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

On the number of universal algebraic geometries

Erhard Aichinger Institut für Algebra, Johannes Kepler Universität Linz, Altenberger Straße 69, 4040 Linz, Austria [email protected]  and  Bernardo Rossi Institut für Algebra, Johannes Kepler Universität Linz, Altenberger Straße 69, 4040 Linz, Austria [email protected]
Abstract.

The algebraic geometry of a universal algebra 𝐀\mathbf{A} is defined as the collection of solution sets of systems of term equations. Two algebras 𝐀1\mathbf{A}_{1} and 𝐀2\mathbf{A}_{2} are called algebraically equivalent if they have the same algebraic geometry. We prove that on a finite set AA with |A|>3\lvert A\rvert>3 there are countably many algebraically inequivalent Mal’cev algebras and that on a finite set AA with |A|>2\lvert A\rvert>2 there are continuously many algebraically inequivalent algebras.

Key words and phrases:
Universal algebraic geometry, Clones, Algebraic sets
2010 Mathematics Subject Classification:
08A40, (08A62, 08B05, 03C05)
Supported by the Austrian Science Fund (FWF): P33878.

1. Introduction

Universal algebraic geometry, introduced in [18, 10], is based on the notion of algebraic sets. Given a universal algebra 𝐀=(A;(fi)iI){\mathbf{A}}=(A;(f_{i})_{i\in I}), a subset BB of AnA^{n} is algebraic if it is the solution set of a system of (possibly infinitely many) term equations in the language of 𝐀{\mathbf{A}}. Following [17], we denote the collection of the algebraic subsets of AnA^{n} by Algn𝐀\operatorname{Alg}_{n}{\mathbf{A}} and we define the universal algebraic geometry of 𝐀{\mathbf{A}} by Alg𝐀:=nAlgn𝐀\operatorname{Alg}{\mathbf{A}}:=\bigcup_{n\in\mathbb{N}}\operatorname{Alg}_{n}{\mathbf{A}}. Clearly, for a universal algebra 𝐀{\mathbf{A}}, Alg𝐀\operatorname{Alg}{\mathbf{A}} is completely determined by its clone of term functions (cf. [8, 19]). We define the universal algebraic geometry of a clone 𝒞\mathcal{C} on a set AA by Alg𝒞:=Alg(A;𝒞)\operatorname{Alg}\mathcal{C}:=\operatorname{Alg}(A;\mathcal{C}). Following [17], we say that two clones 𝒞\mathcal{C} and 𝒟\mathcal{D} on the same set AA are algebraically equivalent, 𝒞alg𝒟\mathcal{C}\operatorname{\sim_{\text{alg}}}\mathcal{D}, if Alg𝒞=Alg𝒟\operatorname{Alg}\mathcal{C}=\operatorname{Alg}\mathcal{D}. In [17, 21] one finds examples of different clones that are algebraically equivalent; in fact Tóth and Waldhauser [21] proved that on the two element set there are only finitely many algebraically inequivalent clones. Pinus showed that on each finite set there are only finitely many universal algebraic geometries closed under taking unions [17, 6]. The aim of this note is to provide infinite families of algebraically inequivalent clones on a finite set.

Analysing the construction of 202^{\aleph_{0}} clones on {0,1,2}\{0,1,2\} in [13], we obtain:

Theorem 1.1.

Let AA be a finite set with |A|3\lvert A\rvert\geq 3. Then there are 202^{\aleph_{0}} algebraically inequivalent clones on AA.

The proof will be given in Section 5.

A clone 𝒞\mathcal{C} is called a Mal’cev clone if it contains a ternary function such that for all a,bAa,b\in A,

d(a,b,b)=d(b,b,a)=a,d(a,b,b)=d(b,b,a)=a,

and it is called constantive if every unary constant function on AA is in 𝒞\mathcal{C}. Using Idziak’s construction from [12], we prove:

Theorem 1.2.

Let AA be a finite set with |A|4\lvert A\rvert\geq 4. Then there are exactly 0\aleph_{0} algebraically inequivalent constantive Mal’cev clones on AA.

The proof will be given in Section 4. It relies on the following fact:

Theorem 1.3.

Let pp be a prime, and let nn\in\mathbb{N}. Then there are exactly 0\aleph_{0} algebraically inequivalent clones that contain Pol(np2,+)\operatorname{Pol}(\mathbb{Z}_{np^{2}},+).

The proof will be given in Section 3.

In Table 1 we summarize our current knowledge on the number of algebraically inequivalent clones on a finite set.

Property of the clone on an nn-element set Number of clones Number of clones modulo alg\operatorname{\sim_{\text{alg}}}
n=2n=2 0\aleph_{0} [20] finite [21]
n3n\geq 3 202^{\aleph_{0}} [13] 202^{\aleph_{0}} (Theorem 1.1)
constantive, n3n\geq 3 202^{\aleph_{0}} [1] ?
equationally additive ? finite [17, 6]
Mal’cev, n3n\leq 3 finite [7] finite
constantive, Mal’cev, n4n\geq 4 0\aleph_{0} [2, 12] 0\aleph_{0} (Theorem 1.2)
contains Clo(n,+)\operatorname{Clo}(\mathbb{Z}_{n},+), nn squarefree finite [11, 4, 15] finite
contains Pol(mp2,+)\operatorname{Pol}(\mathbb{Z}_{mp^{2}},+), m1m\geq 1 0\aleph_{0} [2, 12] 0\aleph_{0} (Theorem 1.3)
Table 1. The first column lists the properties of the clones we consider, the second column reports the number of such clones and the third column lists the number of such clones modulo alg\operatorname{\sim_{\text{alg}}}. The number pp is assumed to be prime. Question marks represent open problems.

2. Notation

We write \mathbb{N} for the set of positive integers and for nn\in\mathbb{N}, [n]:={1,,n}[{n}]:=\{1,\dots,n\}. For a set AA the ii-th component of 𝒂𝑨𝒏\mathbfsl{a}\in A^{n} is denoted by aia_{i} and 𝒂(𝒊)\mathbfsl{a}(i). We will use the notions clone, Poln(𝐀)\operatorname{Pol}_{n}({\mathbf{A}}) and Clon(𝐀)\operatorname{Clo}_{n}({\mathbf{A}}) as they are commonly used in universal algebra [8, 16]. For kk\in\mathbb{N} with knk\leq n, we define the kk-th nn-ary projection πk(n):AnA\pi_{k}^{(n)}:A^{n}\to A by πk(n)(a1,,an)=ak\pi_{k}^{(n)}(a_{1},\dots,a_{n})=a_{k} for all 𝒂𝑨𝒏\mathbfsl{a}\in A^{n}. We set 𝒥A:={πk(n)n and kn}\mathcal{J}_{A}:=\{\pi_{k}^{(n)}\mid n\in\mathbb{N}\text{ and }k\leq n\}. For a clone 𝒞\mathcal{C} on AA, 𝒞[n]\mathcal{C}^{[n]} is the set of nn-ary functions in 𝒞\mathcal{C}. For kk\in\mathbb{N}, for σ:[k][n]\sigma\colon[{k}]\to[{n}] and for f𝒞[k]f\in\mathcal{C}^{[k]}, we define fσ𝒞[n]f_{\sigma}\in\mathcal{C}^{[n]} by fσ(x1,,xn):=f(xσ(1),,xσ(k))f_{\sigma}(x_{1},\dots,x_{n}):=f(x_{\sigma(1)},\dots,x_{\sigma(k)}) for all x1,,xnAx_{1},\dots,x_{n}\in A, and we call fσf_{\sigma} a minor of ff. If tt is a kk-ary term in the language of a universal algebra 𝐀{\mathbf{A}}, we write tσt_{\sigma} for the term t(xσ(1),,xσ(k))t(x_{\sigma(1)},\dots,x_{\sigma(k)}). Observe that in this case (tσ)𝐀=(t𝐀)σ(t_{\sigma})^{\mathbf{A}}=(t^{\mathbf{A}})_{\sigma}.

Let AA be a set and let 𝒞\mathcal{C} be a clone on AA. For nn\in\mathbb{N} and for XAnX\subseteq A^{n}, we define V𝒞(X)\text{V}_{\mathcal{C}}(X) to be the intersection of all the elements of Algn𝒞\operatorname{Alg}_{n}\mathcal{C} that contain XX as a subset. Since the intersection of a collection of elements of Algn𝒞\operatorname{Alg}_{n}\mathcal{C} is an element of Algn𝒞\operatorname{Alg}_{n}\mathcal{C}, we infer that V𝒞(X)Algn𝒞\text{V}_{\mathcal{C}}(X)\in\operatorname{Alg}_{n}\mathcal{C}. The following lemma will be used to assess whether a set XX is algebraic with respect to a clone 𝒞\mathcal{C}, which is equivalent to X=V𝒞(X)X=\text{V}_{\mathcal{C}}(X).

Lemma 2.1.

Let AA be a set, let 𝒞\mathcal{C} be a clone on AA, let nn\in\mathbb{N}, let XAnX\subseteq A^{n}, and let 𝐚𝐀𝐧\mathbfsl{a}\in A^{n}. Then we have

𝒂V𝒞(𝑿)(𝒇,𝒈𝒞[𝒏]:𝒇|𝑿=𝒈|𝑿𝒇(𝒂)=𝒈(𝒂)).\mathbfsl{a}\in\text{V}_{\mathcal{C}}(X)\Longleftrightarrow\Bigl{(}\forall f,g\in\mathcal{C}^{[n]}:f|_{X}=g|_{X}\Rightarrow f(\mathbfsl{a})=g(\mathbfsl{a})\Bigr{)}.
Proof.

