathbf{F}_{2}^{N}\rightarrow\mathbf{F}_{2} be a Boolean function with the following property: There exists a subset S={s1,,sk}𝐙NS=\{s_{1},\cdots,s_{k}\}\subset\mathbf{Z}_{N} such that

P1({1})=prS1({(b1,,bk)})\displaystyle P^{-1}(\{1\})=pr_{S}^{-1}(\{(b_{1},\cdots,b_{k})\})

for some (b1,,bk)𝐅2k(b_{1},\cdots,b_{k})\in\mathbf{F}_{2}^{k}. Then we have

P(x0,,xN1)=xs1(b1)xs2(b2)xsk(bk),\displaystyle P(x_{0},\cdots,x_{N-1})=x_{s_{1}}^{(b_{1})}x_{s_{2}}^{(b_{2})}\cdots x_{s_{k}}^{(b_{k})},

where we put

xi(0)\displaystyle x_{i}^{(0)} =\displaystyle= xi+1,\displaystyle x_{i}+1,
xi(1)\displaystyle x_{i}^{(1)} =\displaystyle= xi,\displaystyle x_{i},

for any Boolean variable xi,i𝐙Nx_{i},i\in\mathbf{Z}_{N}.

Proof.

Note that

prS1({(b1,,bk)})\displaystyle pr_{S}^{-1}(\{(b_{1},\cdots,b_{k})\})
={(a0,,aN1)𝐅2N;(as1,,ask)=(b1,,bk)}.\displaystyle=\{(a_{0},\cdots,a_{N-1})\in\mathbf{F}_{2}^{N};(a_{s_{1}},\cdots,a_{s_{k}})=(b_{1},\cdots,b_{k})\}.

Hence, denoting the complement 𝐙NS\mathbf{Z}_{N}\setminus S by T={j1,,jNk}T=\{j_{1},\cdots,j_{N-k}\}, we see that

P(x0,,xN1)=(xj1+xj1¯)(xjNk+xjNk¯)xs1(b1)xs2(b2)xsk(bk)\displaystyle P(x_{0},\cdots,x_{N-1})=(x_{j_{1}}+\overline{x_{j_{1}}})\cdots(x_{j_{N-k}}+\overline{x_{j_{N-k}}})x_{s_{1}}^{(b_{1})}x_{s_{2}}^{(b_{2})}\cdots x_{s_{k}}^{(b_{k})}

by the algorithm explained in the subsection 4.1. Since xjt+xjt¯=1x_{j_{t}}+\overline{x_{j_{t}}}=1 holds for any t[1,Nk]t\in[1,N-k], we finish the proof. ∎

We also need an equivalent form of the above proposition, where we change the set of indices 𝐙N\mathbf{Z}_{N} to 𝐙N,±\mathbf{Z}_{N,\pm}. Furthermore, since we have observed in the Table 4.5 that the Boolean variable ”y=x+1y=x+1” is superior to the original xx for our study, we employ the negated variables as our basic building blocks. The following proposition is a direct consequence of Proposition 5.5:

Proposition 5.6.

Let P=P(y,,yg):𝐅2N𝐅2P=P(y_{\ell},\cdots,y_{g}):\mathbf{F}_{2}^{N}\rightarrow\mathbf{F}_{2} be a Boolean function with the following property: There exists a subset S={s1,,sk}𝐙N,±S=\{s_{1},\cdots,s_{k}\}\subset\mathbf{Z}_{N,\pm} such that

P1({1})=prS1({(b1,,bk)})\displaystyle P^{-1}(\{1\})=pr_{S}^{-1}(\{(b_{1},\cdots,b_{k})\})

holds for some(b1,,bk)𝐅2k(b_{1},\cdots,b_{k})\in\mathbf{F}_{2}^{k}. Then we have

P(y,,yg)=ys1(b1)ys2(b2)ysk(bk),\displaystyle P(y_{\ell},\cdots,y_{g})=y_{s_{1}}^{(b_{1})}y_{s_{2}}^{(b_{2})}\cdots y_{s_{k}}^{(b_{k})},

where we put

yi(0)\displaystyle y_{i}^{(0)} =\displaystyle= yi,\displaystyle y_{i},
yi(1)\displaystyle y_{i}^{(1)} =\displaystyle= yi+1,\displaystyle y_{i}+1,

for any Boolean variable yi,i𝐙N,±y_{i},i\in\mathbf{Z}_{N,\pm}.

