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

Complex algebras of arithmeticthanks: The authors gratefully acknowledge the support of the EPSRC, grant number EP/F069154/1. Ivo Düntsch also acknowledges support from the Natural Sciences and Engineering Research Council of Canada.

Ivo Düntsch
Department of Computer Science,
Brock University,
St. Catharines, ON, L2S 3A1
Canada
duentsch@brocku.ca
   Ian Pratt–Hartmann
School of Computer Science,
University of Manchester,
Oxford Road, Manchester, M13 9PL
United Kingdom,
ipratt@cs.man.ac.uk
Abstract

An arithmetic circuit is a labeled, acyclic directed graph specifying a sequence of arithmetic and logical operations to be performed on sets of natural numbers. Arithmetic circuits can also be viewed as the elements of the smallest subalgebra of the complex algebra of the semiring of natural numbers. In the present paper we investigate the algebraic structure of complex algebras of natural numbers and make some observations regarding the complexity of various theories of such algebras.

1 Introduction

Let ω\omega be the set of natural numbers {0,1,2,}\{0,1,2,\ldots\}. An arithmetic circuit (AC) [11, 12] is a labeled, acyclic directed graph specifying a sequence of arithmetic and logical operations to be performed on sets of natural numbers. Each node in this graph evaluates to a set of natural numbers, representing a stage of the computation performed by the circuit. Nodes without predecessors in the graph are called input nodes, and their labels are singleton sets of natural numbers. Nodes with predecessors in the graph are called arithmetic gates, and their labels indicate operations to be performed on the values of their immediate predecessors; the results of these operations are then taken to be the values of the arithmetic gates in question. One of the nodes in the graph (usually, a node with no successors) is designated as the circuit output; the set of natural numbers to which it evaluates is taken to be the value of the circuit as a whole.

More formally, an arithmetic circuit is a structure C=G,E,gC,αC=\langle G,E,g_{C},\alpha\rangle, where G,E\langle G,E\rangle is a finite acyclic and asymmetric graph over 2ω2^{\omega}, In(g)2\operatorname{In}(g)\leq 2 for all gGg\in G, and α:G{,,,+,}{{n}:nω}{,ω}\alpha:G\to\{\cup,\cap,{}^{-},\bm{+},\bullet\}\cup\{\{n\}:n\in\omega\}\cup\{\emptyset,\omega\} is a labeling function for which

(1.1) α(g){{{n}:nω}{,},if In(g)=0,{},if In(g)=1,{,,+,},if In(g)=2.\displaystyle\alpha(g)\in\begin{cases}\{\{n\}:n\in\omega\}\cup\{\emptyset,\mathbb{N}\},&\text{if }\operatorname{In}(g)=0,\\ \{{}^{-}\},&\text{if }\operatorname{In}(g)=1,\\ \{\cup,\cap,\bm{+},\bullet\},&\text{if }\operatorname{In}(g)=2.\end{cases}

Here, In(g)\operatorname{In}(g) is the in–degree of gg and +\bm{+} and \bullet are the complex extensions of ++ and \cdot, i.e.

(1.2) a+b:={k+n:ka,nb},ab:={kn:ka,nb}.\displaystyle a\bm{+}b:=\{k+n:k\in a,\ n\in b\},\ a\bullet b:=\{k\cdot n:k\in a,\ n\in b\}.

gCg_{C} is called the output gate; if In(g)=0\operatorname{In}(g)=0, we call gg an input gate or a source.

The arithmetical interpretation of CC is as follows:

  1. (i)

    If In(g)=0\operatorname{In}(g)=0, then I(g)=α(g)I(g)=\alpha(g).

  2. (ii)

    If In(g)=1\operatorname{In}(g)=1, and gg^{\prime} is the unique predecessor of gg, then I(g)=I(g)I(g)=\mathbb{N}\setminus{I(g^{\prime})}.

  3. (iii)

    If In(g)=2\operatorname{In}(g)=2, and g0,g1g_{0},g_{1} are the two predecessors of gg, then I(g):=I(g0)α(g)I(g1)I(g):=I(g_{0})~\alpha(g)~I(g_{1}).

I(C)I(C) is defined as I(gC)I(g_{C}).

Fig. 1 shows two examples of arithmetic circuits, where the output gate is indicated by the double circle. In Fig. 1a, Node 1 evaluates to {1}\{1\}, and Node 2 to ω\omega; hence, Node 3 evaluates to {1}+{1}={2}\{1\}\bm{+}\{1\}=\{2\}, and Node 4, the output of the circuit, to {2}ω\{2\}\bullet\omega, i.e. the set of even numbers. The circuit of Fig. 1b functions similarly: Node 2 evaluates to {0}{nω:n2}\{0\}\cup\{n\in\omega:n\geq 2\}, and Node 3 to {0}{nω:n is composite}\{0\}\cup\{n\in\omega:\text{$n$ is composite}\}; hence, Node 4 evaluates to the set numbers which are either prime or equal to 1, and Node 5, the output of the circuit, to the set of primes. We say that the circuits of Fig. 1a and Fig. 1b define, respectively, the set of even numbers and the set of primes. Any arithmetic circuit defines a set of numbers in this way.

Figure 1: Arithmetic circuits defining: (a) the set of even numbers; (b) the set of primes. The integers next to the nodes are for reference only.
Refer to caption1+\bm{+}\bullet342ω\omega{1}\{1\}

(a)

Refer to caption¯\bar{\ }¯\bar{\ }\cap{1}\{1\}\bullet

(b)

If 𝒪{,,,+,}\mathcal{O}\subseteq\{\cap,\cup,{}^{-},\bm{+},\bullet\}, an 𝒪\mathcal{O} – circuit is an arithmetic circuit whose non–input labels are among those contained in 𝒪\mathcal{O}. Let

(1.3) MC(𝒪)={C,n:C is an 𝒪 – circuit,nI(C)}.\displaystyle MC(\mathcal{O})=\{\langle C,n\rangle:C\text{ is an $\mathcal{O}$ -- circuit},n\in I(C)\}.

The membership problem for 𝒪\mathcal{O} is the question whether MC(𝒪)MC(\mathcal{O}) is decidable [11]. In other words, is there an algorithm which decides membership of an arbitrary nωn\in\omega in an arbitrary output CC of an 𝒪\mathcal{O} – circuit? If the problem is decidable, then its complexity is of interest. For almost all cases of 𝒪\mathcal{O}, the complexities have been determined by McKenzie and Wagner, [11]. The question whether MC(𝒪)MC(\mathcal{O}) is decidable where 𝒪={,,,+,}\mathcal{O}=\{\cap,\cup,{}^{-},\bm{+},\bullet\} is still open. The table of complexities for the membership problem where all Boolean operators are present is given in Table 1.

Table 1: Complexity results for MC [11]
𝒪Lower boundUpper bound,,,+PSPACEPSPACE,,,PSPACEPSPACE,,,+,NEXPTIME?\begin{array}[]{|l|c|c|}\hline\cr\mathcal{O}&\text{Lower bound}&\text{Upper bound}\\ \hline\cr\cap,\cup,{}^{-},\bm{+}&\text{PSPACE}&\text{PSPACE}\\ \cap,\cup,{}^{-},\bullet&\text{PSPACE}&\text{PSPACE}\\ \cap,\cup,{}^{-},\bm{+},\bullet&\text{NEXPTIME}&?\\ \hline\cr\end{array}

Algebraically speaking, an arithmetic circuit can be regarded as a well – formed term over an alphabet 𝒜\mathcal{A} containing operations from {,,,,ω,+,}\{\cap,\cup,{}^{-},\emptyset,\omega,\bm{+},\bullet\} and constants from {{n}:nω}\{\{n\}:n\in\omega\} as input gates. If +\bm{+} is present, then {0}\{0\} will suffice since

(1.4) {1}={0}¯+{0}¯¯{0}¯.\displaystyle\{1\}=\overline{\overline{\{0\}}\bm{+}\overline{\{0\}}}\cap\overline{\{0\}}.

The membership problem now can be seen as a word problem over 𝒜\mathcal{A}:

(1.5) Given nω and a well formed term τ over 𝒜, is {n}τ={n}?\displaystyle\text{Given $n\in\omega$ and a well formed term $\tau$ over $\mathcal{A}$, is $\{n\}\cap\tau=\{n\}$}?

It is natural to generalize the notion of arithmetic circuits by allowing input nodes to represent variable sets of numbers [5]. Logically speaking, we enhance our language by a set VV of variables which are interpreted as sets of natural numbers; arithmetic circuits correspond to the variable free terms of this language. It now makes sense to consider satisfiability and validity of (in–) equations of terms of this language under this interpretation. Furthermore, the operations f:(2ω)k2ωf:(2^{\omega})^{k}\rightarrow 2^{\omega} definable from the given operators 𝒪\mathcal{O} can be studied [16].

In analogy to the membership problem, Glaßer et al., [5] consider the complexity of

SC(𝒪)\displaystyle SC(\mathcal{O}) ={C(x0,,xn),k:C is an 𝒪 circuit and (k0,,kn)[kI(C(k0,,kn)]}\displaystyle=\{\langle C(x_{0},\ldots,x_{n}),k\rangle:C\text{ is an $\mathcal{O}$ circuit and }(\exists k_{0},\ldots,k_{n})[k\in I(C({k_{0}},\ldots,{k_{n}})]\}

for various sets 𝒪\mathcal{O} and determine many of these complexities. The main open problem is the question whether SC(,,,)SC(\cap,\cup,{}^{-},\bullet) is decidable. In other words, is it decidable whether the equation

(1.6) {k}τ(x0,,xn1)={k}\displaystyle\{k\}\cap\tau(x_{0},\ldots,x_{n-1})=\{k\}

has a solution over the subsets of ωn\omega^{n}?

In this paper we shall shed some light on these question and the structure of arithmetic circuits from an algebraic viewpoint. Our main tool will be the apparatus of Boolean algebras with operators, in particular, complex algebras of first order structures, which were introduced by Jónsson and Tarski, [9].

2 Notation and definitions

2.1 Algebras

An algebra 𝔄\mathfrak{A} is a pair 𝔄=A,𝒪\mathfrak{A}=\langle A,\mathcal{O}\rangle, where AA is a set and 𝒪={fi:iI}\mathcal{O}=\{f_{i}:i\in I\} a set of operation symbols ff each having a finite arity α(f)\alpha(f); if we write f(x0,,xn1)f(x_{0},\ldots,x_{n-1}) we implicitly assume that α(f)=n\alpha(f)=n. Operations of arity 0 are called (individual) constants. We will usually denote algebras by gothic letters 𝔄,𝔅,\mathfrak{A},\mathfrak{B},\ldots, and their universes by the corresponding roman letter A,B,A,B,\ldots. 𝔄\mathfrak{A} is called subdirectly irreducible if it has a smallest nontrivial congruence, and congruence–distributive if its congruence lattice is distributive.

Suppose that K\operatorname{{K}} is a class of algebras (of the same type 𝒪\mathcal{O}). For 𝔄,𝔅K\mathfrak{A},\mathfrak{B}\in K, 𝔄𝔅\mathfrak{A}\leq\mathfrak{B} means that 𝔄\mathfrak{A} is a subalgebra of 𝔅\mathfrak{B}. The operators 𝐈,𝐒,𝐇\mathbf{I},\mathbf{S},\mathbf{H} and 𝐏\mathbf{P} have their usual meaning. 𝐕𝐚𝐫(K)\mathbf{Var}(\operatorname{{K}}) is the variety generated by K\operatorname{{K}}, i.e. 𝐕𝐚𝐫(K)=𝐇𝐒𝐏(K)\mathbf{Var}(\operatorname{{K}})=\mathbf{H}\mathbf{S}\mathbf{P}(\operatorname{{K}}). A variety 𝐕\mathbf{V} is called finitely based if there is a finite set Σ\Sigma of equations in the language of 𝐕\mathbf{V} such that 𝔄𝐕\mathfrak{A}\in\mathbf{V} if and only if  𝔄Σ\mathfrak{A}\models\Sigma, and 𝐕\mathbf{V} is called finitely generated if there is a finite set K\operatorname{{K}} of finite algebras such that 𝐕=𝐕𝐚𝐫(K)\mathbf{V}=\mathbf{Var}(\operatorname{{K}}).