\Leftarrow”: We assume that for all f,g𝒞[n]f,g\in\mathcal{C}^{[n]} with f|X=g|Xf|_{X}=g|_{X}, we have f(𝒂)=𝒈(𝒂)f(\mathbfsl{a})=g(\mathbfsl{a}). We show that 𝒂V𝒞(𝑿)\mathbfsl{a}\in\text{V}_{\mathcal{C}}(X). To this end, we show that for each BAlgn𝒞B\in\operatorname{Alg}_{n}\mathcal{C} with XBX\subseteq B, we have 𝒂𝑩\mathbfsl{a}\in B. Since BB is algebraic with respect to 𝒞\mathcal{C}, there exists a system of 𝒞\mathcal{C}-equations {piqiiI}\{p_{i}\approx q_{i}\mid i\in I\} such that B={𝒙𝑨𝒏𝒊𝑰:𝒑𝒊(𝒙)=𝒒𝒊(𝒙)}B=\{\mathbfsl{x}\in A^{n}\mid\forall i\in I\colon p_{i}(\mathbfsl{x})=q_{i}(\mathbfsl{x})\}. Since XBX\subseteq B, we have pi|X=qi|Xp_{i}|_{X}=q_{i}|_{X} for all iIi\in I. Therefore, the assumption yields pi(𝒂)=𝒒𝒊(𝒂)p_{i}(\mathbfsl{a})=q_{i}(\mathbfsl{a}) for all iIi\in I, and thus 𝒂𝑩\mathbfsl{a}\in B.

\Rightarrow”: We assume that there exist f,g𝒞[n]f,g\in\mathcal{C}^{[n]} such that f|X=g|Xf|_{X}=g|_{X} and f(𝒂)𝒈(𝒂)f(\mathbfsl{a})\neq g(\mathbfsl{a}). Then B:={𝒙𝑨𝒏𝒇(𝒙)=𝒈(𝒙)}B:=\{\mathbfsl{x}\in A^{n}\mid f(\mathbfsl{x})=g(\mathbfsl{x})\} satisfies BAlgn𝒞B\in\operatorname{Alg}_{n}\mathcal{C}, XBX\subseteq B, and 𝒂𝑩\mathbfsl{a}\notin B. Thus, 𝒂V𝒞(𝑿)\mathbfsl{a}\notin\text{V}_{\mathcal{C}}(X). ∎

3. Countably many algebraically inequivalent constantive expansions of np2\mathbb{Z}_{np^{2}}

Let 𝐀=(A;+,,0,(fi)iI){\mathbf{A}}=(A;+,-,0,(f_{i})_{i\in I}) be an expanded group. Following [3], a function f:AnAf\colon A^{n}\to A is absorbing if f(a1,,an)=0f(a_{1},\dots,a_{n})=0 for all a1,,anAa_{1},\dots,a_{n}\in A with 0{a1,,an}0\in\{a_{1},\dots,a_{n}\}.

Lemma 3.1.

Let pp be a prime, let n,d,kn,d,k\in\mathbb{N} with d2d\geq 2 and kd+1k\geq d+1, let 𝐀d{\mathbf{A}}_{d} be the algebra (np2;+,,0,fd)(\mathbb{Z}_{np^{2}};+,-,0,f_{d}), where fd:np2dnp2f_{d}\colon\mathbb{Z}_{np^{2}}^{d}\to\mathbb{Z}_{np^{2}} is defined by

fd(x1,,xd)=npj=1dxj for x1,,xdnp2,f_{d}(x_{1},\dots,x_{d})=np\prod_{j=1}^{d}x_{j}\text{ for }x_{1},\dots,x_{d}\in\mathbb{Z}_{np^{2}},

and let gPolk(𝐀d)g\in\operatorname{Pol}_{k}({\mathbf{A}}_{d}). If gg is absorbing, then gg is the constant zero function.

Lemma 3.1 states that the algebra 𝐀d{\mathbf{A}}_{d} is dd-supernilpotent (cf. [5]). We provide a proof that makes no use of the notion of supernilpotency.

Proof.

Let 𝒞d\mathcal{C}_{d} be the set of all operations on np2\mathbb{Z}_{np^{2}} that are induced by a sum of monomials of the form axi1n1xirnrax_{i_{1}}^{n_{1}}\dots x_{i_{r}}^{n_{r}}, where n1,,nrn_{1},\dots,n_{r}\in\mathbb{N}, xi1,,xirx_{i_{1}},\dots,x_{i_{r}} are pairwise distinct variables, anp2a\in\mathbb{Z}_{np^{2}}, and the following property is satisfied:

(3.1) (r=0) or (r=n1=1) or (2n1++nrd and np divides a).(r=0)\text{ or }(r=n_{1}=1)\text{ or }(2\leq n_{1}+\dots+n_{r}\leq d\text{ and }np\text{ divides }a).

It is clear that {+,,0,fd}𝒞d\{+,-,0,f_{d}\}\subseteq\mathcal{C}_{d}, that the projection clone satisfies 𝒥np2𝒞d\mathcal{J}_{\mathbb{Z}_{np^{2}}}\subseteq\mathcal{C}_{d}, and that 𝒞d\mathcal{C}_{d} contains the constant functions. Furthermore, it is straightforward to verify that for t1,t2,,td𝒞dt_{1},t_{2},\dots,t_{d}\in\mathcal{C}_{d} we have that t1+t2𝒞dt_{1}+t_{2}\in\mathcal{C}_{d}, t1t2𝒞dt_{1}-t_{2}\in\mathcal{C}_{d} and fd(t1,,td)𝒞df_{d}(t_{1},\dots,t_{d})\in\mathcal{C}_{d}. Hence Pol(𝐀d)𝒞d\operatorname{Pol}({\mathbf{A}}_{d})\subseteq\mathcal{C}_{d}. We assume that gPolk(𝐀d)g\in\operatorname{Pol}_{k}({\mathbf{A}}_{d}) is absorbing, and consider the representation of gg as a sum of monomials, each one satisfying (3.1). Since gg is absorbing, g(0,x2,,xk)g(0,x_{2},\dots,x_{k}) is the constant 0-function. Thus, the sum of those monomials of gg that do not involve x1x_{1} induces the constant zero function. Therefore, there exists a representation of gg as a sum of monomials of the above form, each of them involving the variable x1x_{1}. A similar argument applies to all the other variables x2,,xkx_{2},\dots,x_{k}. Therefore, there exists a (possibly empty) sum of monomials that induces the same polynomial function as gg such that each monomial involves all the variables x1,,xkx_{1},\dots,x_{k} and satisfies (3.1). If xi1n1xirnrx_{i_{1}}^{n_{1}}\dots x_{i_{r}}^{n_{r}} is such a monomial, then rkr\geq k, and therefore n1++nrkn_{1}+\cdots+n_{r}\geq k. Since k3k\geq 3, (3.1) implies dn1++nrkd\geq n_{1}+\cdots+n_{r}\geq k, contradicting the assumption kd+1k\geq d+1. Hence gg is induced by the empty sum of monomials, and therefore it is the constant 0-function. ∎

Proof of Theorem 1.3.

For each d{1}d\in\mathbb{N}\setminus\{1\}, let 𝐀d{\mathbf{A}}_{d} and fdf_{d} be as in the statement of Lemma 3.1. For d{1}d\in\mathbb{N}\setminus\{1\}, for mm\in\mathbb{N} and for Xnp2mX\subseteq\mathbb{Z}_{np^{2}}^{m}, we use Vd(X)\text{V}_{d}(X) as a shorthand for VPol(𝐀d)(X)\text{V}_{\operatorname{Pol}({\mathbf{A}}_{d})}(X). Let l,il,i\in\mathbb{N} with l>i2l>i\geq 2. We claim that Pol(𝐀i)≁algPol(𝐀l)\operatorname{Pol}({\mathbf{A}}_{i})\not\sim_{\text{alg}}\operatorname{Pol}({\mathbf{A}}_{l}). For proving this, we let

Q:={𝒂𝒏𝒑2𝒍𝒔[𝒍]:𝒂(𝒔)=0}Q:=\{\mathbfsl{a}\in\mathbb{Z}_{np^{2}}^{l}\mid\exists s\in[{l}]:\mathbfsl{a}(s)=0\}

and show that Vi(Q)Vl(Q)\text{V}_{i}(Q)\neq\text{V}_{l}(Q); as a consequence Pol(𝐀i)≁algPol(𝐀l)\operatorname{Pol}({\mathbf{A}}_{i})\not\sim_{\text{alg}}\operatorname{Pol}({\mathbf{A}}_{l}).

First we prove that (1,,1)Vl(Q)(1,\dots,1)\notin\text{V}_{l}(Q). To this end, we consider the term equation fl(x1,,xl)0f_{l}(x_{1},\dots,x_{l})\approx 0. We have that fl(1,,1)=npf_{l}(1,\dots,1)=np and fl|Q=0f_{l}|_{Q}=0. Hence Lemma 2.1 yields (1,,1)Vl(Q)(1,\dots,1)\not\in\text{V}_{l}(Q).

We now prove that Vi(Q)=np2l\text{V}_{i}(Q)=\mathbb{Z}_{np^{2}}^{l}. To this end, let us consider an equation of the form t1(x1,,xl)t2(x1,,xl)t_{1}(x_{1},\dots,x_{l})\approx t_{2}(x_{1},\dots,x_{l}) with t1,t2Poll(𝐀i)t_{1},t_{2}\in\operatorname{Pol}_{l}({\mathbf{A}}_{i}). This equation is equivalent to (t1t2)(x1,,xl)0(t_{1}-t_{2})\,(x_{1},\dots,x_{l})\approx 0. Hence it suffices to consider equations of the form t(x1,,xl)0t(x_{1},\dots,x_{l})\approx 0. If t(x1,,xl)0t(x_{1},\dots,x_{l})\approx 0 is satisfied by all elements of QQ, then the function tt is absorbing and has arity greater than ii. Lemma 3.1 implies that tt is the constant zero function. Therefore, the equation t(x1,,xl)0t(x_{1},\dots,x_{l})\approx 0 is satisfied by all elements in np2l\mathbb{Z}_{np^{2}}^{l}. Hence Lemma 2.1 yields (1,1)Vi(Q)(1,\dots 1)\in\text{V}_{i}(Q) and therefore Vi(Q)Vl(Q)\text{V}_{i}(Q)\neq\text{V}_{l}(Q).