5.6 Ancestors of zero when N=6N=6

Before we give a proof of Theorem 5.2, we examine the case when N=6N=6. This might help the reader to understand our proof for general cases.

The set Par6Par_{6} of parental pairs of zero has been found in Proposition 5.2 as follows:

Par6={(0,0),(0,1),(1,1),(1,2),(2,2),(2,3)}.\displaystyle Par_{6}=\{(0,0),(0,1),(-1,1),(-1,2),(-2,2),(-2,3)\}.

The figure below illustrates the ancestors of zero when N=6N=6:

[Uncaptioned image] [Uncaptioned image] [Uncaptioned image]
[Uncaptioned image] [Uncaptioned image] [Uncaptioned image]

Fig. 5.1. Ancestors of zero for N=6N=6

Here are six families (1)-(6) of ancestors of 0, which are determined in Proposition 5.2 and Corollary 5.2. Each of six large circles represents the unit circle on the complex plane, and each member kk of 𝐙6,±\mathbf{Z}_{6,\pm} are placed at the point exp(kπi/3)\exp(k\pi i/3). The meanings of symbols in the figure are as follows:

(1) The numbers surrounding the unit circle specify the names of the elements in 𝐙6,±\mathbf{Z}_{6,\pm}.
(2) Black dots represent the parental pairs of zero found in Proposition 5.2.
(3) Each of small blank circle signifies that the corresponding element does not appear in the ancestors.
(4) Each of circles having ”\forall” in its interior means that the corresponding element may or may not be present in the ancestors.

We should be careful about the rather exceptional case when n=1n=1, which is depicted as ”(6)”. Thus the 𝐁\mathbf{B}-ancestors of zero which correspond to the six figures are given by the following list:

name𝐁-ancestors of zero in 𝐁6,±number of elements(1){e[2,3]}1(2)pr[2,2]1({e[2,2]})2(3)pr[1,2]1({e[1,2]})4(4)pr[1,1]1({e[1,1]})8(5)pr[0,1]1({e[0,1]})16(6){(0,0,1,0,0,0)}1Total32(=261)\displaystyle\begin{array}[]{lll}\mbox{name}&\mbox{$\mathbf{B}$-ancestors of zero in }\mathbf{B}_{6,\pm}&\mbox{number of elements}\\ \hline\cr(1)&\{e_{[-2,3]}\}&1\\ (2)&pr_{[-2,2]}^{-1}(\{e_{[-2,2]}\})&2\\ (3)&pr_{[-1,2]}^{-1}(\{e_{[-1,2]}\})&4\\ (4)&pr_{[-1,1]}^{-1}(\{e_{[-1,1]}\})&8\\ (5)&pr_{[0,1]}^{-1}(\{e_{[0,1]}\})&16\\ (6)&\{(0,0,1,0,0,0)\}&1\\ Total&&32(=2^{6-1})\\ \end{array}

Table 5.2. Six families of the 𝐁\mathbf{B}-ancestors of zero for N=6N=6

Let SiS_{i} denote the subset of 𝐁6,±\mathbf{B}_{6,\pm} which is specified in the ii-th row (1i6)(1\leq i\leq 6), and let fSif_{S_{i}} be the Boolean function which takes the value 1 (resp. 0) at every element of SiS_{i} (resp. 𝐁6,±Si)\mathbf{B}_{6,\pm}\setminus S_{i}). Then it follows from Proposition 5.6 and Definition 5.11 that