Suppose that K\operatorname{{K}} is a class of algebras of the same type 𝒪\mathcal{O}. We consider the following sets of formulas in the language of 𝒪{\mathcal{O}} (plus equality).

  1. (i)

    The first-order theory 𝐅𝐎K\mathbf{FO}~\operatorname{{K}} of K\operatorname{{K}}: The set of first-order formulas true in each member of K\operatorname{{K}}.

  2. (ii)

    The equational theory 𝐄𝐪K\mathbf{Eq}~\operatorname{{K}} of K\operatorname{{K}}: The set of formulas of the forms τ(x0,,xn)=σ(x0,,xn)\tau(x_{0},\ldots,x_{n})=\sigma(x_{0},\ldots,x_{n}) whose universal closures are true in each member of K\operatorname{{K}}.

  3. (iii)

    The satisfiable equations 𝐄𝐪𝐒𝐚𝐭K\mathbf{EqSat}~\operatorname{{K}} of K\operatorname{{K}}: The set of formulas of the forms τ(x0,,xn)=σ(x0,,xn)\tau(x_{0},\ldots,x_{n})=\sigma(x_{0},\ldots,x_{n}) whose existential closures are true in each member of K\operatorname{{K}}.

If K={𝔄}\operatorname{{K}}=\{\mathfrak{A}\}, we usually write 𝐅𝐎𝔄\mathbf{FO}~\mathfrak{A}, 𝐄𝐪𝔄\mathbf{Eq}~\mathfrak{A}, etc.

2.2 Boolean algebras with operators

In the following, let 𝔅=B,,,,,\mathfrak{B}=\langle B,{\lor,\land,{}^{-},\bot,\top}\rangle be a Boolean algebra (BA); here, \bot is the smallest and \top is the largest element of BB. If a,bBa,b\in B, then aba\vartriangle b denotes the symmetric difference (ab¯)(ba¯)(a\land\overline{b})\lor(b\land\overline{a}); note that a=ba=b if and only if  ab=a\vartriangle b=\bot. If 𝔅\mathfrak{B} is atomic, FC(𝔅)FC(\mathfrak{B}) is the finite–cofinite Boolean subalgebra of 𝔅\mathfrak{B}, i.e. every bFC(𝔅){,}b\in FC(\mathfrak{B})\setminus\{\bot,\top\} is a finite sum of atoms or the complement of such an element.

Suppose that ff is an n–ary operator on BB.

  1. (i)

    ff is called additive in its i–th argument, if

    f(a0,,ai1,x,ai+1,,an1)f(a0,,ai1,y,ai+1,,an1)=f(a0,,ai1,xy,ai+1,,an1).\displaystyle f(a_{0},\ldots,a_{i-1},x,a_{i+1},\ldots,a_{n-1})\lor f(a_{0},\ldots,a_{i-1},y,a_{i+1},\ldots,a_{n-1})=f(a_{0},\ldots,a_{i-1},x\lor y,a_{i+1},\ldots,a_{n-1}).
  2. (ii)

    ff is called normal in its i–th argument if f(a0,,ai1,,ai+1,,an1)=f(a_{0},\ldots,a_{i-1},\bot,a_{i+1},\ldots,a_{n-1})=\bot.

Note that an additive operator is isotone, i.e. it preserves the Boolean order in each of its arguments.

A Boolean algebra with operators (BAO) is a Boolean algebra with additional mappings of finitary rank that are additive and normal in each argument [9].

A (unary) discriminator function on 𝔅\mathfrak{B} is an operation dd on 𝔅\mathfrak{B} such that for all aBa\in B,

(2.1) d(a)={,if a=,,otherwise.\displaystyle d(a)=\begin{cases}\bot,&\text{if }a=\bot,\\ \top,&\text{otherwise}.\end{cases}

If 𝔅\mathfrak{B} has a discriminator function, we call 𝔅\mathfrak{B} a discriminator algebra.

For a class K\operatorname{{K}} of BAOs, a unary term tt is a discriminator term if it represents the discriminator function on each subdirectly irreducible member of K\operatorname{{K}}. A variety of BAOs is called a discriminator variety if it is generated by a class of algebras with a common discriminator term.

Having a discriminator function dd allows us to convert satisfiability (validity) of inequations into satisfiability (validity) of equations: Suppose that τ(x)\tau(\vec{x}) and σ(x)\sigma(\vec{x}) are terms with variables x\vec{x}. Then

(2.2) (x)[τ(x)σ(x)]\displaystyle(\exists\vec{x})[\tau(\vec{x})\neq\sigma(\vec{x})] (x)[τ(x)σ(x)](x)[d(τ(x)σ(x))=],\displaystyle\Longleftrightarrow(\exists\vec{x})[\tau(\vec{x})\vartriangle\sigma(\vec{x})\neq\bot]\Longleftrightarrow(\exists\vec{x})[d(\tau(\vec{x})\vartriangle\sigma(\vec{x}))=\top],
(2.3) (x)[τ(x)σ(x)]\displaystyle(\forall\vec{x})[\tau(\vec{x})\neq\sigma(\vec{x})] (x)[τ(x)σ(x)](x)[d(τ(x)σ(x))=].\displaystyle\Longleftrightarrow(\forall\vec{x})[\tau(\vec{x})\vartriangle\sigma(\vec{x})\neq\bot]\Longleftrightarrow(\forall\vec{x})[d(\tau(\vec{x})\vartriangle\sigma(\vec{x}))=\top].

If K\operatorname{{K}} is a class of algebras of the same type, we denote by Kd\operatorname{{K}}^{d} the class obtained from adding a unary operation symbol which represents the discriminator function on the members of K\operatorname{{K}}.

2.3 Complex algebras

Traditionally, a subset of a group GG is called a complex of GG; the power algebra of GG has 2G2^{G} as its universe, and the group operations lifted to 2G2^{G}. Complex algebras are a generalization of this situation and special instances of BAOs. Suppose that 𝔄,𝒪\langle\mathfrak{A},\mathcal{O}\rangle is an algebra, and f𝒪f\in\mathcal{O} is nn–ary. The complex operation 𝐟:(2A)n2A\mathbf{f}:(2^{A})^{n}\to 2^{A} corresponding to ff is defined by

(2.4) 𝐟(a0,,an1)={f(x0,,xn1):x0a0,,xn1an1}.\displaystyle\mathbf{f}(a_{0},\ldots,a_{n-1})=\{f(x_{0},\ldots,x_{n-1}):x_{0}\in a_{0},\ldots,x_{n-1}\in a_{n-1}\}.

The full complex algebra of 𝔄\mathfrak{A}, denoted by 𝔪𝔄\operatorname{{\mathfrak{Cm}}}\mathfrak{A}, has as its universe the powerset of AA and, besides the Boolean set operations, for each f𝒪f\in\mathcal{O} its complex operator 𝐟\mathbf{f} defined by (2.4).

More generally, the full complex algebra 𝔪𝒰\operatorname{{\mathfrak{Cm}}}\mathcal{U} of a relational structure U,\langle U,\mathcal{R}\rangle is the algebra 2U,,,,,U\langle 2^{U},\cup,\cap,{}^{-},\emptyset,U\rangle, which has for every RR\in\mathcal{R} of, say, arity n+1n+1, an nn – ary operator fR:(2U)n2Uf_{R}:(2^{U})^{n}\to 2^{U} defined by

(2.5) fR(X0,,Xn1)={yU:(x0,,xn1)[x0X0,,xn1Xn1 and R(y,x0,,xn1)]},\displaystyle f_{R}(X_{0},\ldots,X_{n-1})=\{y\in U:(\exists x_{0},\ldots,x_{n-1})[x_{0}\in X_{0},\ldots,x_{n-1}\in X_{n-1}\text{ and }R(y,x_{0},\ldots,x_{n-1})]\},

see e.g. [6].

Each subalgebra of 𝔪𝔄\operatorname{{\mathfrak{Cm}}}\mathfrak{A} is called a complex algebra of 𝔄\mathfrak{A}. Of particular interest for us are the subalgebra of 𝔪𝔄\operatorname{{\mathfrak{Cm}}}\mathfrak{A} generated by the constants, which we denote by 𝔪0𝔄\operatorname{{\mathfrak{Cm}}}_{0}\mathfrak{A}, and the subalgebra of 𝔪𝔄\operatorname{{\mathfrak{Cm}}}\mathfrak{A} generated by the singletons {a}\{a\}, where aAa\in A; we denote this algebra by 𝔪1𝔄\operatorname{{\mathfrak{Cm}}}_{1}\mathfrak{A}. Then, 𝔪0𝔄\operatorname{{\mathfrak{Cm}}}_{0}\mathfrak{A} is the smallest subalgebra of 𝔄\mathfrak{A} and 𝔪1𝔄\operatorname{{\mathfrak{Cm}}}_{1}\mathfrak{A} is the subalgebra of 𝔪𝔄\operatorname{{\mathfrak{Cm}}}\mathfrak{A} generated by the atoms. Clearly, 𝔪0𝔄𝔪1𝔄\operatorname{{\mathfrak{Cm}}}_{0}\mathfrak{A}\leq\operatorname{{\mathfrak{Cm}}}_{1}\mathfrak{A}, but the converse need not be true; an example will be given below.

2.4 Boolean monoids

The complex algebras of the various structures which we will consider have one or more commutative Boolean monoids as a reduct: A commutative Boolean monoid (CBM) is an algebra 𝔄=A,,,,,,,e\mathfrak{A}=\langle A,\lor,\land,{}^{-},\bot,\top,\circ,e\rangle such that

(2.6) A,,,,, is a Boolean algebra.\displaystyle\langle A,\lor,\land,{}^{-},\bot,\top\rangle\text{ is a Boolean algebra}.
(2.7) A,,e is a commutative monoid.\displaystyle\langle A,\circ,e\rangle\text{ is a commutative monoid.}
(2.8) x=.\displaystyle x\circ\bot=\bot.
(2.9) x(yz)=(xy)(xz).\displaystyle x\circ(y\lor z)=(x\circ y)\lor(x\circ z).

In the sequel, we let c(x)=xc(x)=x\circ\top; it is well known that cc is an additive closure operator on CBMs [8]. Furthermore [see e.g. 17],

Lemma 2.1.
  1. (i)

    The class CBM is congruence distributive.

  2. (ii)

    II is a congruence ideal – i.e. the kernel of a congruence – on a CBM 𝔄\mathfrak{A} if and only if  II is a Boolean ideal and xIx\in I implies c(x)Ic(x)\in I for all xAx\in A.

  3. (iii)

    The principal (Boolean) ideal generated by c(x)c(x) is the smallest congruence ideal containing xx.

An element xAx\in A is called a congruence element if c(x)=xc(x)=x. By Lemma 2.1(3), each principal congruence ideal II of 𝔄\mathfrak{A} is of the form I={y:yx}I=\{y:y\leq x\} for some congruence element xx. Note that a CBM is simple – i.e. has only two congruences – if and only if  it satisfies

(2.10) (x)[x=c(x)=].\displaystyle(\forall x)[x=\bot\lor c(x)=\top].

3 Complex algebras of \mathbb{N}

Let =ω,0,+,,1\mathbb{N}=\langle\omega,0,+,\cdot,1\rangle be the semiring of natural numbers, and 𝔪=2ω,,,,,ω,{0},+,{1},\operatorname{{\mathfrak{Cm}}}~\mathbb{N}=\langle 2^{\omega},\cap,\cup,{}^{-},\emptyset,\omega,\{0\},\bm{+},\{1\},\bullet\rangle be its full complex algebra, i.e.

a+b\displaystyle a\bm{+}b ={n+m:na,mb},\displaystyle=\{n+m:n\in a,\ m\in b\},
ab\displaystyle a\bullet b ={nm:na,mb}.\displaystyle=\{n\cdot m:n\in a,\ m\in b\}.