We conclude that (Pol(𝐀i))i{1}(\operatorname{Pol}({\mathbf{A}}_{i}))_{i\in\mathbb{N}\setminus\{1\}} is an infinite family of algebraically inequivalent clones. ∎

4. Countably many algebraically inequivalent constantive Mal’cev clones on finite sets with at least 4 elements

We report a construction from [12, Section 3]: Let AA be a finite set such that |A|4\lvert A\rvert\geq 4 and let 𝒞\mathcal{C} be a constantive clone on AA with a Mal’cev term pp. The construction picks an element 1A1\in A and an element 0A0\notin A and constructs a new clone 𝒞\mathcal{C}^{\oplus} on the set A:=A{0}A^{\oplus}:=A\cup\{0\}.

For all kk\in\mathbb{N} and for all f𝒞f\in\mathcal{C}, the operation f:(A)kAf^{\oplus}\colon(A^{\oplus})^{k}\to A^{\oplus} is defined by

f(x1,,xk):={f(x1,,xk) if (x1,,xk)Ak,0 otherwise.f^{\oplus}(x_{1},\dots,x_{k}):=\begin{cases}f(x_{1},\dots,x_{k})&\text{ if }(x_{1},\dots,x_{k})\in A^{k},\\ 0&\text{ otherwise.}\end{cases}

A binary operation \cdot on AA^{\oplus} is defined by

xy:={0 if (x,y)A2,1 if x=y=0,x if xA and y=0,y if x=0 and yA.x\cdot y:=\begin{cases}0&\text{ if }(x,y)\in A^{2},\\ 1&\text{ if }x=y=0,\\ x&\text{ if }x\in A\text{ and }y=0,\\ y&\text{ if }x=0\text{ and }y\in A.\end{cases}

Finally, 𝒞\mathcal{C}^{\oplus} is defined as the clone on AA^{\oplus} generated by {ff𝒞}{}\{f^{\oplus}\mid f\in\mathcal{C}\}\cup\{\cdot\} and all unary constant operations.

Lemma 4.1.

Let AA be a finite set with |A|4\lvert A\rvert\geq 4, let 𝒞\mathcal{C} be a constantive clone on AA and let f𝒞f\in\mathcal{C}^{\oplus}. Then

(4.1) f|Ak=0 or there exists f^𝒞 such that f|Ak=f^.f|_{A^{k}}=0\text{ or there exists }\hat{f}\in\mathcal{C}\text{ such that }f|_{A^{k}}=\hat{f}.
Proof.

Let KK be the set of all unary constant functions on AA^{\oplus}. We will show that all projections on AA^{\oplus} satisfy (4.1), and that for all k,nk,n\in\mathbb{N}, for all nn-ary

s{ff𝒞}{}Ks\in\{f^{\oplus}\mid f\in\mathcal{C}\}\cup\{\cdot\}\cup K

and for all kk-ary t1,,tnt_{1},\ldots,t_{n} that satisfy (4.1), also s(t1,,tn)s(t_{1},\ldots,t_{n}) satisfies (4.1). From this, we conclude that every function in 𝒞\mathcal{C}^{\oplus} satisfies (4.1).

Clearly, all the projections satisfy (4.1).

Let k,nk,n\in\mathbb{N}, let t1,,tnt_{1},\ldots,t_{n} be kk-ary functions on AA^{\oplus} that satisfy (4.1), and let ss be an nn-ary function on AA^{\oplus} from {ff𝒞}{}K\{f^{\oplus}\mid f\in\mathcal{C}\}\cup\{\cdot\}\cup K. We will show that s(t1,,tn)s(t_{1},\ldots,t_{n}) satisfies (4.1).

We first consider the case that s=fs=f^{\oplus} with f𝒞[n]f\in\mathcal{C}^{[n]}. If there exists i[n]i\in[{n}] such that ti|Ak=0t_{i}|_{A^{k}}=0, then by the definition of ff^{\oplus}, f(t1,,tn)|Ak=0f^{\oplus}(t_{1},\dots,t_{n})|_{A^{k}}=0, and hence s(t1,,tn)s(t_{1},\ldots,t_{n}) satisfies (4.1). If for all i[n]i\in[{n}], there exists ti^𝒞[k]\hat{t_{i}}\in\mathcal{C}^{[k]} such that ti|Ak=ti^t_{i}|_{A^{k}}=\hat{t_{i}}, then by the definition of ff^{\oplus} we have

s(t1,,tn)|Ak=f(t1,,tn)|Ak=f(g1^,,gn^)𝒞.s(t_{1},\ldots,t_{n})|_{A^{k}}=f^{\oplus}(t_{1},\dots,t_{n})|_{A^{k}}=f(\hat{g_{1}},\dots,\hat{g_{n}})\in\mathcal{C}.

Therefore s(t1,,tn)s(t_{1},\ldots,t_{n}) satisfies (4.1).

Now we consider the case that ss is the function \cdot (and n=2n=2). Let t1,t2t_{1},t_{2} be kk-ary functions on AA^{\oplus} satisfying (4.1). We show that t1t2t_{1}\cdot t_{2} satisfies (4.1). Since t1,t2t_{1},t_{2} satisfy (4.1), we have to consider the following cases.

Case 1: t1|Ak=t2|Ak=0t_{1}|_{A^{k}}=t_{2}|_{A^{k}}=0: In this case, the definition of \cdot yields t1t2|Ak=1t_{1}\cdot t_{2}|_{A^{k}}=1. Since 𝒞\mathcal{C} is constantive, t1t2|Ak𝒞t_{1}\cdot t_{2}|_{A^{k}}\in\mathcal{C}, and thus t1t2t_{1}\cdot t_{2} satisfies (4.1).

Case 2: t1|Ak=0t_{1}|_{A^{k}}=0 and there exists t2^𝒞[k]\hat{t_{2}}\in\mathcal{C}^{[k]} such that t2|Ak=t2^t_{2}|_{A^{k}}=\hat{t_{2}}: In this case, the definition of \cdot yields t1t2|Ak=t2^t_{1}\cdot t_{2}|_{A^{k}}=\hat{t_{2}}. This implies that t1t2|Ak𝒞t_{1}\cdot t_{2}|_{A^{k}}\in\mathcal{C}, and thus t1t2t_{1}\cdot t_{2} satisfies (4.1).

Case 3: t2|Ak=0t_{2}|_{A^{k}}=0 and there exists t1^𝒞[k]\hat{t_{1}}\in\mathcal{C}^{[k]} such that t1|Ak=t1^t_{1}|_{A^{k}}=\hat{t_{1}}: In this case, the definition of \cdot yields t1t2|Ak=t1^t_{1}\cdot t_{2}|_{A^{k}}=\hat{t_{1}}. This implies that t1t2|Ak𝒞t_{1}\cdot t_{2}|_{A^{k}}\in\mathcal{C}, and thus t1t2t_{1}\cdot t_{2} satisfies (4.1).

Case 4: There exist g1^,g2^𝒞[k]\hat{g_{1}},\hat{g_{2}}\in\mathcal{C}^{[k]} such that g1|Ak=g1^g_{1}|_{A^{k}}=\hat{g_{1}} and g2|Ak=g2^g_{2}|_{A^{k}}=\hat{g_{2}}: In this case, the definition of \cdot yields t1t2|Ak=0t_{1}\cdot t_{2}|_{A^{k}}=0, thus t1t2t_{1}\cdot t_{2} satisfies (4.1).

Finally, let us consider the case sKs\in K (and n=1n=1). Let aAa\in A^{\oplus} be the unique function value of ss, and let tt be a kk-ary function on AA^{\oplus} that satisfies (4.1). Clearly, s(t)(x1,,xk)=as(t)(x_{1},\dots,x_{k})=a for all x1,,xkAx_{1},\dots,x_{k}\in A^{\oplus}. If aAa\in A, since 𝒞\mathcal{C} is constantive, we have that s(t)s(t) satisfies (4.1). If a=0a=0, then s(t)(x1,,xk)=0s(t)(x_{1},\dots,x_{k})=0 for all (x1,,xk)Ak(x_{1},\dots,x_{k})\in A^{k}. Thus, s(t)s(t) satisfies (4.1). ∎

Lemma 4.2.

Let AA be a finite set with |A|4\lvert A\rvert\geq 4, let 𝒞\mathcal{C} be a constantive clone on AA, let f(𝒞)[k]f\in(\mathcal{C}^{\oplus})^{[k]}, and let 𝐚,𝐛𝐀𝐤\mathbfsl{a},\mathbfsl{b}\in A^{k}. Then f(𝐚)=𝟎f(\mathbfsl{a})=0 if and only if f(𝐛)=𝟎f(\mathbfsl{b})=0.

Proof.

If f(𝒂)0f(\mathbfsl{a})\neq 0, then f|Ak0f|_{A^{k}}\neq 0. Thus by Lemma 4.1, there exists f^𝒞\hat{f}\in\mathcal{C} such that f|Ak=f^f|_{A^{k}}=\hat{f}, and therefore f(𝒃)=𝒇^(𝒃)𝑨f(\mathbfsl{b})=\hat{f}(\mathbfsl{b})\in A, which implies f(𝒃)0f(\mathbfsl{b})\neq 0. ∎

Lemma 4.3.

[12, Section 3] Let AA be a finite set with |A|4\lvert A\rvert\geq 4 and let 𝒞\mathcal{C} be a constantive Mal’cev clone on AA. Then 𝒞\mathcal{C}^{\oplus} is a constantive Mal’cev clone on a set of cardinality |A|+1\lvert A\rvert+1.

Proposition 4.4.

Let AA be a finite set with |A|4\lvert A\rvert\geq 4, and let 𝒞1,𝒞2\mathcal{C}_{1},\mathcal{C}_{2} be two constantive Mal’cev clones on AA. If 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2} are not algebraically equivalent, then 𝒞1\mathcal{C}_{1}^{\oplus} and 𝒞2\mathcal{C}_{2}^{\oplus} are not algebraically equivalent.

Proof.

Without loss of generality, we assume that there exists BAkB\subseteq A^{k} that is algebraic with respect to 𝒞1\mathcal{C}_{1} but not with respect to 𝒞2\mathcal{C}_{2}. Note that then BAkB\neq A^{k}. Since 𝒞2\mathcal{C}_{2} is constantive, BB is not empty. Then there exists 𝒂𝑨𝒌𝑩\mathbfsl{a}\in A^{k}\setminus B such that 𝒂V𝒞2(𝑩)\mathbfsl{a}\in\text{V}_{\mathcal{C}_{2}}(B).