fS1\displaystyle f_{S_{1}} =\displaystyle= f[2,3]=(y2+1)y1y0y1y2(y3+1),\displaystyle f_{[-2,3]}=(y_{-2}+1)y_{-1}y_{0}y_{1}y_{2}(y_{3}+1),
fS2\displaystyle f_{S_{2}} =\displaystyle= f[2,2]=(y2+1)y1y0y1(y2+1),\displaystyle f_{[-2,2]}=(y_{-2}+1)y_{-1}y_{0}y_{1}(y_{2}+1),
fS3\displaystyle f_{S_{3}} =\displaystyle= f[1,2]=(y1+1)y0y1(y2+1),\displaystyle f_{[-1,2]}=(y_{-1}+1)y_{0}y_{1}(y_{2}+1),
fS4\displaystyle f_{S_{4}} =\displaystyle= f[1,1]=(y1+1)y0(y1+1),\displaystyle f_{[-1,1]}=(y_{-1}+1)y_{0}(y_{1}+1),
fS5\displaystyle f_{S_{5}} =\displaystyle= f[0,1]=(y0+1)(y1+1),\displaystyle f_{[0,1]}=(y_{0}+1)(y_{1}+1),
fS6\displaystyle f_{S_{6}} =\displaystyle= f[0,0]6=(y0+1)y2y1y1y2y3.\displaystyle f_{[0,0]}^{6}=(y_{0}+1)y_{-2}y_{-1}y_{1}y_{2}y_{3}.

By summing these up, we have

Bav06=(a,b)Par6{(0,0)}f[a,b]+f[0,0]6,\displaystyle Bav_{0}^{6}=\sum_{(a,b)\in Par_{6}\setminus\{(0,0)\}}f_{[a,b]}+f_{[0,0]}^{6},

which shows the validity of (5.6) of Theorem 5.2 when N=6N=6. Furthermore the total of the numbers of ancestors in the lines (1)-(6) is equal to 32, which is exactly the half of the number 64 of Boolean vectors in 𝐁6,±\mathbf{B}_{6,\pm}. Therefore Bav60Bav_{6}^{0} is balanced as is assured by Theorem 5.1.

5.7 Proof of Theorem 5.2

For any subset U𝐁N,±U\subset\mathbf{B}_{N,\pm}, we denote by PUP_{U} the polynomial function on 𝐁N,±\mathbf{B}_{N,\pm} such that

PU1({1})=U.\displaystyle P_{U}^{-1}(\{1\})=U.

Namely PUP_{U} takes the value 1 on UU and the value 0 outside UU. Recall that we have 𝐁\mathbf{B}-AncN(a,b)=pr[a,b]1({e[a,b]})Anc_{N}(a,b)=pr_{[a,b]}^{-1}(\{e_{[a,b]}\}) for any (a,b)ParN{(0,0)}(a,b)\in Par_{N}\setminus\{(0,0)\} by Corollary 5.2. This implies by Proposition 5.6 and definition 5.11 that

P𝐁AncN(a,b)=f[a,b].\displaystyle P_{\mathbf{B}-Anc_{N}(a,b)}=f_{[a,b]}.

The remaining parental pair (0,0)ParN(0,0)\in Par_{N} corresponds to the one-element rhythm (0)𝐑N1(0)\in\mathbf{R}_{N}^{1} (see Remark 5.1). Since RtoB((0))=(vi)i𝐙N,±RtoB((0))=(v_{i})_{i\in\mathbf{Z}_{N,\pm}}, where

vi={1, if i=0,0, if i0,\displaystyle v_{i}=\left\{\begin{array}[]{ll}1,&\mbox{ if }i=0,\\ 0,&\mbox{ if }i\neq 0,\\ \end{array}\right.

we have

P𝐁AncN(0,0)=f[0,0].\displaystyle P_{\mathbf{B}-Anc_{N}(0,0)}=f_{[0,0]}.

Thus we obtain the equality (5.8). This completes the proof of Theorem 5.2.

We check the validity of Theorem 5.2 by our computational results in Subsection 4.2.

Example 5.1. The case when N=3N=3: The equality (5.8) in Theorem 5.2 asserts that

Bav30=(a,b)Par3{(0,0)}f[a,b]+f[0,0]3.\displaystyle Bav_{3}^{0}=\sum_{(a,b)\in Par_{3}\setminus\{(0,0)\}}f_{[a,b]}+f_{[0,0]}^{3}.