A function F:(2ω)n2ωF:(2^{\omega})^{n}\to 2^{\omega} is called circuit definable if there is a term τ(v0,,vn1)\tau(v_{0},\ldots,v_{n-1}) in the language of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} such that F(s0,,sn1)=τ(s0/v0,,sn1/vn1)F(s_{0},\ldots,s_{n-1})=\tau(s_{0}/v_{0},\ldots,s_{n-1}/v_{n-1}) for all s0,,sn1ωs_{0},\ldots,s_{n-1}\subseteq\omega. A subset aa of ω\omega is called circuit definable, if there is a closed (i.e. variable free) term τ\tau that evaluates to aa. Each element of the smallest subalgebra 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} corresponds to an arithmetic circuit with finite input nodes and vice versa via the interpretation II.

Both 2ω,+,{0}\langle 2^{\omega},\bm{+},\{0\}\rangle and 2ω,,{1}\langle 2^{\omega},\bullet,\{1\}\rangle are commutative monoids. Furthermore, +\bm{+} and \bullet are normal and (completely) additive operators with respect to \cup, so that 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} is a Boolean algebra with operators, and

2ω,,,,,ω,+,{0},2ω,,,,,ω,,{1}\langle 2^{\omega},\cup,\cap,{}^{-},\emptyset,\omega,\bm{+},\{0\}\rangle,\quad\langle 2^{\omega},\cup,\cap,{}^{-},\emptyset,\omega,\bullet,\{1\}\rangle

are CBMs.

Theorem 3.1.
  1. (i)

    𝔪\operatorname{{\mathfrak{Cm}}} \mathbb{N} is a discriminator algebra.

  2. (ii)

    𝔪0=𝔪1\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}=\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}.

  3. (iii)

    𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} is embeddable into any simple algebra of 𝐕𝐚𝐫(𝔪)\mathbf{Var}(\operatorname{{\mathfrak{Cm}}}~\mathbb{N}).

Proof.

(i)  Let f(x)f(x) be the function ω+({0}x)\omega\bm{+}(\{0\}\bullet x). If x=x=\emptyset, then {0}x=\{0\}\bullet x=\emptyset, and thus, f(x)=ω+=f(x)=\omega\bm{+}\emptyset=\emptyset. If xx\neq\emptyset, then {0}x={0}\{0\}\bullet x=\{0\}, hence f(x)=ω+{0}=ωf(x)=\omega\bm{+}\{0\}=\omega.

(ii)  The atoms of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} are the singletons {n}\{n\}, and {n}={1}+{1}n times\{n\}=\underbrace{\{1\}\bm{+}\ldots\{1\}}_{n\text{ times}} if n>0n>0.

(iii) Since 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} is a discriminator algebra, it suffices to show that the smallest subalgebra 𝔄\mathfrak{A} of an ultrapower of copies of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} is isomorphic to 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}. Thus, let 𝔅:=𝔪κ/U\mathfrak{B}:={}^{\kappa}\operatorname{{\mathfrak{Cm}}}~\mathbb{N}/U be an ultrapower of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}. Suppose that e:𝔪𝔅e:\operatorname{{\mathfrak{Cm}}}~\mathbb{N}\to\mathfrak{B} is the canonical embedding, i.e. e(a)=fa/Ue(a)=f_{a}/U, where fa(i)=af_{a}(i)=a for all i<κi<\kappa. Since 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} is generated by {0}\{0\}, e[𝔪0]e[\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}] is generated by e({0})e(\{0\}), and thus, since ee is an embedding, e[𝔪0]e[\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}] is the smallest subalgebra of 𝔅\mathfrak{B}. ∎

Theorem 3.2.

The Boolean reduct of 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} has 2ω2^{\omega} ultrafilters.

Proof.

Let p0,,pk,q0,,qkp_{0},\ldots,p_{k},q_{0},\ldots,q_{k} be different primes; then

p0pk(ωp0)(ωpk)(ω{q0}¯)(ω{qk}¯).\displaystyle p_{0}\cdot\ldots\cdot p_{k}\in(\omega\bullet{p_{0}})\land\ldots\land(\omega\bullet{p_{k}})\land(\overline{\omega\bullet\{q_{0}\}})\land\ldots\land(\overline{\omega\bullet\{q_{k}\}}).

Hence, {ω{p}:p prime}\{\omega\bullet\{p\}:p\text{ prime}\} is an independent set which generates an atomless Boolean subalgebra 𝔄\mathfrak{A} of 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}. 𝔄\mathfrak{A} has 2ω2^{\omega} ultrafilters, and thus, so has 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}. ∎

The atom structure 𝔄𝔱𝔪\operatorname{\mathfrak{At}}\operatorname{{\mathfrak{Cm}}}~\mathbb{N} of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N} has the set Ω={{n}:nω}\Omega=\{\{n\}:n\in\omega\} as its universe, and for each n – ary operator ff an n+1–ary relation Rf:={p,q:pΩn and qΩ,qf(p)}R_{f}:=\{\langle p,q\rangle:p\in\Omega^{n}\text{ and }q\in\Omega,q\subseteq f(p)\}. Then,

R+({k},{n},{m})\displaystyle R_{\bm{+}}(\{k\},\{n\},\{m\}) {m}{k}+{n}k+n=m,\displaystyle\Longleftrightarrow\{m\}\subseteq\{k\}\bm{+}\{n\}\Longleftrightarrow k+n=m,
R({k},{n},{m})\displaystyle R_{\bullet}(\{k\},\{n\},\{m\}) {m}{k}{n}kn=m.\displaystyle\Longleftrightarrow\{m\}\subseteq\{k\}\bullet\{n\}\Longleftrightarrow k\cdot n=m.

It is well known that 𝔄𝔱𝔪\operatorname{\mathfrak{At}}\operatorname{{\mathfrak{Cm}}}~\mathbb{N}\cong\mathbb{N}. Let us call a relation on 𝔄𝔱𝔪\operatorname{\mathfrak{At}}\operatorname{{\mathfrak{Cm}}}~\mathbb{N}, i.e. on \mathbb{N}, circuit definable if it corresponds to a circuit definable operator on 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}. A striking example of the lack of expressiveness of arithmetic circuits is the following:

Theorem 3.3.
  1. (i)

    In \mathbb{N}, the converse \geq of the natural ordering is circuit definable, while \leq is not.

  2. (ii)

    Relative subtraction is not circuit definable.

Proof.

Using (2.5) it is easily seen that \geq is the relation corresponding to the function defined by f(x)=x+ωf(x)=x\bm{+}\omega. The ordering \leq on ω\omega corresponds to the function defined by f(x)={nω:(m)[mx and nm}f(x)=\{n\in\omega:(\exists m)[m\in x\text{ and }n\leq m\}, and we have shown in [16] that this function is not circuit definable. In the same paper we have proved (ii). ∎

3.1 Complex algebras of ω,+,0\langle\omega,+,0\rangle

Let +=ω,+,0,\mathbb{N}^{\bm{+}}=\langle\omega,+,0,\rangle, 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} be its full complex algebra, and 𝐕\mathbf{V} be the variety generated by 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}. Furthermore, set c(x)=x+ωc(x)=x\bm{+}\omega. Recall that the constant {1}\{1\} is definable in 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} by

{1}=ω((ω{0}+ω{0}){0}).\displaystyle\{1\}=\omega\setminus((\omega\setminus\{0\}\bm{+}\omega\setminus\{0\})\cup\{0\}).

Note that for all aωa\subseteq\omega,

(3.1) c(a)=a+ω={k:(n,m)[ma and k=m+n]}={k:min(a)k}=c({min(a)}),\displaystyle c(a)=a\bm{+}\omega=\{k:(\exists n,m)[m\in a\text{ and }k=m+n]\}=\{k:\min(a)\leq k\}=c(\{\min(a)\}),
(3.2) c(a)+{1}=a+c({1})=a+{0}¯.\displaystyle c(a)+\{1\}=a\bm{+}c(\{1\})=a\bm{+}\overline{\{0\}}.

The following observation will be useful:

Lemma 3.4.

Let a,bωa,b\subseteq\omega. Then, a=a=\emptyset or b=b=\emptyset if and only if  c(a)c(b)=c(a)\cap c(b)=\emptyset.

Proof.

If, say, a=a=\emptyset, then c(a)=c(a)=\emptyset. Conversely, if c(a)c(b)=c(a)\cap c(b)=\emptyset, then one of c(a)c(a) or c(b)c(b) must be empty, since the intersection of any two cofinite sets is not empty. Hence, a=a=\emptyset or b=b=\emptyset. ∎

Recall that 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}} is the smallest subalgebra of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}. The following result is well known:

Lemma 3.5.

The universe of 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}} is the finite – cofinite subalgebra of 2ω2^{\omega}.

Next, we describe the congruences of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}:

Theorem 3.6.

The congruences of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} form a chain of order type 1+ω1+\omega^{*}.

Proof.

By Lemma 2.1, c({n})c(\{n\}) is a congruence element generating the congruence θn\theta_{n}. Conversely, suppose that \equiv is a congruence induced by the non–trivial ideal II; then, II\neq\emptyset, and II is closed under cc. Since II is also closed under \subseteq, {min(a)}I\{\min(a)\}\in I for every aIa\in I, and therefore, n:=min({min(a):aI,a})n:=\min(\{\min(a):a\in I,\ a\neq\emptyset\}) exists, and c({n})Ic(\{n\})\in I. If aI,aa\in I,\ a\neq\emptyset, then nmin(a)n\leq\min(a), and it follows that ac(a)=c({min(a)})c({n})a\subseteq c(a)=c(\{\min(a)\})\subseteq c(\{n\}). Hence, II is the principal ideal of 2ω2^{\omega} generated by c({n})c(\{n\}).

Observing that c({n})={m:nm}c(\{n\})=\{m:n\leq m\}, we see that

c({n+1})c({n})c({1})c({0})=ω,\displaystyle\emptyset\subsetneq\ \ldots\subsetneq c(\{n+1\})\subsetneq c(\{n\})\subsetneq\ldots\subsetneq c(\{1\})\subsetneq c(\{0\})=\omega,

and thus,

(3.3) 1θn+1θnθ1θ0=V,\displaystyle 1^{\prime}\subsetneq\ldots\subsetneq\theta_{n+1}\subsetneq\theta_{n}\subsetneq\ldots\subsetneq\theta_{1}\subsetneq\theta_{0}=V,

where 11^{\prime} is the identity and VV the universal congruence. Clearly, this chain has order type 1+ω1+\omega^{*}. It follows that 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+} has no smallest nontrivial congruence, and therefore, 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+} is not subdirectly irreducible. ∎

Corollary 3.7.

The congruences of 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}} form a chain of order type 1+ω1+\omega^{*}.

Proof.

Each congruence θn\theta_{n} of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} is generated by a cofinite congruence element, which is in 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}} by Lemma 3.5. ∎

Let 𝔅n:=𝔪+/θn+1\mathfrak{B}_{n}:=\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}/\theta_{n+1}, and πn:𝔪+𝔅n\pi_{n}:\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}\twoheadrightarrow\mathfrak{B}_{n} be the quotient mapping. Note that the kernel of θn+1\theta_{n+1} is the ideal of 2ω2^{\omega} generated by c({n+1})={n+1}+ω=ω[0,n]c(\{n+1\})=\{n+1\}\bm{+}\omega=\omega\setminus[0,n]. Thus, the Boolean part of 𝔅n\mathfrak{B}_{n} is isomorphic to the powerset algebra of {0,,n}\{0,\ldots,n\} with atoms gi:=πn({i})g_{i}:=\pi_{n}(\{i\}) for ini\leq n. In particular, 𝔅0\mathfrak{B}_{0} is isomorphic to the two element Boolean algebra, since c({1})=ω{0}c(\{1\})=\omega\setminus\{0\} generates a prime ideal of 2ω2^{\omega}.