Let B0:=B((A)kAk)B_{0}:=B\cup((A^{\oplus})^{k}\setminus A^{k}). We claim that

(4.2) B0Alg𝒞1.B_{0}\in\operatorname{Alg}\mathcal{C}_{1}^{\oplus}.

To prove (4.2), let 𝒄𝑩0\mathbfsl{c}\notin B_{0}. By Lemma 2.1 it suffices to show that there are two functions p1,p2𝒞1p_{1},p_{2}\in\mathcal{C}_{1}^{\oplus} such that p1|B0=p2|B0p_{1}|_{B_{0}}=p_{2}|_{B_{0}} and p1(𝒄)𝒑2(𝒄)p_{1}(\mathbfsl{c})\neq p_{2}(\mathbfsl{c}). Since 𝒄𝑨𝒌𝑩\mathbfsl{c}\in A^{k}\setminus B and BB is algebraic with respect to 𝒞1\mathcal{C}_{1}, Lemma 2.1 implies that there are f1,f2𝒞1[k]f_{1},f_{2}\in\mathcal{C}_{1}^{[k]} such that f1|B=f2|Bf_{1}|_{B}=f_{2}|_{B} and f1(𝒄)𝒇2(𝒄)f_{1}(\mathbfsl{c})\neq f_{2}(\mathbfsl{c}). We set p1=f1p_{1}=f_{1}^{\oplus} and p2=f2p_{2}=f_{2}^{\oplus}. Then p1|B0=p2|B0p_{1}|_{B_{0}}=p_{2}|_{B_{0}} and p1(𝒄)𝒑2(𝒄)p_{1}(\mathbfsl{c})\neq p_{2}(\mathbfsl{c}), which concludes the proof of (4.2).

We now claim that

(4.3) 𝒂V𝒞2(𝑩0).\mathbfsl{a}\in\text{V}_{\mathcal{C}_{2}^{\oplus}}(B_{0}).

Suppose that 𝒂V𝒞2(𝑩0)\mathbfsl{a}\notin\text{V}_{\mathcal{C}_{2}^{\oplus}}(B_{0}). Then Lemma 2.1 implies that there are f1,f2(𝒞2)[k]f_{1},f_{2}\in(\mathcal{C}_{2}^{\oplus})^{[k]} such that f1|B0=f2|B0f_{1}|_{B_{0}}=f_{2}|_{B_{0}} and f1(𝒂)𝒇2(𝒂)f_{1}(\mathbfsl{a})\neq f_{2}(\mathbfsl{a}). Since BB0B\subseteq B_{0} we infer that f1|B=f2|Bf_{1}|_{B}=f_{2}|_{B}. From Lemma 4.2 and BB\neq\emptyset, we deduce that f1|B=f2|B0f_{1}|_{B}=f_{2}|_{B}\neq 0. Moreover, by Lemma 4.1, there are f1^,f2^𝒞2[k]\hat{f_{1}},\hat{f_{2}}\in\mathcal{C}_{2}^{[k]} such that f1|Ak=f1^f_{1}|_{A^{k}}=\hat{f_{1}} and f2|Ak=f2^f_{2}|_{A^{k}}=\hat{f_{2}}. Therefore, f1^|B=f2^|B\hat{f_{1}}|_{B}=\hat{f_{2}}|_{B} and f1^(𝒂)𝒇2^(𝒂)\hat{f_{1}}(\mathbfsl{a})\neq\hat{f_{2}}(\mathbfsl{a}). Hence by Lemma 2.1, 𝒂V𝒞2(𝑩)\mathbfsl{a}\notin\text{V}_{\mathcal{C}_{2}}(B), contradicting the choice of 𝒂\mathbfsl{a}.

From (4.2) and (4.3) we obtain Alg𝒞1Alg𝒞2\operatorname{Alg}\mathcal{C}_{1}^{\oplus}\neq\operatorname{Alg}\mathcal{C}_{2}^{\oplus}. ∎

Proof of Theorem 1.2.

We proceed by induction on n=|A|n=\lvert A\rvert. For the base case n=4n=4, we use Theorem 1.3 with n=1n=1 and p=2p=2. For the induction step we let n4n\geq 4 and {𝒞ii}\{\mathcal{C}_{i}\mid i\in\mathbb{N}\} be a collection of pairwise algebraically inequivalent constantive Mal’cev clones on a set of cardinality nn. Then Proposition 4.4 and Lemma 4.3 imply that {𝒞ii}\{\mathcal{C}_{i}^{\oplus}\mid i\in\mathbb{N}\} is a collection of pairwise algebraically inequivalent constantive Mal’cev clones on a set of cardinality n+1n+1. ∎

5. Continuously many algebraically inequivalent clones on sets with at least 3 elements

Let Z={0,1,2}Z=\{0,1,2\}. For nn\in\mathbb{N} and for ini\leq n, a function f:ZnZf:Z^{n}\to Z depends on its ii-th argument if there exist 𝒂𝒁𝒏\mathbfsl{a}\in Z^{n} and bZb\in Z such that f(𝒂)𝒇(𝒂1,,𝒂𝒊1,𝒃,𝒂𝒊+1,,𝒂𝒏)f(\mathbfsl{a})\neq f(a_{1},\dots,a_{i-1},b,a_{i+1},\dots,a_{n}). If ff depends on exactly one of its arguments, then ff is called essentially unary. The essential arity of ff is defined by essArity(f):=|{i[n]:f depends on its i-th argument}|\text{essArity}(f):=\lvert\{i\in[{n}]:f\text{ depends on its }i\text{-th argument}\}\rvert. Let \mathcal{F} be a set of finitary functions on ZZ. We say that ff belongs to \mathcal{F} up to inessential arguments if there exist ll\in\mathbb{N}, 1i1<<iln1\leq i_{1}<\dots<i_{l}\leq n and a function g[l]g\in\mathcal{F}^{[l]} such that f(x1,,xn)=g(xi1,,xil)f(x_{1},\dots,x_{n})=g(x_{i_{1}},\dots,x_{i_{l}}) for all 𝒙𝒁𝒏\mathbfsl{x}\in Z^{n}.

Let D1=D_{1}=\emptyset and for each m{1}m\in\mathbb{N}\setminus\{1\}, let Dm:={(1,2,,2),,(2,,2,1)}D_{m}:=\{(1,2,\dots,2),\dots,(2,\dots,2,1)\}. For each nn\in\mathbb{N}, let FnF_{n} be the set consisting of all the functions f:ZnZf\colon Z^{n}\to Z such that

(5.1) f(Zn){0,1}, and\displaystyle f(Z^{n})\subseteq\{0,1\},\text{ and}
(5.2) f1({1})Dn.\displaystyle f^{-1}(\{1\})\subseteq D_{n}.

We set F:=nFnF:=\bigcup_{n\in\mathbb{N}}F_{n} and

F:=n{f:ZnZf belongs to F up to inessential arguments}𝒥Z.F^{\prime}:=\bigcup_{n\in\mathbb{N}}\{f\colon Z^{n}\to Z\mid f\text{ belongs to }F\text{ up to inessential arguments}\}\cup\mathcal{J}_{Z}.
Lemma 5.1.

The sets FF and FF^{\prime} satisfy the following properties:

  1. (1)

    For all fFf\in F with essArity(f)1\text{essArity}(f)\leq 1, we have f=0f=0.

  2. (2)

    For all nn\in\mathbb{N} and for all fFnf\in F_{n} with essArity(f)>0\text{essArity}(f)>0, we have that essArity(f)=n\text{essArity}(f)=n and n2n\geq 2.

  3. (3)

    For all fFf\in F^{\prime} with essArity(f)=1\text{essArity}(f)=1, we have f𝒥Zf\in\mathcal{J}_{Z}.

  4. (4)

    For all l,nl,n\in\mathbb{N}, for all gFlg\in F_{l} and for all ρ:[l][n]\rho\colon[{l}]\to[{n}], we have gρFng_{\rho}\in F^{\prime}_{n}.

  5. (5)

    For all k,nk,n\in\mathbb{N}, for all fFkf\in F^{\prime}_{k} and for all σ:[k][n]\sigma\colon[{k}]\to[{n}], we have fσFnf_{\sigma}\in F^{\prime}_{n}.

Proof.

(1) Let nn\in\mathbb{N}, let fFnf\in F_{n} with essArity(f)1\text{essArity}(f)\leq 1 and let 𝒙𝒁𝒏\mathbfsl{x}\in Z^{n}. We prove that f(𝒙)=0f(\mathbfsl{x})=0. If essArity(f)1\text{essArity}(f)\leq 1, then there exists i[n]i\in[{n}] such that for all j[n]{i}j\in[{n}]\setminus\{i\} the function ff does not depend on its jj-th argument. Let us define fZZf^{\prime}\in Z^{Z} by f(y)=f(y,,y)f^{\prime}(y)=f(y,\dots,y) for all yZy\in Z. We prove that f=0f^{\prime}=0. To this end, let aZa\in Z. Since (a,,a)Dn(a,\dots,a)\notin D_{n}, we have f(a)=f(a,,a)=0f^{\prime}(a)=f(a,\dots,a)=0. Moreover, since for all jij\neq i, ff does not depend on its jj-th argument, f(𝒙)=𝒇(𝒙𝒊,,𝒙𝒊)=𝒇(𝒙𝒊)=0f(\mathbfsl{x})=f(x_{i},\dots,x_{i})=f^{\prime}(x_{i})=0. This concludes the proof of (1).

(2) Let nn\in\mathbb{N} and let fFnf\in F_{n}. If ff is not constantly zero, then there exists 𝒂𝑫𝒏\mathbfsl{a}\in D_{n} such that f(𝒂)=1f(\mathbfsl{a})=1. Changing any one of the coordinates of 𝒂\mathbfsl{a} to 0, the value of ff changes from 11 to 0. Hence ff depends on all of its arguments and (1) implies that it cannot be essentially unary. This concludes the proof of (2).