Recall that

Par3={(0,0),(0,1),(1,1)}\displaystyle Par_{3}=\{(0,0),(0,1),(-1,1)\}

by Proposition 5.2, and the corresponding polynomials f[a,b]f_{[a,b]}, (a,b)Par3{(0,0)}(a,b)\in Par_{3}\setminus\{(0,0)\} and f[0,0]3f_{[0,0]}^{3} are defined in Definition 5.11 as follows:

f[0,1]\displaystyle f_{[0,1]} =\displaystyle= (y0+1)(y1+1),\displaystyle(y_{0}+1)(y_{1}+1),
f[1,1]\displaystyle f_{[-1,1]} =\displaystyle= (y1+1)y0(y1+1),\displaystyle(y_{-1}+1)y_{0}(y_{1}+1),
f[0,0]3\displaystyle f_{[0,0]}^{3} =\displaystyle= (y0+1)y1y1.\displaystyle(y_{0}+1)y_{-1}y_{1}.

Hence we have

Bav30\displaystyle Bav_{3}^{0} =\displaystyle= (y0+1)(y1+1)+(y1+1)y0(y1+1)+(y0+1)y1y1\displaystyle(y_{0}+1)(y_{1}+1)+(y_{-1}+1)y_{0}(y_{1}+1)+(y_{0}+1)y_{-1}y_{1}
=\displaystyle= (1+y0¯+y1+y0y1¯)+y0(1¯+y1+y1¯+y1y1¯)+y1(y0¯+1)y1\displaystyle(1+\underline{y_{0}}+y_{1}+\underline{y_{0}y_{1}})+y_{0}(\underline{1}+y_{-1}+\underline{y_{1}}+\underline{y_{-1}y_{1}})+y_{-1}(\underline{y_{0}}+1)y_{1}
=\displaystyle= 1+y1+y1y0+y1y1,\displaystyle 1+y_{1}+y_{-1}y_{0}+y_{-1}y_{1},
(by cancelling the underlined terms)\displaystyle\hskip 85.35826pt(\mbox{by cancelling the underlined terms})

and we see that the rightmost side coincides with Table 4.5 for N=3N=3.

Example 5.2. The case when N=4N=4: By Theorem 5.2, Proposition 5.2, and Definition 5.11, we can compute as follows:

Bav40\displaystyle Bav_{4}^{0} =\displaystyle= (a,b)Par4{(0,0)}f[a,b]+f[0,0]4\displaystyle\sum_{(a,b)\in Par_{4}\setminus\{(0,0)\}}f_{[a,b]}+f_{[0,0]}^{4}
=\displaystyle= f[0,1]+f[1,1]+f[1,2]+f[0,0]4\displaystyle f_{[0,1]}+f_{[-1,1]}+f_{[-1,2]}+f_{[0,0]}^{4}
=\displaystyle= (y0+1)(y1+1)+(y1+1)y0(y1+1)+(y1+1)y0y1(y2+1)\displaystyle(y_{0}+1)(y_{1}+1)+(y_{-1}+1)y_{0}(y_{1}+1)+(y_{-1}+1)y_{0}y_{1}(y_{2}+1)
+(y0+1)y1y1y2\displaystyle\hskip 14.22636pt+(y_{0}+1)y_{-1}y_{1}y_{2}
=\displaystyle= (1+y0¯+y1+y0y1¯)+y0(1¯+y1+y1¯+y1y1¯)\displaystyle(1+\underline{y_{0}}+y_{1}+\underline{y_{0}y_{1}})+y_{0}(\underline{1}+y_{-1}+\underline{y_{1}}+\underline{y_{-1}y_{1}})
+y0y1(1+y1¯+y2+y1y2¯))+(y0¯+1)y1y1y2\displaystyle+y_{0}y_{1}(1+\underline{y_{-1}}+y_{2}+\underline{y_{-1}y_{2}}))+(\underline{y_{0}}+1)y_{-1}y_{1}y_{2}
=\displaystyle= 1+y1+y1y0+y0y1+y0y1y2+y1y1y2.\displaystyle 1+y_{1}+y_{-1}y_{0}+y_{0}y_{1}+y_{0}y_{1}y_{2}+y_{-1}y_{1}y_{2}.
(by cancelling the underlined terms)\displaystyle\hskip 85.35826pt(\mbox{by cancelling the underlined terms})

The rightmost side coincides with Table 4.5 for N=4N=4.