The composition table for \circ on the atoms of 𝔅n\mathfrak{B}_{n} is given below. Observe that g0=πn({0})g_{0}=\pi_{n}(\{0\}) is the identity element ee of 𝔅n,\langle\mathfrak{B}_{n},\circ\rangle, and gm=g0g1g1m – timesg_{m}=g_{0}\circ\underbrace{g_{1}\circ\ldots\circ g_{1}}_{\text{m -- times}}.

g0g1g2gn1gng0g0g1g2gn1gng1g1g2g3gng2g2g3g4gngn\begin{array}[]{c|cccccc}\circ&{g_{0}}&{g_{1}}&g_{2}&\ldots&g_{n-1}&g_{n}\\ \hline\cr g_{0}&g_{0}&{g_{1}}&g_{2}&\ldots&g_{n-1}&g_{n}\\ g_{1}&g_{1}&g_{2}&g_{3}&\ldots&g_{n}&\bot\\ g_{2}&g_{2}&g_{3}&g_{4}&\ldots&\bot&\bot\\ \ldots\\ g_{n}&g_{n}&\bot&\bot&\ldots&\bot&\bot\end{array}
Theorem 3.8.
  1. (i)

    Each 𝔅n\mathfrak{B}_{n} is subdirectly irreducible.

  2. (ii)

    𝐕𝐚𝐫(𝔅n)𝐕𝐚𝐫(𝔅n+1)\mathbf{Var}(\mathfrak{B}_{n})\subsetneq\mathbf{Var}(\mathfrak{B}_{n+1}).

  3. (iii)

    𝐕=𝐕𝐚𝐫{𝔅n:nω}\mathbf{V}=\mathbf{Var}\{\mathfrak{B}_{n}:n\in\omega\}, and thus, 𝐕\mathbf{V} is generated by its finite members.

Proof.

(i)  The congruences of 𝔅n\mathfrak{B}_{n} are in 1–1 correspondence to the congruences of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} containing θn\theta_{n}. This is a finite chain, and the smallest nonzero congruence element of 𝔅n\mathfrak{B}_{n} is gng_{n}.

(ii)  Clearly, 𝐕𝐚𝐫(𝔅n)𝐕𝐚𝐫(𝔅n+1)\mathbf{Var}(\mathfrak{B}_{n})\subseteq\mathbf{Var}(\mathfrak{B}_{n+1}). In 𝔅n\mathfrak{B}_{n}, g1++g1n+1 times=\underbrace{g_{1}\bm{+}\ldots\bm{+}g_{1}}_{n+1\text{ times}}=\bot, and g1++g1n+1 times=gn+1\underbrace{g_{1}\bm{+}\ldots\bm{+}g_{1}}_{n+1\text{ times}}=g_{n+1}\neq\bot in 𝔅n+1\mathfrak{B}_{n+1}.

(iii)  Clearly, 𝔅n𝐕\mathfrak{B}_{n}\in\mathbf{V} for each nωn\in\omega. Conversely, by Birkhoff’s subdirect representation theorem [3], 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} is isomorphic to a subdirect product of its subdirectly irreducible quotients, see e.g. [4], Theorem 8.6. By Theorem 3.6, the only proper quotients of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} are the algebras 𝔅n\mathfrak{B}_{n}, and these are subdirectly irreducible by 1. above. ∎

𝐕\mathbf{V} contains all Boolean algebras for which the extra operator \circ is the Boolean meet and e=e=\top, since the universe of 𝔅0\mathfrak{B}_{0} is the two element Boolean algebra, and 𝔅0𝐕\mathfrak{B}_{0}\in\mathbf{V}. Moreover,

Theorem 3.9.

𝐕𝐚𝐫(𝔅n)\mathbf{Var}(\mathfrak{B}_{n}) is finitely based for each nωn\in\omega. Hence, 𝐄𝐪𝔅n\mathbf{Eq}~\mathfrak{B}_{n} is decidable for all nωn\in\omega.

Proof.

Since 𝐕𝐚𝐫(𝔅n)\mathbf{Var}(\mathfrak{B}_{n}) is congruence distributive and 𝔅n\mathfrak{B}_{n} is finite, Baker’s finite basis theorem [1] implies that 𝐕𝐚𝐫(𝔅n)\mathbf{Var}(\mathfrak{B}_{n}) is finitely based for each nωn\in\omega. The second claim follows from the fact that a finitely based variety which is generated by a finite algebra has a decidable equational theory. ∎

Corollary 3.10.

𝐄𝐪𝐕\mathbf{Eq}~\mathbf{V} is co – r.e.

Proof.

Given an equation τ=σ\tau=\sigma we can check whether τ=σ\tau=\sigma holds in 𝔅0,𝔅1,,\mathfrak{B}_{0},\mathfrak{B}_{1},\ldots,, since 𝐄𝐪Bn\mathbf{Eq}~B_{n} is decidable. Since 𝐕\mathbf{V} is generated by {𝔅n:nω}\{\mathfrak{B}_{n}:n\in\omega\}, any equation that fails in 𝐕\mathbf{V} must fail in some 𝔅n\mathfrak{B}_{n}. ∎

Let gg be the term

(3.4) g:=e¯e¯¯e¯.\displaystyle g:=\overline{\overline{e}\circ\overline{e}}\land\overline{e}.

In 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}, gg evaluates to {1}\{1\}. Furthermore, we set

gn:={e,if n=0,gggntimes,otherwise.\displaystyle g^{n}:=\begin{cases}e,&\text{if }n=0,\\ \underbrace{g\circ g\circ\ldots\circ g}_{n-\text{times}},&\text{otherwise.}\end{cases}

Consider the following identities in the language of 𝐕\mathbf{V}:

(3.5) e(xy)=exy.\displaystyle e\land(x\circ y)=e\land x\land y.
(3.6) c(gn+1)=g0gn¯ for all nω.\displaystyle c(g^{n+1})=\overline{g^{0}\lor\ldots\lor g^{n}}\text{ for all }n\in\omega.
(3.7) c[c(x)c(y)¯]c[c(y)c(x)¯]=.\displaystyle c[c(x)\land\overline{c(y)}]\land c[c(y)\land\overline{c(x)}]=\bot.
(3.8) g(xy)=[(ex)(gy)][(gx)(ey)]\displaystyle g\land(x\circ y)=[(e\land x)\circ(g\land y)]\lor[(g\land x)\circ(e\land y)]
(3.9) (xgn)(x¯gn)= for all nω.\displaystyle(x\land g^{n})\circ(\overline{x}\land g^{n})=\bot\text{ for all }n\in\omega.
(3.10) c(x)=c(xxe¯¯).\displaystyle c(x)=c(x\land\overline{x\circ\overline{e}}).
Lemma 3.11.

(3.5) – (3.10) hold in 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}, and thus, in 𝐕\mathbf{V}.

Proof.

(3.5):  Just note that 0a+b0a and 0b0\in a\bm{+}b\Longleftrightarrow 0\in a\text{ and }0\in b so that {0}(a+b)\{0\}\cap(a\bm{+}b)\neq\emptyset if and only if  {0}a\{0\}\cap a\neq\emptyset and {0}b\{0\}\cap b\neq\emptyset.

(3.6):  c({n+1})={n+1}+ω={n+1}={0,,n}¯c(\{n+1\})=\{n+1\}\bm{+}\omega=\uparrow\{n+1\}=\overline{\{0,\ldots,n\}}.

(3.7):  The set {c(a):aω}\{c(a):a\subseteq\omega\} is a chain, thus, c(x)c(y)¯=c(x)\cap\overline{c(y)}=\emptyset or c(y)c(x)¯=c(y)\cap\overline{c(x)}=\emptyset; hence, c(c(x)c(y)¯)=c(c(x)\cap\overline{c(y)})=\emptyset or c(c(y)c(x)¯)=c(c(y)\cap\overline{c(x)})=\emptyset. Now apply Lemma 3.4.

(3.8):  This follows immediately from the definition of +\bm{+}.

(3.9):  Each gng^{n} is an atom of 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}, so xgn=x\land g^{n}=\bot or x¯gn=\overline{x}\land g^{n}=\bot for all x𝔪0+x\in\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}.

(3.10):  If aωa\subseteq\omega and a=a=\emptyset, the claim clearly holds. If aa\neq\emptyset, then aa+{0}¯¯=min(a)a\cap\overline{a\bm{+}\overline{\{0\}}}=\min(a) whence the conclusion follows. ∎

We do not know whether (3.5) – (3.10) are sufficient to axiomatize 𝐕\mathbf{V}.

Theorem 3.12.

Let 𝔄𝐕\mathfrak{A}\in\mathbf{V} be subdirectly irreducible and suppose that dd is the smallest nonzero congruence element in 𝔄\mathfrak{A}.

  1. (i)

    ee is an atom of AA.

  2. (ii)

    The congruence elements of 𝔄\mathfrak{A} are linearly ordered.

  3. (iii)

    If 𝔄\mathfrak{A} is finite, then it is isomorphic to some 𝔅n\mathfrak{B}_{n}.

Proof.

(i)   Assume that there are a,bAa,b\in A such that a,b\bot\lneq a,b, ab=a\land b=\bot, and ab=ea\lor b=e. Then, the monotonicity of \circ implies that abee=ea\circ b\leq e\circ e=e, and by (3.5), ab=(ab)e=abe=a\circ b=(a\circ b)\land e=a\land b\land e=\bot.

Since aa\neq\bot, we have aa\circ\top\neq\bot, and the fact that dd is the smallest non–zero congruence element implies dad\ \leq a\circ\top. Now,

dadbab==,\displaystyle d\leq a\circ\top\Rightarrow d\circ b\leq a\circ b\circ\top=\bot\circ\top=\bot,

and, similarly, da=d\circ a=\bot. But then,

d=de=d(ab)=(da)(db)=,\displaystyle d=d\circ e=d\circ(a\lor b)=(d\circ a)\lor(d\circ b)=\bot,

contradicting our hypothesis that dd\neq\bot.

(ii)  Assume there are nonzero congruence elements x,yx,y such that xy¯x\land\overline{y}\neq\bot and yx¯y\land\overline{x}\neq\bot. Then, both c(xy¯)c(x\land\overline{y}) and c(yx¯)c(y\land\overline{x}) are nonzero congruence elements, and therefore, dc(xy¯)c(yx¯)d\leq c(x\land\overline{y})\land c(y\land\overline{x}). On the other hand,

c(xy¯)c(yx¯)=c(c(x)c(y)¯)c(c(y)c(x)¯)=\displaystyle c(x\land\overline{y})\land c(y\land\overline{x})=c(c(x)\land\overline{c(y)})\land c(c(y)\land\overline{c(x)})=\bot

by (3.7) which contradicts dd\neq\bot.

(iii)  By (3.6), mnm\neq n implies that gmgn=g^{m}\land g^{n}=\bot. Therefore, since 𝔄\mathfrak{A} is finite, there exists a smallest nn such that gn+1=g^{n+1}=\bot. We will prove that 𝔄=𝔅n\mathfrak{A}=\mathfrak{B}_{n}.