(3) Let fFnf\in F^{\prime}_{n} with essArity(f)=1\text{essArity}(f)=1. Seeking a contradiction, we suppose that ff is not a projection. Then ff belongs to FF up to inessential arguments. Thus, there exist 1j1<<jkn1\leq j_{1}<\dots<j_{k}\leq n and f^F\hat{f}\in F such that f(x1,,xn)=f^(xj1,,xjk)f(x_{1},\dots,x_{n})=\hat{f}(x_{j_{1}},\dots,x_{j_{k}}) for all 𝒙𝒁𝒏\mathbfsl{x}\in Z^{n}. If essArity(f^)1\text{essArity}(\hat{f})\leq 1, then (1) yields f^=0\hat{f}=0, and so ff is the constant 0-function, contradicting essArity(f)=1\text{essArity}(f)=1. If essArity(f^)>1\text{essArity}(\hat{f})>1, then essArity(f)>1\text{essArity}(f)>1, contradicting essArity(f)=1\text{essArity}(f)=1. This concludes the proof of (3).

(4) Let us assume that the image of ρ\rho is {j1,,jk}[n]\{j_{1},\dots,j_{k}\}\subseteq[{n}], with j1<<jkj_{1}<\dots<j_{k}, and let us define

h:={((xj1,,xjk),g(xρ(1),,xρ(l)))(x1,,xn)Zn}.h:=\Set{\bigl{(}(x_{j_{1}},\dots,x_{j_{k}}),g(x_{\rho(1)},\dots,x_{\rho(l)})\bigr{)}\mid(x_{1},\dots,x_{n})\in Z^{n}}.

We first prove that hh is a functional relation. To this end, let 𝒙,𝒚𝒁𝒏\mathbfsl{x},\mathbfsl{y}\in Z^{n} be such that (xj1,,xjk)=(yj1,,yjk)(x_{j_{1}},\dots,x_{j_{k}})=(y_{j_{1}},\dots,y_{j_{k}}). We show that g(xρ(1),,xρ(l))=g(yρ(1),,yρ(l))g(x_{\rho(1)},\dots,x_{\rho(l)})=g(y_{\rho(1)},\dots,y_{\rho(l)}). Clearly, if (xj1,,xjk)=(yj1,,yjk)(x_{j_{1}},\dots,x_{j_{k}})=(y_{j_{1}},\dots,y_{j_{k}}), then (xρ(1),,xρ(l))=(yρ(1),,yρ(l))(x_{\rho(1)},\ldots,x_{\rho(l)})=(y_{\rho(1)},\ldots,y_{\rho(l)}), and therefore g(xρ(1),,xρ(l))=g(yρ(1),,yρ(l))g(x_{\rho(1)},\dots,x_{\rho(l)})=g(y_{\rho(1)},\dots,y_{\rho(l)}). This concludes the proof that hh is a functional relation. Now we prove that hFkh\in F_{k}. Condition (5.1) is clearly satisfied since gFlg\in F_{l}. We prove that hh satisfies (5.2). To this end, let 𝒙𝑨𝒏\mathbfsl{x}\in A^{n} with g(xρ(1),,xρ(l))=1g(x_{\rho(1)},\dots,x_{\rho(l)})=1. We show that (xj1,,xjk)Dk(x_{j_{1}},\dots,x_{j_{k}})\in D_{k}. Since gFlg\in F_{l} and g(xρ(1),,xρ(l))=1g(x_{\rho(1)},\dots,x_{\rho(l)})=1, there exists a[l]a\in[{l}] such that xρ(a)=1x_{\rho(a)}=1 and xρ(b)=2x_{\rho(b)}=2 for all b[l]{a}b\in[{l}]\setminus\{a\}. Since j1<<jkj_{1}<\dots<j_{k} we have |{r[k]:jr=ρ(a)}|1\lvert\{r\in[{k}]\colon j_{r}=\rho(a)\}\rvert\leq 1. Moreover, since {j1,,jk}\{j_{1},\dots,j_{k}\} is the image of ρ\rho, we have |{r[k]:jr=ρ(a)}|1\lvert\{r\in[{k}]\colon j_{r}=\rho(a)\}\rvert\geq 1. Thus, |{r[k]:jr=ρ(a)}|=1\lvert\{r\in[{k}]\colon j_{r}=\rho(a)\}\rvert=1. Hence there exists i[k]i\in[{k}] such that xji=xρ(a)=1x_{j_{i}}=x_{\rho(a)}=1. Then for each r[k]{i}r\in[{k}]\setminus\{i\} there exists b[l]{a}b\in[{l}]\setminus\{a\} such that jr=ρ(b)j_{r}=\rho(b), and then xjr=xρ(b)=2x_{j_{r}}=x_{\rho(b)}=2. Thus, (xj1,,xjk)Dk(x_{j_{1}},\dots,x_{j_{k}})\in D_{k}, and so hh satisfies (5.2). This proves that hFkh\in F_{k}. Since j1<<jkj_{1}<\dots<j_{k} and gρ(𝒂)=𝒉(𝒂𝒋1,,𝒂𝒋𝒌)g_{\rho}(\mathbfsl{a})=h(a_{j_{1}},\dots,a_{j_{k}}) for all 𝒂𝑨𝒏\mathbfsl{a}\in A^{n}, gρg_{\rho} belongs to FF up to inessential arguments. Hence gρFng_{\rho}\in F_{n}^{\prime}. This concludes the proof of (4).

(5) Let k,nk,n\in\mathbb{N}, let fFkf\in F^{\prime}_{k} and let σ:[k][n]\sigma\colon[{k}]\to[{n}]. We prove that fσFnf_{\sigma}\in F^{\prime}_{n}. By the definition of FF^{\prime}, there exist ll\in\mathbb{N}, gFlg\in F_{l} and τ:[l][k]\tau\colon[{l}]\to[{k}] injective and increasing such that f(y1,,yk)=g(yτ(1),,yτ(l))f(y_{1},\dots,y_{k})=g(y_{\tau(1)},\dots,y_{\tau(l)}) for all 𝒚𝒁𝒌\mathbfsl{y}\in Z^{k}. Therefore, for all 𝒙𝒁𝒏\mathbfsl{x}\in Z^{n} we have

fσ(x1,,xn)=f(xσ(1),,xσ(k))=g(xσ(τ(1)),,xσ(τ(l))).f_{\sigma}(x_{1},\dots,x_{n})=f(x_{\sigma(1)},\dots,x_{\sigma(k)})=g(x_{\sigma(\tau(1))},\dots,x_{\sigma(\tau(l))}).

Since gFlg\in F_{l}, (4) implies that gστFng_{\sigma\circ\tau}\in F^{\prime}_{n}, and thus fσFnf_{\sigma}\in F^{\prime}_{n}. This concludes the proof of (5). ∎

Lemma 5.2.

The set FF^{\prime} is a clone.

Proof.

We prove that FF^{\prime} is closed under ζ\zeta, τ\tau, Δ\Delta, \nabla and * as they are defined in [14] (cf. [9, 19]). To this end, let fFnf\in F^{\prime}_{n}. Lemma 5.1(5) yields that 𝜁fFn\mathop{\zeta}f\in F^{\prime}_{n}, 𝜏fFn\mathop{\tau}f\in F^{\prime}_{n}, ΔfFn1\mathop{\Delta}f\in F^{\prime}_{n-1}, and fFn+1\mathop{\nabla}f\in F^{\prime}_{n+1}. Let gFmg\in F^{\prime}_{m} and let hh be the function defined by h(𝒙)=(𝒇𝒈)(𝒙)=𝒇(𝒈(𝒙1,,𝒙𝒎),𝒙𝒎+1,,𝒙𝒎+𝒏1)h(\mathbfsl{x})=(f*g)(\mathbfsl{x})=f(g(x_{1},\dots,x_{m}),x_{m+1},\dots,x_{m+n-1}) for all 𝒙𝒁𝒎+𝒏1\mathbfsl{x}\in Z^{m+n-1}. We prove that hh belongs to FF^{\prime}. If ff is constant, then hh is a constant mapping with the same function value as ff. Hence hh is a minor of ff, and thus by Lemma 5.1(5), hFh\in F^{\prime}. If ff is a projection, then either hh is a projection or h=gh=g. In both cases hh belongs to FF^{\prime}. If gg is a projection, then hh is a minor of ff. Hence Lemma 5.1(5) yields hFh\in F^{\prime}. Finally, let us assume that both ff and gg belong to FF up to inessential arguments and ff is not constant. Then there exist 1j1<<jnn1\leq j_{1}<\dots<j_{n^{\prime}}\leq n, 1i1<<imm1\leq i_{1}<\dots<i_{m^{\prime}}\leq m, f^Fn\hat{f}\in F_{n^{\prime}} and g^Fm\hat{g}\in F_{m^{\prime}} such that for all 𝒙𝒁𝒏\mathbfsl{x}\in Z^{n} and for all 𝒚𝒁𝒎\mathbfsl{y}\in Z^{m} we have f(𝒙)=𝒇^(𝒙𝒋1,,𝒙𝒋𝒏)f(\mathbfsl{x})=\hat{f}(x_{j_{1}},\dots,x_{j_{n^{\prime}}}) and g(𝒚)=𝒈^(𝒚𝒊1,,𝒚𝒊𝒎)g(\mathbfsl{y})=\hat{g}(y_{i_{1}},\dots,y_{i_{m^{\prime}}}). If j12j_{1}\geq 2, then for all 𝒙𝒁𝒎+𝒏1\mathbfsl{x}\in Z^{m+n-1} we have h(𝒙)=𝒇^(𝒙𝒎+𝒋11,,𝒙𝒎+𝒋𝒏1)h(\mathbfsl{x})=\hat{f}(x_{m+j_{1}-1},\dots,x_{m+j_{n^{\prime}}-1}). Thus, hh is a minor of f^\hat{f} and Lemma 5.1(4) yields hFh\in F^{\prime}. Let us now assume that j1=1j_{1}=1 and let h^\hat{h} be defined by h^(x1,,xm+n1)=f^(g^(x1,,xm),xm+1,,xm+n1)\hat{h}(x_{1},\dots,x_{m^{\prime}+n^{\prime}-1})=\hat{f}(\hat{g}(x_{1},\dots,x_{m^{\prime}}),x_{m^{\prime}+1},\dots,x_{m^{\prime}+n^{\prime}-1}) for all 𝒙𝒁𝒎+𝒏1\mathbfsl{x}\in Z^{m^{\prime}+n^{\prime}-1}. We now prove that h^F\hat{h}\in F. Since f^\hat{f} belongs to FF, h^\hat{h} satisfies (5.1). We now prove that h^\hat{h} satisfies (5.2). To this end, let 𝒂𝒁𝒎+𝒏1\mathbfsl{a}\in Z^{m^{\prime}+n^{\prime}-1} be such that h^(𝒂)=1\hat{h}(\mathbfsl{a})=1. Then, since f^\hat{f} satisfies (5.2), we infer that (g(a1,,am),am+1,,am+n1)Dn(g(a_{1},\dots,a_{m^{\prime}}),a_{m^{\prime}+1},\dots,a_{m^{\prime}+n^{\prime}-1})\in D_{n^{\prime}}. Moreover, since gg satisfies (5.1) and (5.2), we have that (a1,,am)Dm(a_{1},\dots,a_{m^{\prime}})\in D_{m^{\prime}} and (am+1,,am+n1)=(2,,2)(a_{m^{\prime}+1},\dots,a_{m^{\prime}+n^{\prime}-1})=(2,\dots,2). Thus, there exists r[m]r\in[{m^{\prime}}] such that ar=1a_{r}=1 and for all k[m+n1]{r}k\in[{m^{\prime}+n^{\prime}-1}]\setminus\{r\} we have ak=2a_{k}=2. This proves that 𝒂𝑫𝒎+𝒏1\mathbfsl{a}\in D_{m^{\prime}+n^{\prime}-1}. Thus, h^\hat{h} satisfies (5.2), and so h^F\hat{h}\in F. Moreover, we have that for all 𝒙𝒁𝒎+𝒏1\mathbfsl{x}\in Z^{m+n-1}