5.8 Recurrence formula

Inspecting the shapes of BavN0Bav_{N}^{0} found in Theorem 5.2, we are led naturally to the following:

Theorem 5.3.

When NN is even and N=2mN=2m, we have

Bav2m0=Bav2m10+ym+2y1y1ym1(ym+1)(y0+ym+1).\displaystyle Bav_{2m}^{0}=Bav_{2m-1}^{0}+y_{-m+2}\cdots y_{-1}y_{1}\cdots y_{m-1}(y_{m}+1)(y_{0}+y_{-m+1}). (5.17)

When NN is odd and N=2m+1N=2m+1, we have

Bav2m+10=Bav2m0+ym+1y1y1ym1(ym+1)(y0+ym).\displaystyle Bav_{2m+1}^{0}=Bav_{2m}^{0}+y_{-m+1}\cdots y_{-1}y_{1}\cdots y_{m-1}(y_{-m}+1)(y_{0}+y_{m}). (5.18)
Proof.

It follows from Proposition 5.2 that

ParNParN+1\displaystyle Par_{N}\subset Par_{N+1}

holds for any N3N\geq 3. Hence we have only to consider their difference. We divide the proof into two parts according to the parity of NN.

1) The case when N=2m4N=2m\geq 4. It follows from Proposition 5.2 that

Par2mPar2m1={[m+1,m]}.\displaystyle Par_{2m}\setminus Par_{2m-1}=\{[-m+1,m]\}.

Before we start our computation, for the convenience of the reader, we recall what members are in 𝐙2m,±\mathbf{Z}_{2m,\pm} or in 𝐙2m1,±\mathbf{Z}_{2m-1,\pm}:

𝐙2m,±\displaystyle\mathbf{Z}_{2m,\pm} =\displaystyle= {x𝐙;m+1xm},\displaystyle\{x\in\mathbf{Z};-m+1\leq x\leq m\},
𝐙2m1,±\displaystyle\mathbf{Z}_{2m-1,\pm} =\displaystyle= {x𝐙;m+1xm1}.\displaystyle\{x\in\mathbf{Z};-m+1\leq x\leq m-1\}.

Keeping in mind the fact that 1=1-1=1 in 𝐅2\mathbf{F}_{2}, we can compute as follows:

Bav2m0Bav2m10\displaystyle Bav_{2m}^{0}-Bav_{2m-1}^{0}
=\displaystyle= f[m+1,m]+f[0,0]2mf[0,0]2m1\displaystyle f_{[-m+1,m]}+f_{[0,0]}^{2m}-f_{[0,0]}^{2m-1}
=\displaystyle= (ym+1+1)(k=m+2m1yk)(ym+1)\displaystyle(y_{-m+1}+1)\left(\prod_{k=-m+2}^{m-1}y_{k}\right)(y_{m}+1)
+((y0+1)k𝐙2m,±{0}yk(y0+1)k𝐙2m1,±{0}yk)\displaystyle+\left((y_{0}+1)\prod_{k\in\mathbf{Z}_{2m,\pm}\setminus\{0\}}y_{k}-(y_{0}+1)\prod_{k\in\mathbf{Z}_{2m-1,\pm}\setminus\{0\}}y_{k}\right)
=\displaystyle= (ym+1+1)(k=m+2m1yk)(ym+1)\displaystyle(y_{-m+1}+1)\left(\prod_{k=-m+2}^{m-1}y_{k}\right)(y_{m}+1)
+(y0+1)(k𝐙2m1,±{0}yk)(ym1)\displaystyle\hskip 14.22636pt+(y_{0}+1)\left(\prod_{k\in\mathbf{Z}_{2m-1,\pm}\setminus\{0\}}y_{k}\right)(y_{m}-1)
=\displaystyle= ym+2y1y1ym1\displaystyle y_{-m+2}\cdots y_{-1}y_{1}\cdots y_{m-1}
×((ym+1+1)y0(ym+1)+ym+1(y0+1)(ym+1))\displaystyle\times\left((y_{-m+1}+1)y_{0}(y_{m}+1)+y_{-m+1}(y_{0}+1)(y_{m}+1)\right)
=\displaystyle= ym+2y1y1ym1(ym+1)\displaystyle y_{-m+2}\cdots y_{-1}y_{1}\cdots y_{m-1}(y_{m}+1)
×((ym+1+1)y0+ym+1(y0+1))\displaystyle\times\left((y_{-m+1}+1)y_{0}+y_{-m+1}(y_{0}+1)\right)
=\displaystyle= ym+2y1y1ym1(ym+1)(y0+ym+1).\displaystyle y_{-m+2}\cdots y_{-1}y_{1}\cdots y_{m-1}(y_{m}+1)(y_{0}+y_{-m+1}).