  1. 1.

    c(gn)=gnc(g^{n})=g^{n}: Again by (3.6) we have g0gn1c(gn)=g^{0}\lor\ldots g^{n-1}\lor c(g^{n})=\top, and c(gn)gm=c(g^{n})\land g^{m}=\bot for all mnm\lneq n. Suppose there is some sAs\in A such that sgn=s\land g^{n}=\bot and sgn=c(gn)s\lor g^{n}=c(g^{n}). Then,

    gs=(gs)=(ggn)(gs)=g(gns)=gc(gn)=g(gn)=(ggn)=,\displaystyle g\circ s=\bot\lor(g\circ s)=(g\circ g^{n})\lor(g\circ s)=g\circ(g^{n}\lor s)=g\circ c(g^{n})=g\circ(g^{n}\circ\top)=(g\circ g^{n})\circ\top=\bot,

    and, by the normality of \circ we obtain gms=g^{m}\circ s=\bot for all 1mn1\leq m\leq n. Now,

    c(gn)=gn(g0gns)=(gng0)gng1=gngn=gns==gn.\displaystyle c(g^{n})=g^{n}\circ(g^{0}\lor\ldots\lor g^{n}\lor s)=(g^{n}\circ g^{0})\lor\underbrace{g^{n}\circ g^{1}}_{=\bot}\lor\ldots\lor\underbrace{g^{n}\circ g^{n}}_{=\bot}\lor\underbrace{g^{n}\circ s}_{=\bot}=g^{n}.

    It follows that s=s=\bot and also that g0gn=g^{0}\lor\ldots\lor g^{n}=\top.

  2. 2.

    d=gnd=g^{n}:  Since dd is the smallest congruence element, we have dgnd\leq g^{n}. Assume there is some tt\neq\bot such that dt=d\land t=\bot and dt=gnd\lor t=g^{n}. Then, for x{d,t}x\in\{d,t\} and y{g1,gn}y\in\{g^{1},\ldots g^{n}\} we have xy=x\circ y=\bot. Furthermore, dd=dt=tt=d\circ d=d\circ t=t\circ t=\bot, since d,tgnd,t\leq g^{n} and gn+1=g^{n+1}=\bot. This implies that dd and tt are disjoint nonzero congruence elements, contradicting the subdirect irreducibility of 𝔄\mathfrak{A}. It follows that d=gnd=g^{n}.

  3. 3.

    Each gmg^{m} is an atom of 𝔄\mathfrak{A}: Assume that there are s,tgm\bot\lneq s,t\lneq g^{m} with st=s\land t=\bot and st=gms\lor t=g^{m} for some mnm\leq n. By (i) above, we have 1m1\leq m. From sgms\leq g^{m} it follows that sgkgmgk=gk+mgms\circ g^{k}\leq g^{m}\circ g^{k}=g^{k+m}\neq g^{m} for k0k\neq 0. Therefore,

    gm(s)=gm(s(sg1)(sgn))=s.\displaystyle g^{m}\land(s\circ\top)=g^{m}\land(s\lor(s\circ g^{1})\lor\ldots\lor(s\circ g^{n}))=s.

    Similarly we obtain gm(t)=tg^{m}\land(t\circ\top)=t. Since tt and ss are nonzero and disjoint, ss\circ\top and tt\circ\top are incomparable congruence elements, contradicting (ii).

Theorem 3.13.

𝔄𝐕\mathfrak{A}\in\mathbf{V} is simple if and only if  |A|2\lvert A\rvert\leq 2.

Proof.

Clearly, 𝔄\mathfrak{A} is simple if it has at most two elements. Conversely, let 𝔄\mathfrak{A} be simple. If gg\neq\bot, then c(g)=c(g)=\top by (2.10), and thus, =c(g)¯=e\bot=\overline{c(g)}=e by (3.6). The normality of \circ implies that, for all xAx\in A, x=ex=x=x=e\circ x=\bot\circ x=\bot, and therefore, 𝔄\mathfrak{A} has only one element.

Now, suppose that g=g=\bot; then, =c(g)=e¯\bot=c(g)=\overline{e} by (3.6), and thus, e=e=\top. If xx\neq\bot, then

x=xe=x=c(x)=,\displaystyle x=x\circ e=x\circ\top=c(x)=\top,

the latter by the simplicity of 𝔄\mathfrak{A}. ∎

Since every nontrivial variety contains a nontrivial simple algebra, it follows that the subvariety 𝐕0\mathbf{V}_{0} of 𝐕\mathbf{V} generated by 𝔅0\mathfrak{B}_{0} is smallest nontrivial subvariety of 𝐕\mathbf{V}.

If 𝔄\mathfrak{A} is a CBM, we call zAz\in A an annihilator of \circ, if xz=zx\circ z=z for all xA,xx\in A,\ x\neq\bot. The complex algebra of ω,1,\langle\omega,1,\cdot\rangle has {0}\{0\} as a nonzero annihilator. This cannot happen in 𝐕\mathbf{V}:

Theorem 3.14.

Suppose that 𝔄𝐕\mathfrak{A}\in\mathbf{V} and that |A|>2\lvert A\rvert>2. Then, 𝔄\mathfrak{A} has no nonzero annihilator.

Proof.

Since 𝐕=𝐇𝐒𝐏{𝔅n:nω}\mathbf{V}=\mathbf{H}\mathbf{S}\mathbf{P}\{\mathfrak{B}_{n}:n\in\omega\}, there are a sequence {α:α<κ}\{\mathfrak{C}_{\alpha}:\alpha<\kappa\} of algebras from {𝔅n:nω}\{\mathfrak{B}_{n}:n\in\omega\}, a subalgebra 𝔇\mathfrak{D} of :=α<κα\mathfrak{C}:=\prod_{\alpha<\kappa}\mathfrak{C}_{\alpha}, and an onto homomorphism π:𝔇𝔄\pi:\mathfrak{D}\twoheadrightarrow\mathfrak{A} with kernel II. Let g=e¯e¯¯e¯g=\overline{\overline{e}\circ\overline{e}}\land\overline{e} in \mathfrak{C}, and gα=e¯e¯¯e¯g_{\alpha}=\overline{\overline{e}\circ\overline{e}}\land\overline{e} in α\mathfrak{C}_{\alpha}. Since 𝔇\mathfrak{D} is a subalgebra of \mathfrak{C} and gg is a constant term, we have g𝔇g\in\mathfrak{D}; furthermore, g(α)=gαg(\alpha)=g_{\alpha} for all α<κ\alpha<\kappa.

Assume that zz is a nonzero annihilator of 𝔄\mathfrak{A}, and let f𝔇f\in\mathfrak{D} with z=π(f)z=\pi(f); since zz\neq\bot we have fIf\not\in I, in particular, ff\neq\bot. Now, z=z=π(f)π()=π(f)z=z\circ\top=\pi(f)\circ\pi(\top)=\pi(f\circ\top), and we may suppose that ff is a congruence element. Since 𝔄\mathfrak{A} has more than two elements, <g𝔄\bot<g_{\mathfrak{A}}, and therefore zg𝔄=zz\circ g_{\mathfrak{A}}=z. Hence, there is some iIi\in I such that (fg)i=fi(f\circ g)\lor i=f\lor i, in particular, f(fg)if\leq(f\circ g)\lor i; since II is a congruence ideal, we may suppose w.l.o.g. that i=c(i)i=c(i).

Let α<κ\alpha<\kappa such that f(α)f(\alpha)\neq\bot, and suppose that α=𝔅n\mathfrak{C}_{\alpha}=\mathfrak{B}_{n}; then, f(α)(f(α)gα)i(α)f(\alpha)\leq(f(\alpha)\circ g_{\alpha})\lor i(\alpha). Since ff is a congruence element, so is f(α)f(\alpha), and it follows from the definition of 𝔅n\mathfrak{B}_{n} that there is some m<nm<n such that f(α)=c(gαm)f(\alpha)=c(g_{\alpha}^{m}). Now,

f(α)\displaystyle f(\alpha) (f(α)g(α))i(α)\displaystyle\leq(f(\alpha)\circ g(\alpha))\lor i(\alpha)
=(c(gαm)gα)i(α)\displaystyle=(c(g_{\alpha}^{m})\circ g_{\alpha})\lor i(\alpha)
=(gαmgα)i(α)\displaystyle=(g_{\alpha}^{m}\circ\top\circ g_{\alpha})\lor i(\alpha)
=c(gαm+1)i(α)\displaystyle=c(g_{\alpha}^{m+1})\lor i(\alpha)
=(3.6)gα0gαm¯i(α)\displaystyle\overset{\eqref{ge1}}{=}\overline{g_{\alpha}^{0}\lor\ldots\lor g_{\alpha}^{m}}\lor i(\alpha)
=(gα0gαm1¯gαm¯)i(α)\displaystyle=(\overline{g_{\alpha}^{0}\lor\ldots\lor g_{\alpha}^{m-1}}\land\overline{g_{\alpha}^{m}})\lor i(\alpha)
=(f(α)gαm¯)i(α),\displaystyle=(f(\alpha)\land\overline{g_{\alpha}^{m}})\lor i(\alpha),
which implies
f(α)gαm\displaystyle f(\alpha)\land g_{\alpha}^{m} i(α).\displaystyle\leq i(\alpha).

Now, f(α)=gα0gαm1¯f(\alpha)=\overline{g_{\alpha}^{0}\lor\ldots\lor g_{\alpha}^{m-1}} implies gαmf(α)g_{\alpha}^{m}\leq f(\alpha), and thus, gαmi(α)g_{\alpha}^{m}\leq i(\alpha). Since i(α)i(\alpha) is a congruence element, we have i=ii=i\circ\top, and therefore,

f(α)=c(gαm)=gαmi(α)=i(α).\displaystyle f(\alpha)=c(g_{\alpha}^{m})=g_{\alpha}^{m}\circ\top\leq i(\alpha)\circ\top=i(\alpha).

Thus, f(α)i(α)f(\alpha)\leq i(\alpha) for all α<κ\alpha<\kappa and it follows that fIf\in I, contradicting our hypothesis. ∎

Let us briefly look at the complex algebra 𝔪+,\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+,\leq} of ω,+,,0\langle\omega,+,\leq,0\rangle. We have seen earlier that the complex version of \leq is the operator :2ω2ω\downarrow:2^{\omega}\to 2^{\omega} defined by a={nω:(m)[ma and nm]}\downarrow~a=\{n\in\omega:(\exists m)[m\in a\text{ and }n\leq m]\}; thus, the universe of 𝔪0+,\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{+,\leq} is FC(ω)FC(\omega).

Since \leq is first order definable in ω,+,0\langle\omega,+,0\rangle, one might suspect that 𝔪+,\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+,\leq} and 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+} are not “too far apart”. It turns out, however that 𝔪+,\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+,\leq} has much stronger properties than 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+}.

Theorem 3.15.
  1. (i)

    𝔪+,\operatorname{{\mathfrak{Cm}}}\mathbb{N}^{+,\leq} is a discriminator algebra.

  2. (ii)

    𝐄𝐪𝔪+,𝐄𝐪𝔪0+,\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}\mathbb{N}^{+,\leq}\neq\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{+,\leq}.

Proof.

(i)  Set d(x):=ω+xd(x):=\omega\bm{+}\downarrow x. If x=x=\emptyset, then x=\downarrow x=\emptyset, and thus, d(x)=d(x)=\emptyset. Otherwise, 0x0\in\downarrow x, hence, d(x)=ω+x=ωd(x)=\omega\bm{+}x=\omega.

(ii)  Consider the function 𝖿𝗂𝗇:2ω{,ω}\mathsf{fin}:2^{\omega}\to\{\emptyset,\omega\} defined by 𝖿𝗂𝗇(a):=d(a¯)¯\mathsf{fin}(a):=\overline{d(\overline{\downarrow a})}. Then,

𝖿𝗂𝗇(a)={ω,if |a|=ω,,if a is finite.\displaystyle\mathsf{fin}(a)=\begin{cases}\omega,&\text{if }\lvert a\rvert=\omega,\\ \emptyset,&\text{if }a\text{ is finite.}\end{cases}

Since for each a𝔪0+,a\in\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{+,\leq}, either aa finite or a¯\overline{a} is finite, the equation 𝖿𝗂𝗇(a)𝖿𝗂𝗇(a¯)=\mathsf{fin}(a)\cap\mathsf{fin}(\overline{a})=\emptyset holds in 𝔪0+,\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{+,\leq}, but not in 𝔪+,\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+,\leq}. ∎

3.2 Complex algebras of ω,,1\langle\omega,\cdot,1\rangle

Let =ω,,1\mathbb{N}^{\bullet}=\langle\omega,\cdot,1\rangle, 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet} be its complex algebra, and 𝐕\mathbf{V} be the variety generated by 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet}. Furthermore, let c(a):=ωac(a):=\omega\bullet a for every aωa\subseteq\omega.