h(x1,,xm+n1)=f^(g^(xi1,,xim),xm+j21,,xm+jn1).h(x_{1},\dots,x_{m+n-1})=\hat{f}(\hat{g}(x_{i_{1}},\dots,x_{i_{m^{\prime}}}),x_{m+j_{2}-1},\dots,x_{m+j_{n^{\prime}}-1}).

Therefore hh belongs to FF up to inessential arguments, and so hFh\in F^{\prime}. Therefore, FF^{\prime} is closed under *. ∎

We make use of the following construction from [13] (cf. [19, Chapter 3]). For each i{1}i\in\mathbb{N}\setminus\{1\}, the function gi:ZiZg_{i}\colon Z^{i}\to Z is defined by

gi(x1,,xi)={1 if (x1,,xi)Di,0 otherwise.g_{i}(x_{1},\dots,x_{i})=\begin{cases}1&\text{ if }(x_{1},\dots,x_{i})\in D_{i},\\ 0&\text{ otherwise}.\end{cases}

For each I{1}I\subseteq\mathbb{N}\setminus\{1\}, 𝐙I{\mathbf{Z}}_{I} is defined as (Z;(gi)iI)(Z;(g_{i})_{i\in I}).

Lemma 5.3.

Let I{1}I\subseteq\mathbb{N}\setminus\{1\}, let kk\in\mathbb{N} and let gClok(𝐙I)g\in\operatorname{Clo}_{k}({\mathbf{Z}}_{I}). Then we have:

  1. (1)

    If gg is constant, then g=0g=0.

  2. (2)

    If gg is essentially unary, then gg is a projection.

  3. (3)

    If gg depends exactly on the arguments i1<<ili_{1}<\dots<i_{l} with l2l\geq 2, then the following two conditions are satisfied:

    1. (a)

      For all 𝒃𝒁𝒌\mathbfsl{b}\in Z^{k} we have g(𝒃)2g(\mathbfsl{b})\neq 2.

    2. (b)

      For all 𝒂𝒁𝒌\mathbfsl{a}\in Z^{k} with (ai1,,ail)Dl(a_{i_{1}},\dots,a_{i_{l}})\notin D_{l}, we have g(𝒂)=0g(\mathbfsl{a})=0.

Proof.

Equations (5.1) and (5.2) imply that for all i{1}i\in\mathbb{N}\setminus\{1\}, giFg_{i}\in F. Thus, Lemma 5.2 yields that for all I{1}I\subseteq\mathbb{N}\setminus\{1\}, Clo(𝐙I)F\operatorname{Clo}({\mathbf{Z}}_{I})\subseteq F^{\prime}. Since gClok(𝐙I)Fg\in\operatorname{Clo}_{k}({\mathbf{Z}}_{I})\subseteq F^{\prime}, then either g𝒥Zg\in\mathcal{J}_{Z} or gg belongs to FF up to inessential arguments. If gg is a constant, then it cannot be a projection. Hence it belongs to FF up to inessential arguments. Thus, by Lemma 5.1(1), is the constant zero. This proves (1). We now prove (2). Let us assume that gg is essentially unary. Then by Lemma 5.1(3), gg is a projection. This proves (2). We now prove (3). Let us assume that gg depends on its arguments i1<<ili_{1}<\dots<i_{l}, let 𝒃𝒁𝒌\mathbfsl{b}\in Z^{k} and let 𝒂𝒁𝒌\mathbfsl{a}\in Z^{k} be such that (ai1,,ail)Dl(a_{i_{1}},\dots,a_{i_{l}})\notin D_{l}. Since l2l\geq 2, gg is not a projection. Since gFg\in F^{\prime}, gg belongs to FF up to inessential arguments. This, together with Lemma 5.1(2), implies that there exists g^Fl\hat{g}\in F_{l} such that g(x1,,xk)=g^(xi1,,xil)g(x_{1},\dots,x_{k})=\hat{g}(x_{i_{1}},\dots,x_{i_{l}}) for all 𝒙𝒁𝒌\mathbfsl{x}\in Z^{k}, and g^\hat{g} depends on all of its arguments. Since g^F\hat{g}\in F, (5.1) implies that g(𝒃)=𝒈^(𝒃𝒊1,,𝒃𝒊𝒍){0,1}g(\mathbfsl{b})=\hat{g}(b_{i_{1}},\dots,b_{i_{l}})\in\{0,1\}. Moreover, since (ai1,,ail)Dl(a_{i_{1}},\dots,a_{i_{l}})\notin D_{l}, (5.2) and (5.1) yield g(𝒂)=𝒈^(𝒂𝒊1,,𝒂𝒊𝒍)=0g(\mathbfsl{a})=\hat{g}(a_{i_{1}},\dots,a_{i_{l}})=0. This proves (3). ∎

Proposition 5.4.

Let I{1}I\subseteq\mathbb{N}\setminus\{1\}, let i{1}i\in\mathbb{N}\setminus\{1\} and let gi={(x1,,xi,xi+1)Zi+1xi+1=g(x1,,xi)}g_{i}^{\bullet}=\{(x_{1},\dots,x_{i},x_{i+1})\in Z^{i+1}\mid x_{i+1}=g(x_{1},\dots,x_{i})\}. Then giAlg𝐙Ig_{i}^{\bullet}\in\operatorname{Alg}{\mathbf{Z}}_{I} if and only if iIi\in I.

Proof.

If iIi\in I, then giAlg𝐙Ig_{i}^{\bullet}\in\operatorname{Alg}{\mathbf{Z}}_{I} by definition. We now assume that giAlg𝐙Ig_{i}^{\bullet}\in\operatorname{Alg}{\mathbf{Z}}_{I} and prove that iIi\in I.

If giAlg𝐙Ig_{i}^{\bullet}\in\operatorname{Alg}{\mathbf{Z}}_{I}, since (1,,1)gi(1,\dots,1)\notin g_{i}^{\bullet}, Lemma 2.1 implies that there exist f1,f2Cloi+1(𝐙I)f_{1},f_{2}\in\operatorname{Clo}_{i+1}({\mathbf{Z}}_{I}) such that f1|gi=f2|gif_{1}|_{g_{i}^{\bullet}}=f_{2}|_{g_{i}^{\bullet}} and f1(1,,1)f2(1,,1)f_{1}(1,\dots,1)\neq f_{2}(1,\dots,1). We now prove that one of the two is a projection and the other depends on at least two of its arguments. Seeking contradictions, let us suppose that this is not the case. We distinguish cases accordingly to the essential arity of f1f_{1} and f2f_{2}.

Case 1: f1f_{1} and f2f_{2} are constant: In this case f1(1,,1)=f2(1,,1)f_{1}(1,\dots,1)=f_{2}(1,\dots,1).

Case 2: f1f_{1} is constant and f2f_{2} is essentially unary: In this case, Lemma 5.3 yields that f1f_{1} is the constant 0 function and f2f_{2} is a projection. Let 𝒂=(1,2,,2,1)\mathbfsl{a}=(1,2,\dots,2,1). Since g(a1,,ai)=1g(a_{1},\dots,a_{i})=1, 𝒂𝒈𝒊\mathbfsl{a}\in g_{i}^{\bullet}. Lemma 5.3 yields f1(𝒂)=0f_{1}(\mathbfsl{a})=0 and f2(𝒂){1,2}f_{2}(\mathbfsl{a})\in\{1,2\}. Hence f1(𝒂)𝒇2(𝒂)f_{1}(\mathbfsl{a})\neq f_{2}(\mathbfsl{a}). Thus, f1|gif2|gif_{1}|_{g_{i}^{\bullet}}\neq f_{2}|_{g_{i}^{\bullet}}.

Case 3: f1f_{1} is essentially unary and f2f_{2} is constant: This case is symmetric to Case 2.

Case 4: f1f_{1} is constant and f2f_{2} depends on at least two of its arguments: In this case Lemma 5.3 yields f1(1,,1)=0=f2(1,,1)f_{1}(1,\dots,1)=0=f_{2}(1,\dots,1).

Case 5: f1f_{1} is depends on at least two of its arguments and f2f_{2} is constant: This case is symmetric to Case 4.