This shows the validity of (5.9).

2) The case when N=2m+15N=2m+1\geq 5. It follows Proposition 5.2 that

Par2m+1Par2m={[m,m]}.\displaystyle Par_{2m+1}\setminus Par_{2m}=\{[-m,m]\}.

This time, the members of 𝐙2m+1,±\mathbf{Z}_{2m+1,\pm} and 𝐙2m,±\mathbf{Z}_{2m,\pm} are given by

𝐙2m+1,±\displaystyle\mathbf{Z}_{2m+1,\pm} =\displaystyle= {x𝐙;mxm},\displaystyle\{x\in\mathbf{Z};-m\leq x\leq m\},
𝐙2m,±\displaystyle\mathbf{Z}_{2m,\pm} =\displaystyle= {x𝐙;m+1xm}.\displaystyle\{x\in\mathbf{Z};-m+1\leq x\leq m\}.

Hence we can compute as follows:

Bav2m+10Bav2m0\displaystyle Bav_{2m+1}^{0}-Bav_{2m}^{0}
=\displaystyle= f[m,m]+f[0,0]2m+1f[0,0]2m\displaystyle f_{[-m,m]}+f_{[0,0]}^{2m+1}-f_{[0,0]}^{2m}
=\displaystyle= (ym+1)(k=m+1m1yk)(ym+1)\displaystyle(y_{-m}+1)\left(\prod_{k=-m+1}^{m-1}y_{k}\right)(y_{m}+1)
+((y0+1)k𝐙2m+1,±{0}yk(y0+1)k𝐙2m,±{0}yk)\displaystyle+\left((y_{0}+1)\prod_{k\in\mathbf{Z}_{2m+1,\pm}\setminus\{0\}}y_{k}-(y_{0}+1)\prod_{k\in\mathbf{Z}_{2m,\pm}\setminus\{0\}}y_{k}\right)
=\displaystyle= (ym+1)(k=m+1m1yk)(ym+1)\displaystyle(y_{-m}+1)\left(\prod_{k=-m+1}^{m-1}y_{k}\right)(y_{m}+1)
+(y0+1)(k𝐙2m,±{0}yk)(ym1)\displaystyle+(y_{0}+1)\left(\prod_{k\in\mathbf{Z}_{2m,\pm}\setminus\{0\}}y_{k}\right)(y_{-m}-1)
=\displaystyle= ym+1y1y1ym1\displaystyle y_{-m+1}\cdots y_{-1}y_{1}\cdots y_{m-1}
×((ym+1)y0(ym+1)+(ym+1)(y0+1)ym)\displaystyle\times\left((y_{-m}+1)y_{0}(y_{m}+1)+(y_{-m}+1)(y_{0}+1)y_{m}\right)
=\displaystyle= ym+1y1y1ym1(ym+1)\displaystyle y_{-m+1}\cdots y_{-1}y_{1}\cdots y_{m-1}(y_{-m}+1)
×(y0(ym+1)+(y0+1)ym)\displaystyle\times\left(y_{0}(y_{m}+1)+(y_{0}+1)y_{m}\right)
=\displaystyle= ym+1y1y1ym1(ym+1)(y0+ym).\displaystyle y_{-m+1}\cdots y_{-1}y_{1}\cdots y_{m-1}(y_{-m}+1)(y_{0}+y_{m}).