We will first describe the smallest subalgebra of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet}.

Theorem 3.16.

𝔪0+𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}\cong\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}.

Proof.

For each nωn\in\omega, let

an:={mω:m has exactly n (possibly repeated) prime divisors}.\displaystyle a_{n}:=\{m\in\omega:m\text{ has exactly $n$ (possibly repeated) prime divisors}\}.

Then, a0={1}a_{0}=\{1\}, and the set of primes is circuit definable by

a1=(ω{1})(ω{1})¯{1}\displaystyle a_{1}=\overline{(\omega\setminus\{1\})\bullet(\omega\setminus\{1\})}\setminus\{1\}

It comes as no surprise that a1a_{1} is nothing else than the constant gg defined in (3.4). Each ana_{n} is circuit definable, since an=a1a1n–timesa_{n}=\underbrace{a_{1}\bullet\ldots\bullet a_{1}}_{n\text{--times}}. Clearly, aiaj=a_{i}\cap a_{j}=\emptyset for iji\neq j, and nωan=ω{0}\bigcup_{n\in\omega}a_{n}=\omega\setminus\{0\}; the latter can be shown via induction on the degree of a term.

Let A0A_{0} be the Boolean algebra with atoms {{1},ω{1}}\{\{1\},\omega\setminus\{1\}\}, and for n+1n+1 let An+1A_{n+1} be the Boolean closure of {ab:a,bAn}\{a\bullet b:a,b\in A_{n}\}.Furthermore, for each nωn\in\omega, let

bn+1=a0an¯.\displaystyle b_{n+1}=\overline{a_{0}\cup\ldots\cup a_{n}}.
Claim.

For 0<n0<n each AnA_{n} is finite with atoms a0,,a2n1,b2n1+1a_{0},\ldots,a_{2^{n-1}},b_{2^{n-1}+1}.

First, we consider n=1n=1. Computing {ab:a,bA0}\{a\bullet b:a,b\in A_{0}\}), we retain A0A_{0} (since {1}A0\{1\}\in A_{0}) and, obtain additionally, (ω{1})(ω{1})(\omega\setminus\{1\})\bullet(\omega\setminus\{1\}) which is the set of all positive composite numbers. Thus, the atom ω{1}\omega\setminus\{1\} of A0A_{0} splits into a1a_{1}, the set of all prime numbers, and b2b_{2}, the set of all composite numbers (including 0). Since 1=2111=2^{1-1}, the claim is true for n=1n=1.

Suppose that the claim is true for AnA_{n}, i.e. that the atoms of AnA_{n} are a0,a1,,a2n1,b2n1+1a_{0},a_{1},\ldots,a_{2^{n-1}},b_{2^{n-1}+1}. We need to show that the closure of {ab:a,bAn}\{a\bullet b:a,b\in A_{n}\} under the Boolean operations gives us An+1A_{n+1}. Since \bullet distributes over \cup it is sufficient to find aiaja_{i}\bullet a_{j} and aibn+1a_{i}\bullet b_{n+1} for i,jni,j\leq n. Now, if i,jni,j\leq n, then aiaj=ai+ja_{i}\bullet a_{j}=a_{i+j}, and thus, from aiaja_{i}\bullet a_{j} we obtain the disjoint sets

a0,a1,,a2n1,a2n1+1,,a2n1+2n1=a2(n+1)1.\displaystyle a_{0},\ a_{1},\ldots,a_{2^{n-1}},\ a_{2^{n-1}+1},\ldots,a_{2^{n-1}+2^{n-1}}=a_{2^{(n+1)-1}}.

From aib2n1+1a_{i}\bullet b_{2^{n-1}+1} we obtain

b2n1+1b2n1+2b2n1+1+2n1=b2(n+1)1+1\displaystyle b_{2^{n-1}+1}\supseteq b_{2^{n-1}+2}\supseteq b_{2^{n-1}+1+2^{n-1}}=b_{2^{(n+1)-1}+1}

The claim now follows from bmbm+1=amb_{m}\setminus b_{m+1}=a_{m}.

Clearly, {an:nω}\{a_{n}:n\in\omega\} is the set of atoms of 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}. Let f:𝔪0+𝔪0f:\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}\to\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet} be the mapping induced by f({n})=anf(\{n\})=a_{n}. Then, ff is bijective, and

f({n}+{m})=f({n+m})=an+m=anam=f({n})f({m}).\displaystyle f(\{n\}\bm{+}\{m\})=f(\{n+m\})=a_{n+m}=a_{n}\bullet a_{m}=f(\{n\})\bullet f(\{m\}).

Since +\bm{+} and \bullet are (completely) additive, ff is an isomorphism. ∎

It may be noted that that 0an¯0\in\overline{a_{n}} for all nωn\in\omega. Thus, {0}\{0\} is not definable from the constants, and

ω=𝔪0{an:nω}𝔪{an:nω}=ω{0}.\displaystyle\omega=\sum\nolimits^{\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}}\{a_{n}:n\in\omega\}\neq\sum\nolimits^{\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet}}\{a_{n}:n\in\omega\}=\omega\setminus\{0\}.

It follows that 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet} as a Boolean algebra is not a regular Boolean subalgebra of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet} [for the definition see 10].

Let us now consider the algebra 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}, i.e. the subalgebra of 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet} generated by its atoms {n}\{n\}. We note that {0}\{0\} is a nonzero annihilator, and thus is a proper congruence element - indeed, the smallest nonzero congruence element. Therefore, 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet} is subdirectly irreducible. By Theorem 3.14, no element of 𝐕𝐚𝐫𝔪+\mathbf{Var}\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} with more than two elements has a nonzero annihilator. Together with 𝔪0+𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}\cong\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet} we obtain that 𝔪1𝐕𝐚𝐫𝔪0\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}\not\in\mathbf{Var}\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}, and therefore, 𝐕𝐚𝐫𝔪1𝐕𝐚𝐫𝔪0\mathbf{Var}\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}\neq\mathbf{Var}\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}

Let θ\theta be the congruence generated by {0}\{0\}, and 𝔄\mathfrak{A} be the complex algebra of ω{0},,1\langle\omega\setminus\{0\},\cdot,1\rangle. Then, clearly, aθba{0}=b{0}a\theta b\Longleftrightarrow a\cup\{0\}=b\cup\{0\}, and 𝔪1/θ\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}/\theta is isomorphic to the singleton algebra 𝔄1\mathfrak{A}_{1} of 𝔄\mathfrak{A}; furthermore, 𝔪0𝔄0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}\cong\mathfrak{A}_{0}.

Owing to the presence of the nonzero annihilator {0}\{0\} we can still turn satisfiability (validity) of inequations into satisfiability (validity) of equations even though 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet} is not a discriminator algebra - it is subdirectly irreducible, but not simple:

(3.11) (x)[τ(x)σ(x)]\displaystyle(\exists\vec{x})[\tau(\vec{x})\neq\sigma(\vec{x})] (x)[τ(x)σ(x)](x)[{0}(τ(x)σ(x))={0}],\displaystyle\Longleftrightarrow(\exists\vec{x})[\tau(\vec{x})\vartriangle\sigma(\vec{x})\neq\bot]\Longleftrightarrow(\exists\vec{x})[\{0\}\bullet(\tau(\vec{x})\vartriangle\sigma(\vec{x}))=\{0\}],
(3.12) (x)[τ(x)σ(x)]\displaystyle(\forall\vec{x})[\tau(\vec{x})\neq\sigma(\vec{x})] (x)[τ(x)σ(x)](x)[{0}(τ(x)σ(x))={0}].\displaystyle\Longleftrightarrow(\forall\vec{x})[\tau(\vec{x})\vartriangle\sigma(\vec{x})\neq\bot]\Longleftrightarrow(\forall\vec{x})[\{0\}\bullet(\tau(\vec{x})\vartriangle\sigma(\vec{x}))=\{0\}].

We know already that the set of primes is definable in 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}. This can be generalized as follows: For nωn\in\omega let 𝐏𝐨(n)\mathbf{Po}(n) be the set of all powers of nn.

Theorem 3.17.

Let p0,,pnp_{0},\ldots,p_{n} be primes, and b=𝐏𝐨(p0)𝐏𝐨(pn)b=\mathbf{Po}(p_{0})\bullet\ldots\bullet\mathbf{Po}(p_{n}). Then, for all aωa\subseteq\omega,

ab𝔪1abFC(b).\displaystyle a\cap b\in\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}\Longleftrightarrow a\cap b\in FC(b).

Here, FC(b)FC(b) is the set of all finite or cofinite subsets of bb.

Proof.

\Leftarrow”:  We first show that 𝐏𝐨(p)𝔪1\mathbf{Po}(p)\in\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet} for every prime pp. Consider the following sequence:

ω{p}\displaystyle\omega\bullet\{p\} All multiples of pp
ω{p}¯\displaystyle\overline{\omega\bullet\{p\}} All nn not divisible by pp
ω{p}¯{1}¯\displaystyle\overline{\omega\bullet\{p\}}\cap\overline{\{1\}} All n1n\neq 1 not divisible by pp, i.e. coprime to pp, since pp is prime
ω(ω{p}¯{1}¯)\displaystyle\omega\bullet(\overline{\omega\bullet\{p\}}\cap\overline{\{1\}}) All nn with a factor 1\neq 1 coprime to pp
ω(ω{p}¯{1}¯)¯\displaystyle\overline{\omega\bullet(\overline{\omega\bullet\{p\}}\cap\overline{\{1\}})} All nn with (n1n\neq 1\Rightarrow no mm coprime to pp divides nn),

which defines the set of all powers of pp. It follows that b𝔪1b\in\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}. Since all singletons are in 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}, each finite or cofinite subset of bb is in 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}.

\Rightarrow”:  Consider the condition

(3.13) xbFC(b).\displaystyle x\cap b\not\in FC(b).

Suppose there are a term of minimal length τ(x0,,xk)\tau(x_{0},\ldots,x_{k}) and a0,,akωa_{0},\ldots,a_{k}\subseteq\omega such that a:=τ(a0,,ak)a:=\tau(a_{0},\ldots,a_{k}) satisfies (3.13). If a=sta=s\cup t, then ss or tt satisfy (3.13), contradicting the minimality of τ\tau; similarly, aa is not of the form s¯\overline{s}. Finally, let a=sta=s\bullet t. By the minimality of τ\tau, both sbs\cap b and tbt\cap b are in FC(b)FC(b), and by our assumption one must be cofinite in bb, say, ss. The cofinality implies there are q0,,qnωq_{0},\ldots,q_{n}\in\omega such that such that for all m0,,mnωm_{0},\ldots,m_{n}\in\omega,

(3.14) m0q0mnqnpm0pmns.\displaystyle m_{0}\geq q_{0}\land\ldots\land m_{n}\geq q_{n}\Rightarrow p^{m_{0}}\cdot\ldots\cdot p^{m_{n}}\in s.

Let q=p0j0pnjntbq=p_{0}^{j_{0}}\cdot\ldots p_{n}^{j_{n}}\in t\cap b. Then, {q}s\{q\}\bullet s is cofinite in ss by (3.14), and thus, cofinite in bb. It follows that a=sta=s\bullet t is cofinite in bb, contradicting our assumption. ∎

Theorem 3.17 does not hold in 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}: If a:=({3}ω)+{1}a:=(\{3\}\bullet\omega)\bm{+}\{1\}, then 𝐏𝐨(2)a\mathbf{Po}(2)\cap a is the set of all powers of 44. This also shows that 𝔪1𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet}\subsetneq\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}.