Case 6: f1f_{1} and f2f_{2} are both essentially unary: In this case Lemma 5.3(2) implies that both f1f_{1} and f2f_{2} are projections. Thus f1(1,,1)=1=f2(1,,1)f_{1}(1,\dots,1)=1=f_{2}(1,\dots,1).

Case 7: f1f_{1} and f2f_{2} both depend on at least two of their arguments: In this case Lemma 5.3(3) yields f1(1,,1)=0=f2(1,,1)f_{1}(1,\dots,1)=0=f_{2}(1,\dots,1).

Thus, we have proved that one among f1f_{1} and f2f_{2} is a projection, while the other depends on at least two of its arguments. Without loss of generality, let us assume that f2f_{2} is a projection and that f1f_{1} depends on at least two of its arguments. Let

T:={𝒂𝒁𝒊+1𝒂𝒊+1=1 and (𝒂1,,𝒂𝒊)𝑫𝒊}.T:=\{\mathbfsl{a}\in Z^{i+1}\mid a_{i+1}=1\text{ and }(a_{1},\dots,a_{i})\in D_{i}\}.

Note that TgiT\subseteq g_{i}^{\bullet}. Let j{1,,i,i+1}j\in\{1,\dots,i,i+1\} be such that f2(x1,,xi+1)=xjf_{2}(x_{1},\dots,x_{i+1})=x_{j} for all 𝒙𝒁𝒊+1\mathbfsl{x}\in Z^{i+1}. We claim j=i+1j=i+1. Seeking a contradiction, let us suppose that jij\leq i. Let 𝒙𝑻\mathbfsl{x}\in T be such that 𝒙(𝒋)=2\mathbfsl{x}(j)=2. Then, since f1|T=f2|Tf_{1}|_{T}=f_{2}|_{T}, we have f1(𝒙)=2f_{1}(\mathbfsl{x})=2. On the other hand, Lemma 5.3(3) implies that f1(𝒙)2f_{1}(\mathbfsl{x})\neq 2. Thus we get the desired contradiction and deduce that f2f_{2} is the (i+1)(i+1)-th projection. Next, we prove that f1f_{1} does not depend on its (i+1)(i+1)-th argument. Seeking a contradiction, let us suppose that f1f_{1} depends on its (i+1)(i+1)-th argument. Then, since f1f_{1} depends on at least two of its arguments, there exists j[i]j\in[{i}] such that f1f_{1} depends also on its jj-th argument. Let 𝒂𝑻\mathbfsl{a}\in T be such that 𝒂(𝒋)=1\mathbfsl{a}(j)=1. Since f1|T=f2|Tf_{1}|_{T}=f_{2}|_{T}, we have that f1(𝒂)=𝒇2(𝒂)=𝒂𝒊+1=1f_{1}(\mathbfsl{a})=f_{2}(\mathbfsl{a})=a_{i+1}=1. On the other hand, Lemma 5.3 implies that f1(𝒂)=0f_{1}(\mathbfsl{a})=0 because aj=ai+1=1a_{j}=a_{i+1}=1. From this contradiction, we conclude that f1f_{1} does not depend on its (i+1)(i+1)-th argument.

Let f:ZiZf\colon Z^{i}\to Z be defined by f(x1,xi):=f1(x1,,xi,x1)f(x_{1},\dots x_{i}):=f_{1}(x_{1},\dots,x_{i},x_{1}). Since f1f_{1} does not depend on its (i+1)(i+1)-th argument, for each 𝒂𝒈𝒊\mathbfsl{a}\in g_{i}^{\bullet}, we have

(5.3) f(a1,,ai)=f1(a1,,ai,a1)=f1(a1,,ai+1)=f2(a1,,ai+1)=ai+1.f(a_{1},\dots,a_{i})=f_{1}(a_{1},\dots,a_{i},a_{1})=f_{1}(a_{1},\dots,a_{i+1})=f_{2}(a_{1},\dots,a_{i+1})=a_{i+1}.

We show that gi=fg_{i}=f. To this end, let 𝒂𝒁𝒊\mathbfsl{a}\in Z^{i}. Then (a1,,ai,gi(𝒂))𝒈𝒊(a_{1},\dots,a_{i},g_{i}(\mathbfsl{a}))\in g_{i}^{\bullet}, and thus (5.3) yields f(𝒂)=𝒈𝒊(𝒂)f(\mathbfsl{a})=g_{i}(\mathbfsl{a}). Since fCloi(𝐙I)f\in\operatorname{Clo}_{i}({\mathbf{Z}}_{I}), we deduce that giCloi(𝐙I)g_{i}\in\operatorname{Clo}_{i}({\mathbf{Z}}_{I}). We now apply [19, Theorem 3.1.4] and deduce that iIi\in I. ∎

Corollary 5.5.

On the set Z={0,1,2}Z=\{0,1,2\} there are exactly 202^{\aleph_{0}} algebraically inequivalent clones.

Proof.

Let Z\mathcal{L}_{Z} be the set of all clones on ZZ and let ψ:𝒫({1})Z/alg\psi\colon\mathcal{P}(\mathbb{N}\setminus\{1\})\to\mathcal{L}_{Z}/{\operatorname{\sim_{\text{alg}}}} be defined by IClo(𝐙I)/algI\mapsto\operatorname{Clo}({\mathbf{Z}}_{I})/{\operatorname{\sim_{\text{alg}}}}. Proposition 5.4 yields that for all IJ𝒫({1})I\neq J\in\mathcal{P}(\mathbb{N}\setminus\{1\}) we have Clo(𝐙I)algClo(𝐙J)\operatorname{Clo}({\mathbf{Z}}_{I})\operatorname{\sim_{\text{alg}}}\operatorname{Clo}({\mathbf{Z}}_{J}) if and only if I=JI=J. Therefore, ψ\psi is injective. Hence there are at least 202^{\aleph_{0}} algebraically inequivalent clones on ZZ. Since by [19, Theorem 3.1.4] there are at most 202^{\aleph_{0}} clones on ZZ, we can conclude that there are exactly 202^{\aleph_{0}} algebraically inequivalent clones on ZZ. ∎

Proof of Theorem 1.1.

We first show that for a set AA and an element uAu\notin A, A{u}A\cup\{u\} has at least as many algebraically inequivalent clones as AA. To this end, let AA be a finite set with |A|3\lvert A\rvert\geq 3, let A\mathcal{L}_{A} be the set of all clones on AA, let uAu\notin A, let A=A{u}A^{\oplus}=A\cup\{u\}, let A\mathcal{L}_{A^{\oplus}} be the set of all clones on AA^{\oplus}, and let Φ:AA\Phi\colon\mathcal{L}_{A}\to\mathcal{L}_{A^{\oplus}} be defined by

Φ(𝒞)=n{f:(A)nAf|An𝒞}.\Phi(\mathcal{C})=\underset{n\in\mathbb{N}}{\bigcup}\{f:(A^{\oplus})^{n}\to A^{\oplus}\,\mid\,f|_{A^{n}}\in\mathcal{C}\}.

We first prove that for all 𝒞A\mathcal{C}\in\mathcal{L}_{A}, Φ(𝒞)\Phi(\mathcal{C}) is a clone on AA^{\oplus}. Clearly, all the projections belong to Φ(𝒞)\Phi(\mathcal{C}). Let fΦ(𝒞)[n]f\in\Phi(\mathcal{C})^{[n]}, let t1,,tnΦ(𝒞)[k]t_{1},\dots,t_{n}\in\Phi(\mathcal{C})^{[k]}. Then

f(t1,,tn)|Ak=f(t1|Ak,,tn|Ak)=f|An(t1|Ak,,tn|Ak).f(t_{1},\dots,t_{n})|_{A^{k}}=f(t_{1}|_{A^{k}},\dots,t_{n}|_{A^{k}})=f|_{A^{n}}(t_{1}|_{A^{k}},\dots,t_{n}|_{A^{k}}).

The definition of Φ\Phi yields f|An,t1|Ak,,tn|Ak𝒞f|_{A^{n}},t_{1}|_{A^{k}},\dots,t_{n}|_{A^{k}}\in\mathcal{C}. Thus, f|An(t1|Ak,,tn|Ak)f|_{A^{n}}(t_{1}|_{A^{k}},\dots,t_{n}|_{A^{k}}) belongs to 𝒞\mathcal{C} and f(t1,,tn)Φ(𝒞)f(t_{1},\dots,t_{n})\in\Phi(\mathcal{C}). Therefore, Φ(𝒞)\Phi(\mathcal{C}) is a clone on AA^{\oplus}.

Next, we prove that Φ\Phi satisfies

(5.4) 𝒞A,n,B(A)n:BAlgΦ(𝒞)BAnAlg𝒞.\forall\mathcal{C}\in\mathcal{L}_{A},\forall n\in\mathbb{N},\forall B\subseteq(A^{\oplus})^{n}\colon B\in\operatorname{Alg}\Phi(\mathcal{C})\Leftrightarrow B\cap A^{n}\in\operatorname{Alg}\mathcal{C}.