This shows the validity of (5.10). This completes the proof. ∎

We have displayed in Table 4.5 the rhythm polynomials BavN0Bav_{N}^{0} for N=3,,6N=3,\cdots,6. We reexamine the results through Theorem 5.3.

Example 5.3. The case when N=4N=4. It follows from (5.9) with m=2m=2 that

Bav40\displaystyle Bav_{4}^{0} =\displaystyle= Bav30+y1(y2+1)(y0+y1)\displaystyle Bav_{3}^{0}+y_{1}(y_{2}+1)(y_{0}+y_{-1})
=\displaystyle= (1+y1+y1y0+y1y1¯)+y1(y1y2+y0y2+y1¯+y0)\displaystyle(1+y_{1}+y_{-1}y_{0}+\underline{y_{-1}y_{1}})+y_{1}(y_{-1}y_{2}+y_{0}y_{2}+\underline{y_{-1}}+y_{0})
=\displaystyle= 1+y1+y1y0+y0y1+y1y1y2+y0y1y2,\displaystyle 1+y_{1}+y_{-1}y_{0}+y_{0}y_{1}+y_{-1}y_{1}y_{2}+y_{0}y_{1}y_{2},
(by canceling the underlined terms)\displaystyle\hskip 85.35826pt(\mbox{by canceling the underlined terms})

where the last polynomial coincides with the result in Table 4.5 for N=4N=4.

Example 5.4. The case when N=5N=5. It follows from (5.10) with m=2m=2 that

Bav50\displaystyle Bav_{5}^{0} =\displaystyle= Bav40+y1y1(y2+1)(y0+y2)\displaystyle Bav_{4}^{0}+y_{-1}y_{1}(y_{-2}+1)(y_{0}+y_{2})
=\displaystyle= (1+y1+y1y0+y0y1+y1y1y2¯+y0y1y2)\displaystyle(1+y_{1}+y_{-1}y_{0}+y_{0}y_{1}+\underline{y_{-1}y_{1}y_{2}}+y_{0}y_{1}y_{2})
+y1y1(y2y0+y2y2+y0+y2¯)\displaystyle\hskip 56.9055pt+y_{-1}y_{1}(y_{-2}y_{0}+y_{-2}y_{2}+y_{0}+\underline{y_{2}})
=\displaystyle= 1+y1+y1y0+y0y1+y1y0y1+y0y1y2\displaystyle 1+y_{1}+y_{-1}y_{0}+y_{0}y_{1}+y_{-1}y_{0}y_{1}+y_{0}y_{1}y_{2}
+y2y1y0y1+y2y1y1y2,\displaystyle\hskip 56.9055pt+y_{-2}y_{-1}y_{0}y_{1}+y_{-2}y_{-1}y_{1}y_{2},
(by canceling the underlined terms)\displaystyle\hskip 85.35826pt(\mbox{by canceling the underlined terms})

which coincides with the result in Table 4.5 for N=5N=5.

References
[1][1]
Cusick, T. W., Stănică, Cryptographic Boolean Functions and Applications. Elsevier/Academic Press, Amsterdam, the Netherlands, 2009.
[2][2] Hazama, F., Iterative method of construction for amooth rhythms. Journal of Mathematics and Music(2021), On-Line-First, 1-20,
https://doi.org/10.1080/17459737.2021.1924303
[3][3] MacWilliams, F. J., Sloane, N. J. A., The Theory of Error-Correcting Codes. Elsevier/NorthHolland, Amsterdam, the Netherlands, 1977.
[4][4] Whitesitt, J. E., Boolean Algebra and Its Application, Dover Bookds on Computer science, 2010.