Theorem 3.18.
  1. (i)

    For each n>0n>0, 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet} contains an idempotent subsemigroup with nn generators and 2n12^{n}-1 elements.

  2. (ii)

    Suppose that GG is a subsemigroup of 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet} and a group. Then, |G|=1\lvert G\rvert=1.

Proof.

(i)  Let P={p1,,pn}P=\{p_{1},\ldots,p_{n}\} be a set of nn primes, and for each nonempty M={pi1,pik}PM=\{p_{i_{1}},\ldots p_{i_{k}}\}\subseteq P let aM:=𝐏𝐨pi1𝐏𝐨pika_{M}:=\mathbf{Po}p_{i_{1}}\bullet\ldots\bullet\mathbf{Po}p_{i_{k}}. Then, S={aM:MP}S=\{a_{M}:\emptyset\neq M\subseteq P\} is the desired semigroup generated by {a{pi}:1in}\{a_{\{p_{i}\}}:1\leq i\leq n\}; the identity element is aPa_{P}.

(ii)  Let ee be the neutral element of GG. If e=e=\emptyset, then a=ae=a=a=a\bullet e=a\bullet\emptyset=\emptyset for all aGa\in G, and thus, |G|=1\lvert G\rvert=1. Similarly, if e={0}e=\{0\} we have |G|=1\lvert G\rvert=1. Thus, suppose that e{0}e\not\subseteq\{0\}; it is easy to see that then a{0}a\not\subseteq\{0\} for all aGa\in G. Let n=min(e{0})n=\min(e\setminus\{0\}). Since ee=ee\bullet e=e, there are k,mek,m\in e with n=kmn=k\cdot m. Minimality of nn and n0n\neq 0 imply n=kn=k and m=1m=1 or n=mn=m and k=1k=1. In any case, n=1n=1, and thus, 1e1\in e.

Suppose that aGa\in G. Since aa1=ea\bullet a^{-1}=e and 1e1\in e, we have 1aa11\in a\cap a^{-1} and hence, a=a{1}aa1=ea=a\bullet\{1\}\subseteq a\bullet a^{-1}=e.

Conversely, e=e{1}ea=ae=e\bullet\{1\}\subseteq e\bullet a=a, so that altogether a=ea=e. ∎

4 Decidability of theories

Recall that for a BAO 𝔅\mathfrak{B}, we denote by 𝔅0\mathfrak{B}_{0} the smallest subalgebra of 𝔅\mathfrak{B}, i.e. the subalgebra of 𝔅\mathfrak{B} generated by the constants. In this section we consider the problems 𝐅𝐎𝔅\mathbf{FO}~\mathfrak{B}, 𝐄𝐪𝔅\mathbf{Eq}~\mathfrak{B}, and 𝐄𝐪𝐒𝐚𝐭𝔅\mathbf{EqSat}~\mathfrak{B} for the algebras 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}, 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}, 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}, 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}, 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet}, and 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bullet}. If 𝔅\mathfrak{B} is one of these algebras, we denote by 𝔅d\mathfrak{B}^{d} the algebra enhanced by an additional operator dd which represents a discriminator function on 𝔅\mathfrak{B}. A conjunctive grammar is a context–free grammar with an explicit intersection operation [13]. This section largely draws together work by Okhotin, [14], Jeż and Okhotin, [7], and Pinus and Vazhenin, [15].

We have the following undecidability results. If T is a Turing Machine, we can define the language VALC(T)\mbox{VALC}(T) of computations of TT, over the alphabet Σ={0,,k1}\Sigma=\{0,\ldots,k-1\}, for some k>0k>0. It does not really matter how these computations are encoded: the important point here is that VALC(T)=\mbox{VALC}(T)=\emptyset if and only if the language accepted by TT is empty. We may assume without loss of generality that no strings in VALC(T)\mbox{VALC}(T) begin with the letter 0. Any string sΣs\in\Sigma^{*} which does not begin with 0 may be regarded as a base-kk representation of a positive integer (s)\sharp(s). Thus, we obtain a 1–1 mapping fk:VALC(T){a}f_{k}:\mbox{VALC}(T)\rightarrow\{a\}^{*} given by fk(s)=a(s)f_{k}(s)=a^{\sharp(s)}. Thus, fk(VALC(T))f_{k}(\mbox{VALC}(T)) is a language over the 1-element alphabet {a}\{a\}.

Lemma 4.1.

[7])

  1. (i)

    For every Turing Machine TT, we can effectively construct conjunctive grammars GG and GG^{\prime} over the alphabet {a}\{a\} such that L(G)=fk(VALC(T))L(G)=f_{k}(\mbox{VALC}(T)).

  2. (ii)

    If aωa\subseteq\omega is recursive, there exists a finite system of equations of the form τi(y,x1,,xn)=σi(y,x1,,xn)\tau_{i}(y,x_{1},\ldots,x_{n})=\sigma_{i}(y,x_{1},\ldots,x_{n}) in the language with ,,+\cup,\cap,\bm{+} such that its unique solution is y=ay=a and xi=bix_{i}=b_{i} for some b1,,bn(2ω)n\langle b_{1},\ldots,b_{n}\rangle\in(2^{\omega})^{n}.

First, we compare the theories of these algebras.

Theorem 4.2.
  1. (i)

    𝐄𝐪(+)=𝐄𝐪(2ω,+,{0})\mathbf{Eq}~(\mathbb{N}^{\bm{+}})=\mathbf{Eq}~(2^{\omega},\bm{+},\{0\}).

  2. (ii)

    𝐄𝐪𝔪0+=𝐄𝐪𝔪+\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}=\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}.

  3. (iii)

    𝐄𝐪𝐒𝐚𝐭𝔪0+𝐄𝐪𝐒𝐚𝐭𝔪+\mathbf{EqSat}~\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}\neq\mathbf{EqSat}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}.

  4. (iv)

    𝐄𝐪𝔪0+,d𝐄𝐪𝔪+,d\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+},d}\neq\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+,d}.

  5. (v)

    𝐄𝐪𝔪0𝐄𝐪𝔪\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}\neq\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}.

Proof.

(i)  The mapping f:ω{aω:a is finite}f:\omega\to\{a\subseteq\omega:a\text{ is finite}\} which maps nn to {n}\{n\} is an embedding of monoids, and thus, 𝐄𝐪(2ω,+,{0})𝐄𝐪(+)\mathbf{Eq}~(2^{\omega},\bm{+},\{0\})\subseteq\mathbf{Eq}~(\mathbb{N}^{+}). The reverse inclusion follows from the fact that +\mathbb{N}^{+} is the free monoid on a single generator.

(ii)  Since 𝔪0+𝔪+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}\leq\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}, it follows that 𝔪0+𝐕𝐚𝐫(𝔪+)\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}\in\mathbf{Var}(\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}). Conversely, each 𝔅n\mathfrak{B}_{n} is in 𝐕𝐚𝐫(𝔪0+)\mathbf{Var}(\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}) by Corollary 3.7, and thus, 𝔪+𝐕𝐚𝐫(𝔪0+)\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}\in\mathbf{Var}(\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}).

(iii)  The equation

(4.1) x+{1}=x¯\displaystyle x\bm{+}\{1\}=\overline{x}

has a unique solution in 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}, namely, the set of even numbers, which is not in 𝔪0+\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}}.

(iv) This is a slight generalization of Theorem 3.15(2). The equation (4.1) has no solution in FC(ω)FC(\omega), i.e. (x)[x+{1}x¯](\forall x)[x\bm{+}\{1\}\neq\overline{x}] holds in 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}. This is equivalent to the equation d((x+{1})x¯)=ωd((x\bm{+}\{1\})\vartriangle\overline{x})=\omega which is not valid in 𝔪+,d\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{+,d}.

(v)  Let a𝔪𝔪0a\in\operatorname{{\mathfrak{Cm}}}~\mathbb{N}\setminus\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} be recursive. Such set exists, since every every set definable by an arithmetic circuit is in the bounded hierarchy BH [16], and the bounded hierarchy is known to be contained within the zeroth Grzegorczyk class, 0\mathcal{E}^{0}_{*}. By Lemma 4.1 there is a first order sentence (x)φ(x)(\exists x)\varphi(x) such that 𝔪φ(x/s)\operatorname{{\mathfrak{Cm}}}~\mathbb{N}\models\varphi(x/s) if and only if  s=as=a. It follows that 𝔪0⊧̸(x)φ(x)\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}\not\models(\exists x)\varphi(x), i.e. 𝔪0(x)¬φ(x)\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}\models(\forall x)\neg\varphi(x). Since 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} is a discriminator algebra, there is an equation τ(x)=σ(x)\tau(x)=\sigma(x), such that 𝔪0(x)¬φ(x)\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}\models(\forall x)\neg\varphi(x) if and only if  𝔪0τ(x)=σ(x)\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}\models\tau(x)=\sigma(x). Since 𝔪(x)φ(x)\operatorname{{\mathfrak{Cm}}}~\mathbb{N}\models(\exists x)\varphi(x), τ(x)=σ(x)\tau(x)=\sigma(x) cannot hold in 𝔪\operatorname{{\mathfrak{Cm}}}~\mathbb{N}. ∎

Given any conjunctive grammar GG with non-terminals X1,,XnX_{1},\ldots,X_{n} over the alphabet {a}\{a\}, we may effectively construct a system of language equations \mathcal{E} in variables V1,,VnV_{1},\ldots,V_{n}, with the property that \mathcal{E} has a unique least (under componentwise-inclusion) solution S10,,Sn0S_{1}^{0},\ldots,S_{n}^{0} and, moreover, for all ii (1in1\leq i\leq n), SiS_{i} is the set of strings of {a}\{a\}^{*} to which GG assigns the category XiX_{i}. Let us assume that X1X_{1} is the start-symbol of GG; i.e., L(G)L(G) is the set of strings to which GG assigns category X1X_{1}.

Theorem 4.3.

Let OO be any collection of isotone operators on \mathbb{N} with +O\bm{+}\in O. Then 𝐄𝐪𝐒𝐚𝐭𝔪(,O)\mathbf{EqSat}~\operatorname{{\mathfrak{Cm}}}(\mathbb{N},O) is co–r.e.-complete.

Proof.

For the lower bound, it suffices to establish the result in the case 𝔪(,{+})=𝔪+\operatorname{{\mathfrak{Cm}}}(\mathbb{N},\{+\})=\operatorname{{\mathfrak{Cm}}}\mathbb{N}^{\bm{+}}. We use the fact that the emptiness of the languages accepted by a Turing machine TT is equivalent to the validity of the language equations \mathcal{E}, as outlined above. We must translate the language equations in \mathcal{E} in the logical signature {ε,{a},,,}\{\varepsilon,\{a\},\cup,\cap,\cdot\}, (where \cdot denotes concatenation) into integer-set equations, by replacing ε\varepsilon by {0}\{0\}, {a}\{a\} by {1}\{1\}, and \cdot by +\bm{+}. Let the result of this translation be \mathcal{E}^{*}. If g:{a}g:\{a\}^{*}\rightarrow\mathbb{N} is the isomorphism given by akka^{k}\mapsto k, then S1,,SnS_{1},\ldots,S_{n} is a solution of \mathcal{E} if and only if g(S1),,g(Sn)g(S_{1}),\ldots,g(S_{n}) is a solution of \mathcal{E}^{*}.

Altogether, we have:

Acc(T)=\displaystyle\mbox{Acc}(T)=\emptyset \displaystyle\Leftrightarrow VALC(T)=\displaystyle\mbox{VALC}(T)=\emptyset
\displaystyle\Leftrightarrow fk(VALC(T))=\displaystyle f_{k}(\mbox{VALC}(T))=\emptyset
\displaystyle\Leftrightarrow LG=\displaystyle L_{G}=\emptyset
\displaystyle\Leftrightarrow S10=\displaystyle S_{1}^{0}=\emptyset
\displaystyle\Leftrightarrow {X1=}\mathcal{E}\cup\{X_{1}=\emptyset\} has a solution
\displaystyle\Leftrightarrow {X1=}\mathcal{E}^{*}\cup\{X_{1}=\emptyset\} has a solution.