For proving (5.4), let 𝒞A\mathcal{C}\in\mathcal{L}_{A}, let nn\in\mathbb{N}, let B(A)nB\subseteq(A^{\oplus})^{n} and let us assume that BAlgnΦ(𝒞)B\in\operatorname{Alg}_{n}\Phi(\mathcal{C}). We prove that BAnAlgn𝒞B\cap A^{n}\in\operatorname{Alg}_{n}\mathcal{C}. To this end, let 𝒃𝑨𝒏𝑩\mathbfsl{b}\in A^{n}\setminus B. By Lemma 2.1 it suffices to prove that there are f1^,f2^𝒞\hat{f_{1}},\hat{f_{2}}\in\mathcal{C} such that f1^|AnB=f2^|AnB\hat{f_{1}}|_{A^{n}\cap B}=\hat{f_{2}}|_{A^{n}\cap B} and f1^(𝒃)𝒇2^(𝒃)\hat{f_{1}}(\mathbfsl{b})\neq\hat{f_{2}}(\mathbfsl{b}). Since BAlgnΦ(𝒞)B\in\operatorname{Alg}_{n}\Phi(\mathcal{C}) and 𝒃𝑩\mathbfsl{b}\notin B, Lemma 2.1 yields that there exist f1,f2Φ(𝒞)f_{1},f_{2}\in\Phi(\mathcal{C}) such that f1|B=f2|Bf_{1}|_{B}=f_{2}|_{B} and f1(𝒃)𝒇2(𝒃)f_{1}(\mathbfsl{b})\neq f_{2}(\mathbfsl{b}). Let f1^=f1|An\hat{f_{1}}=f_{1}|_{A^{n}} and let f2^=f2|An\hat{f_{2}}=f_{2}|_{A^{n}}. Then clearly f1^,f2^𝒞\hat{f_{1}},\hat{f_{2}}\in\mathcal{C}, f1^|AnB=f2^|AnB\hat{f_{1}}|_{A^{n}\cap B}=\hat{f_{2}}|_{A^{n}\cap B} and f1^(𝒃)𝒇2^(𝒃)\hat{f_{1}}(\mathbfsl{b})\neq\hat{f_{2}}(\mathbfsl{b}). Thus, BAnAlg𝒞B\cap A^{n}\in\operatorname{Alg}\mathcal{C}.

Let us now assume that BAnAlg𝒞B\cap A^{n}\in\operatorname{Alg}\mathcal{C}. We prove that BAlgnΦ(𝒞)B\in\operatorname{Alg}_{n}\Phi(\mathcal{C}). Let 𝒂(𝑨)𝒏𝑩\mathbfsl{a}\in(A^{\oplus})^{n}\setminus B. By Lemma 2.1 it suffices to show that there are f1,f2Φ(𝒞)f_{1},f_{2}\in\Phi(\mathcal{C}) such that f1|B=f2|Bf_{1}|_{B}=f_{2}|_{B} and f1(𝒂)𝒇2(𝒂)f_{1}(\mathbfsl{a})\neq f_{2}(\mathbfsl{a}). We split the proof into two cases.

Case 1: 𝒂𝑨𝒏\mathbfsl{a}\in A^{n}: In this case Lemma 2.1 implies that there are g1,g2𝒞g_{1},g_{2}\in\mathcal{C} such that g1|BAn=g2|BAng_{1}|_{B\cap A^{n}}=g_{2}|_{B\cap A^{n}} and g1(𝒂)𝒈2(𝒂)g_{1}(\mathbfsl{a})\neq g_{2}(\mathbfsl{a}). Then for i{1,2}i\in\{1,2\}, we set

fi(𝒙)={𝒈𝒊(𝒙) if 𝒙𝑨𝒏,𝒖 otherwise.f_{i}(\mathbfsl{x})=\begin{cases}g_{i}(\mathbfsl{x})&\text{ if }\mathbfsl{x}\in A^{n},\\ u&\text{ otherwise}.\end{cases}

By the definition of Φ(𝒞)\Phi(\mathcal{C}), both f1f_{1} and f2f_{2} belong to Φ(𝒞)\Phi(\mathcal{C}), f1|B=f2|Bf_{1}|_{B}=f_{2}|_{B} and, since 𝒂𝑨𝒏\mathbfsl{a}\in A^{n}, f1(𝒂)𝒇2(𝒂)f_{1}(\mathbfsl{a})\neq f_{2}(\mathbfsl{a}).

Case 2: 𝒂𝑨𝒏\mathbfsl{a}\notin A^{n}: Let v1,v2Av_{1},v_{2}\in A with v1v2v_{1}\neq v_{2}. For i{1,2}i\in\{1,2\}, we define

fi(𝒙)={𝒖 if 𝒙(𝑨)𝒏(𝑨𝒏{𝒂}),𝒙1 if 𝒙𝑨𝒏,𝒗𝒊 if 𝒙{𝒂}.f_{i}(\mathbfsl{x})=\begin{cases}u&\text{ if }\mathbfsl{x}\in(A^{\oplus})^{n}\setminus(A^{n}\cup\{\mathbfsl{a}\}),\\ x_{1}&\text{ if }\mathbfsl{x}\in A^{n},\\ v_{i}&\text{ if }\mathbfsl{x}\in\{\mathbfsl{a}\}.\end{cases}

By construction, we have that f1|An,f2|An𝒞f_{1}|_{A^{n}},f_{2}|_{A^{n}}\in\mathcal{C}. Hence f1,f2Φ(𝒞)f_{1},f_{2}\in\Phi(\mathcal{C}). Moreover, f1|B=f2|Bf_{1}|_{B}=f_{2}|_{B} and f1(𝒂)𝒇2(𝒂)f_{1}(\mathbfsl{a})\neq f_{2}(\mathbfsl{a}). Thus, we can conclude that BAlgΦ(𝒞)B\in\operatorname{Alg}\Phi(\mathcal{C}). Equation (5.4) implies that for all 𝒞,𝒟A\mathcal{C},\mathcal{D}\in\mathcal{L}_{A}, if 𝒟≁alg𝒞\mathcal{D}\not\sim_{\text{alg}}\mathcal{C} then Φ(𝒞)≁algΦ(𝒟)\Phi(\mathcal{C})\not\sim_{\text{alg}}\Phi(\mathcal{D}).

Since by Corollary 5.5 there are 202^{\aleph_{0}} algebraically inequivalent clones on the set {0,1,2}\{0,1,2\} and since there are at most 202^{\aleph_{0}} clones on a finite set, we deduce that on each finite set AA with |A|3\lvert A\rvert\geq 3 there are exactly 202^{\aleph_{0}} algebraically inequivalent clones. ∎

Acknowledgements

The authors thank Tamás Waldhauser for discussions on the questions addressed in the present note and the referee, who has suggested the content of Lemmas 3.1 and 5.3 to replace our originally more involved argument for proving Theorems 1.3 and 1.1.

References

  • [1] István Ágoston, János Demetrovics, and László Hannák, On the number of clones containing all constants (a problem of R. McKenzie), Lectures in universal algebra (Szeged, 1983), Colloq. Math. Soc. János Bolyai, vol. 43, North-Holland, Amsterdam, 1986, pp. 21–25.
  • [2] Erhard Aichinger, Constantive Maľcev clones on finite sets are finitely related, Proc. Amer. Math. Soc. 138 (2010), no. 10, 3501–3507.
  • [3] by same author, On the direct decomposition of nilpotent expanded groups, Comm. Algebra 42 (2014), no. 6, 2651–2662.
  • [4] Erhard Aichinger and Peter Mayr, Polynomial clones on groups of order pqpq, Acta Math. Hungar. 114 (2007), no. 3, 267–285.
  • [5] Erhard Aichinger and Nebojša Mudrinski, On various concepts of nilpotence for expansions of groups, Publ. Math. Debrecen 83 (2013), 583–604.
  • [6] Erhard Aichinger and Bernardo Rossi, A clonoid based approach to some finiteness results in universal algebraic geometry, Algebra Universalis 81 (2020), no. 1, Paper No. 8, 7.
  • [7] Andrei Bulatov, On the number of finite Maľtsev algebras, Contributions to General Algebra 13 (Velké Karlovice, 1999/Dresden, 2000), Heyn, Klagenfurt, 2001, pp. 41–54.
  • [8] Stanley Burris and Hanamantagouda P. Sankappanavar, A course in universal algebra, Graduate Texts in Mathematics, vol. 78, Springer-Verlag, New York, 1981, available on-line at http://www.math.uwaterloo.ca/~snburris/htdocs/UALG/univ-algebra.pdf.
  • [9] Miguel Couceiro and Erkko Lehtonen, Galois theory for sets of operations closed under permutation, cylindrification, and composition, Algebra Universalis 67 (2012), no. 3, 273–297.
  • [10] Èvelina Yu. Daniyarova, Alexei G. Myasnikov, and Vladimir N. Remeslennikov, Algebraic geometry over algebraic structures. II. Foundations, Fundam. Prikl. Mat. 17 (2011/12), no. 1, 65–106, translation in J. Math. Sci. (N.Y.) 185 (2012), no. 3, 389–416.
  • [11] Stefano Fioravanti, Expansions of abelian square-free groups, Internat. J. Algebra Comput. 31 (2021), no. 4, 623–638.
  • [12] Paweł M. Idziak, Clones containing Maľtsev operations, Internat. J. Algebra Comput. 9 (1999), no. 2, 213–226.
  • [13] Jurij I. Janov and Aľbert A. Mučnik, Existence of kk-valued closed classes without a finite basis, Dokl. Akad. Nauk SSSR 127 (1959), no. 1, 44–46, (Russian).
  • [14] Anatolij I. Maľcev, Iterative Post algebras, Nauka, Novosibirsk, 1976.
  • [15] Peter Mayr, Polynomial clones on squarefree groups, Internat. J. Algebra Comput. 18 (2008), no. 4, 759–777.
  • [16] Ralph N. McKenzie, George F. McNulty, and Walter F. Taylor, Algebras, lattices, varieties. Vol. I, The Wadsworth & Brooks/Cole Mathematics Series, vol. I, Wadsworth & Brooks/Cole Advanced Books & Software, Monterey, CA, 1987.
  • [17] Aleksandr G. Pinus, On algebraically equivalent clones, Algebra Logika 55 (2016), no. 6, 760–768, translation in Algebra Logic 55 (2016), no. 6, 501–506.
  • [18] Boris I. Plotkin, Some concepts of algebraic geometry in universal algebra, Algebra i Analiz 9 (1997), no. 4, 224–248, translation in St. Petersburg Math. J. 9 (1998), no. 4, 859–879.
  • [19] Reinhard Pöschel and Lev A. Kalužnin, Funktionen- und Relationenalgebren, Mathematische Monographien, vol. 15, VEB Deutscher Verlag der Wissenschaften, Berlin, 1979.
  • [20] Emil L. Post, The two-valued iterative systems of mathematical logic, Annals of Mathematics Studies, vol. 5, Princeton University Press, Princeton, N. J., 1941.
  • [21] Endre Tóth and Tamás Waldhauser, On the shape of solution sets of systems of (functional) equations, Aequationes Math. 91 (2017), no. 5, 837–857.