This establishes that 𝐄𝐪𝐒𝐚𝐭𝔪(,O)\mathbf{EqSat}~\operatorname{{\mathfrak{Cm}}}(\mathbb{N},O) is co-r.e.-hard, as required.

To show that 𝐄𝐪𝐒𝐚𝐭𝔪(,O)\mathbf{EqSat}~\operatorname{{\mathfrak{Cm}}}(\mathbb{N},O) is co-r.e., it suffices to prove that, for any mm-tuple of variables x¯\bar{x} and any term τ(x¯)\tau(\bar{x}),

(4.2) τ(x¯)= has a solution in (2ω)m\tau(\bar{x})=\emptyset\text{ has a solution in $(2^{\omega})^{m}$}

if and only if, for all nn,

(4.3) τ(x¯)[0,n]= has a solution in (2[0,n])m,\tau(\bar{x})\cap[0,n]=\emptyset\text{ has a solution in $(2^{[0,n]})^{m}$},

since the condition (4.3) is evidently decidable for fixed nn. In the sequel, if s¯=(s1,,sm)\bar{s}=(s_{1},\ldots,s_{m}) and t¯=(t1,,tm)\bar{t}=(t_{1},\ldots,t_{m}) are mm-tuples of sets, we write s¯[0,n]\bar{s}\cap[0,n] for the mm-tuple (s1[0,n],,sn[0,n])(s_{1}\cap[0,n],\ldots,s_{n}\cap[0,n]), s¯t¯\bar{s}\cup\bar{t} for the mm-tuple (s1t¯1,,snt¯m)(s_{1}\cup\bar{t}_{1},\ldots,s_{n}\cup\bar{t}_{m}) and s¯t¯\bar{s}\subseteq\bar{t} for the condition s1t1smtms_{1}\subseteq t_{1}\wedge\cdots\wedge s_{m}\subseteq t_{m}.

The direction from (4.2) to (4.3) is easy. For suppose τ(s¯)=\tau(\bar{s})=\emptyset. Then, for all nn, τ(s¯)[0,n]=\tau(\bar{s})\cap[0,n]=\emptyset, whence, by the monotonicity of the operators on OO, τ(s¯[0,n])=\tau(\bar{s}\cap[0,n])\cap=\emptyset. To show the converse, let VnV_{n} denote, for any nn, the set of pairs s¯,n\langle\bar{s},n\rangle where s¯\bar{s} is a solution of τ(x¯)[0,n]=\tau(\bar{x})\cap[0,n]=\emptyset in (2[0,n])m(2^{[0,n]})^{m}. Thus, VnV_{n} is finite, and, assuming (4.3) for all nn, non-empty. Define the directed graph (V,E)(V,E) by setting V=VnV=\bigcup V_{n} and

E={(s¯,n,t¯,n+1):s¯,nVn,t¯,n+1Vn+1 and s¯t¯}.E=\left\{\left(\langle\bar{s},n\rangle,\langle\bar{t},n+1\rangle\right):\langle\bar{s},n\rangle\in V_{n},\ \ \langle\bar{t},n+1\rangle\in V_{n+1}\text{ and }\bar{s}\subseteq\bar{t}\right\}.

Thus, (V,E)(V,E) is a finitely branching, infinite tree, and so has an infinite path s¯0,0,s¯1,1,\langle\bar{s}_{0},0\rangle,\langle\bar{s}_{1},1\rangle,\ldots, where s¯0s¯1\bar{s}_{0}\subseteq\bar{s}_{1}\subseteq\cdots. Letting s¯=s¯n\bar{s}=\bigcup\bar{s}_{n}, we have, for all nn, τ(s¯)[0,n]=τ(s¯n)[0,n]=\tau(\bar{s})\cap[0,n]=\tau(\bar{s}_{n})\cap[0,n]=\emptyset. Hence τ(s¯)=\tau(\bar{s})=\emptyset, whence (4.2) holds. ∎

It immediately follows from Theorem 4.3 that

Corollary 4.4.

𝐄𝐪𝐒𝐚𝐭𝔪\mathbf{EqSat}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N} is co-r.e.-hard.

In Corollary 3.10 we showed that 𝐄𝐪𝔪0+\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N}^{\bm{+}} is co–re. On the other hand, it is not obvious that we can find a (computable) bound for the smallest witnesses of inequations in these languages.

While the membership problem for 𝔪0\operatorname{{\mathfrak{Cm}}}_{0}\mathbb{N} is a word problem, the satisfaction problem (1.6) is related to the equational theory:

Theorem 4.5.

The equational theory of 𝔪1\operatorname{{\mathfrak{Cm}}}_{1}\mathbb{N}^{\bullet} is decidable if and only if  the satisfaction problem (1.6) is decidable.

Proof.

\Rightarrow”:  Let nωn\in\omega and τ(x)\tau(\vec{x}) be a term with variables x\vec{x}. Then,

(x)[{n}τ(x)]¬((x)[{n}τ(x)=]).\displaystyle\exists(\vec{x})[\{n\}\cap\tau(\vec{x})\neq\emptyset]\Longleftrightarrow\neg((\forall\vec{x})[\{n\}\cap\tau(\vec{x})=\emptyset]).

\Leftarrow”:  Suppose that τ(x),σ(x)\tau(\vec{x}),\ \sigma(\vec{x}) are terms with variables among x\vec{x}; w.l.o.g. we may suppose that σ(x)=\sigma(\vec{x})=\emptyset. Then,

(x)[τ(x)=](x)[0{0}τ(x)]¬((x)[0{0}τ(x)]).\displaystyle(\forall\vec{x})[\tau(\vec{x})=\emptyset]\Longleftrightarrow(\forall\vec{x})[0\not\in\{0\}\bullet\tau(\vec{x})]\Longleftrightarrow\neg((\exists\vec{x})[0\in\{0\}\bullet\tau(\vec{x})]).

As for equational theories, results are known as long as we have the wherewithal to convert equations into inequations. Determining whether an equation belongs to the equational theory of a language \mathcal{L} over some interpretation 𝔄\mathfrak{A} is the co-problem of determining whether an inequation in \mathcal{L} is satisfiable in 𝔄\mathfrak{A}. If we have a discriminator at our disposal, then (2.2), (2.3) and Theorem 4.3 imply

Theorem 4.6.

The set 𝐄𝐪𝔪+,d\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+},d} is r.e.-hard. Hence, 𝐄𝐪𝔪\mathbf{Eq}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N} is r.e.-hard.

If S,\langle S,\circ\rangle is a semigroup, then its power structure is the semigroup of complexes of SS. The following result is quoted by Pinus and Vazhenin, [15, Theorem 2.3.2]:

Theorem 4.7.

[2]  For a variety 𝐕\mathbf{V} of semigroups the class of power structures of elements of 𝐕\mathbf{V} has a decidable elementary theory if and only if 𝐕𝐕𝐚𝐫({xyz=xz})\mathbf{V}\subseteq\mathbf{Var}(\{x\circ y\circ z=x\circ z\}).

Neither ω,+,0\langle\omega,+,0\rangle nor ω,,1\langle\omega,\cdot,1\rangle satisfy xyz=xzx\circ y\circ z=x\circ z. Since the power structure of ω,+,0\langle\omega,+,0\rangle is a reduct of 𝔪+\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}}, this is another way to show that 𝐅𝐎𝔪+\mathbf{FO}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bm{+}} is undecidable. It also applies to ω,,1\langle\omega,\cdot,1\rangle:

Corollary 4.8.

𝐅𝐎𝔪\mathbf{FO}~\operatorname{{\mathfrak{Cm}}}~\mathbb{N}^{\bullet} is undecidable.

References

  • Baker, [1977] Baker, K. A. (1977). Finite equational bases for finite algebras in a congruence–distributive equational class. Advances in Mathematics, 24(3):207–243.
  • Bayasgalan, [1988] Bayasgalan, B. (1988). Decidability of theories of derived structures of semigroups. Algebraic Systems and their Varieties, Ural. Gos. Univ. Sverdlovsk, pages 4–13. In Russian.
  • Birkhoff, [1935] Birkhoff, G. (1935). On the structure of abstract algebras. Proceedings of the Cambridge Philosophical Society, 31:433–454.
  • Burris and Sankappanavar, [1981] Burris, S. and Sankappanavar, H. P. (1981). A Course in Universal Algebra. Springer-Verlag, New York.
  • Glaßer et al., [2007] Glaßer, C., Reitwießner, C., Travers, S. D., and Waldherr, M. (2007). Satisfiability of algebraic circuits over sets of natural numbers. In Arvind, V. and Prasad, S., editors, Foundations of Software Technology and Theoretical Computer Science, 27th International Conference (FSTTCS 2007), New Delhi, India, December 12-14, Proceedings, volume 4855 of Lecture Notes in Computer Science, pages 253–264. Springer.
  • Goldblatt, [1989] Goldblatt, R. (1989). Varieties of complex algebras. Annals of Pure and Applied Logic, 44:173–242.
  • Jeż and Okhotin, [2008] Jeż, A. and Okhotin, A. (2008). On the computational completeness of equations over sets of natural numbers. In Aceto, L., Damgård, I., Halldórsson, L. G. M., Ingólfsdóttir, A., and Walukiewicz, I., editors, Proceedings of ICALP 2008, Part II, volume 5126 of LNCS, pages 63–74. Springer.
  • Jipsen, [1992] Jipsen, P. (1992). Computer aided investigations of relation algebras. PhD thesis, Vanderbilt University.
  • Jónsson and Tarski, [1951] Jónsson, B. and Tarski, A. (1951). Boolean algebras with operators I. American Journal of Mathematics, 73:891–939.
  • Koppelberg, [1989] Koppelberg, S. (1989). General Theory of Boolean Algebras, volume 1 of Handbook on Boolean Algebras. North–Holland.
  • McKenzie and Wagner, [2003] McKenzie, P. and Wagner, K. W. (2003). The complexity of membership problems for circuits over sets of natural numbers. In Alt, H. and Habib, M., editors, STACS 2003, 20th Annual Symposium on Theoretical Aspects of Computer Science, Berlin, Germany, February 27 - March 1, 2003, Proceedings, volume 2607 of Lecture Notes in Computer Science, pages 571–582. Springer–Verlag.
  • McKenzie and Wagner, [2007] McKenzie, P. and Wagner, K. W. (2007). The complexity of membership problems for circuits over sets of natural numbers. Computational Complexity, 16(3):211–244.
  • Okhotin, [2001] Okhotin, A. (2001). Conjunctive grammars. Journal of Automata, Languages and Combinatorics, 4:519–535.
  • Okhotin, [2003] Okhotin, A. (2003). Decision problems for language equations with boolean operations. In Baeten, J. C. M., Lenstra, J. K., Parrow, J., and Woeginger, G. J., editors, ICALP, volume 2719 of Lecture Notes in Computer Science, pages 239–251. Springer.
  • Pinus and Vazhenin, [2005] Pinus, A. G. and Vazhenin, Y. M. (2005). Elementary classification and decidability of theories of derived structures. Russian Math. Surveys, 60:395–432.
  • Pratt-Hartmann and Düntsch, [2009] Pratt-Hartmann, I. and Düntsch, I. (2009). Functions definable by arithmetic circuits. In Ambos-Spies, K., Löwe, B., and Merkle, W., editors, Mathematical Theory and Computational Practice: 5th Conference on Computability in Europe, CiE 2009, volume 5635 of Lecture Notes in Computer Science, pages 409–418, Heidelberg. Springer Verlag.
  • Reich, [1996] Reich, P. (1996). Complex algebras of semigroups. PhD thesis, Iowa State University, Ames.