Profinite semigroups and applications in Computer Science Jorge Almeida Departamento Matem´atica, CMUP Faculdade de Ciˆencias, Universidade do Porto Porto, Portugal http://www.fc.up.pt/cmup/jalmeida/ October, 2011 ´Ustav matematiky a statistiky Pˇr´ırodovˇedeck´a fakulta Masarykova univerzita 1 / 332 Abstract Finite semigroups appear naturally in Computer Science, namely as syntactic semigroups of regular languages, transition semigroups of finite automata, or as finite recognizing devices on their own. Eilenberg’s correspondence theorem gives a general framework for the classification of regular languages through algebraic properties of their syntactic semigroups. Here is the resulting typical problem on the algebraic side: a recursively enumerable set R of finite semigroups is given and one wishes to decide whether a given finite semigroup is a homomorphic image of a subsemigroup of a finite product of members of R. Since such a problem is often undecidable, special techniques have been devised to handle special cases. Relatively free profinite semigroups turn out to be quite useful in this context. They play the role of free algebras in Universal Algebra, capturing in their algebraic-topological/metric structure combinatorial properties of the corresponding classes of languages. The aim of this short course is to introduce reltively free profinite semigroups and to explore two topics in which there have been significant recent developments, namely the separation of a given word from a given regular language by a regular language of a special type (for instance, a group language), and connections with symbolic dynamics. Tentative syllabus and preliminary references: Part 1 Relatively free profinite semigroups. (1 lecture) Reference: [1] J. Almeida, Profinite semigroups and applications, in ”Structural Theory of Automata, Semigroups, and Universal Algebra”, V. B. Kudryavtsev and I. G. Rosenberg (eds.), Proceedings of the NATO Advanced Study Institute on Structural Theory of Automata, Semigroups and Universal Algebra (Montr´eal, Qu´ebec, Canada, 7-18 July 2003), Springer, New York, 2005, pp. 1-45. Part 2 Separating words and regular languages. (2 lectures) Reference: [2] S. Margolis, M. Sapir, and P. Weil, Closed subgroups in pro-V topologies and the extension problem for inverse automata, Int. J. Algebra and Comput. 11 (2001) 405-455. Part 3 Relatively free profinite semigroups and Symbolic Dynamics. (2 lectures) Reference: [1] (see above). 2 / 332 Part I Relatively free profinite semigroups 3 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 4 / 332 A regular language is a subset of the free monoid A∗ on an alphabet A admitting a regular expression, i.e., a formal expression describing it in terms of the empty set ∅ and the letters a ∈ A using the following operations: (K, L) → K ∪ L (union) (K, L) → KL (concatenation) L → L∗ (Kleene star) The syntactic congruence of the language L ⊆ A∗ is the binary relation σL on A∗ defined by: u σL v if ∀ x, y ∈ A∗ (xuy ∈ L ⇔ xvy ∈ L). The syntactic monoid M(L) of the language L ⊆ A∗ is the quotient monoid A∗/σL. 5 / 332 Theorem 1.1 The following conditions are equivalent for a language L over a finite alphabet A: (1) L is regular; (2) L is recognized by some finite automaton; (3) L is recognized by some finite complete deterministic automaton; (4) the syntactic monoid A∗/σL on A∗ is finite; (5) L is recognized by some homomorphism ϕ : A∗ → M into a finite monoid, in the sense that L = ϕ−1ϕL. Corollary 1.2 The set Reg(A∗) of all regular languages over the alphabet A is a Boolean subalgebra of the Boolean algebra of all subsets of A∗. 6 / 332 Example: (restricted) Dyck languages Regular expression: L1 = (ab)∗ Minimal (incomplete) automaton: 0 1 a bTransition monoid (M(L1)): 0 1 a 1 b - 0 ab 0 ba - 1 0 - a b ab ba 0 a 0 ab 0 a 0 b ba 0 b 0 0 ab a 0 ab 0 0 ba 0 b 0 ba 0 0 0 0 0 0 0 Presentation: a, b; aba = a, bab = b, a2 = b2 = 0 . One may then compute Green’s relations, which are summarized in the following eggbox picture: • same row: elements generate the same right ideal (R) • same column: elements generate the same left ideal (L) • elements above are factors of elements below (≥J ) • ∗ e marks an idempotent (e2 = e) • the“eggboxes”are the J -classes (J = ≥J ∩ ≤J ) • D = R ◦ L = L ◦ R • in a finite monoid, D = J *0 a *ab *ba b *1 7 / 332 Regular expression: L2 = (a(ab)∗ b)∗ Minimal (incomplete) automaton: 0 1 2 a b a b Presentation of syntactic monoid M(L2): a, b; aba = a, bab = b, a2 b2 a2 = a2 , b2 a2 b2 = b2 , ab2 a = ba2 b, a3 = b3 = 0 Eggbox picture: *0 aa aab *aabb baa *abba abb *bbaa bba bb a *ab *ba b *1 8 / 332 Dyck language: L∞ = n≥0 Ln, where L0 = {1}, Ln+1 = (aLnb)∗. Recognition by infinite automaton: 0 1 2 · · · n n + 1 · · · a b a b a b a b a b a b Syntactic monoid: M(L∞) = a, b; ab = 1 . Eggbox picture: ∗ 1 a a2 · · · an an+1 · · · b ∗ ba ba2 · · · ban ban+1 · · · b2 b2 a ∗ b2 a2 · · · b2 an b2 an+1 · · · ... ... ... ... ... ... bn bn a bn a2 · · · ∗ bn an bn an+1 · · · bn+1 bn+1 a bn+1 a2 · · · bn+1 an ∗ bn+1 an+1 · · · ... ... ... ... ... ... 9 / 332 Exercise 1.3 Consider the transition semigroup S of the following infinite automaton: 2 4 c ÐÐ 6 c ÐÐ 8 c ÐÐ · · · c ~~ 1 a yy b GG 3 a yy b GG 5 a yy b GG 7 a yy b GG · · · 1. Note that, in S, aca is a factor of a but a is not regular. 2. Verify that S admits the following presentation: a, b, c; baca = a, bacb = b2 ac = b, cbac = c, a2 = ab = bc = c2 = 0 . 3. Show that S has two J -classes, one of which is reduced to zero. 4. Show that the non-trivial J -class of S consists of two infinite D-classes, one of which is regular and a bicyclic monoid, while the other is not regular and has only one L-class. All H-classes of S are trivial. 10 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 11 / 332 A variety of languages is a correspondence V associating with each finitely generated free monoid A∗ a set V(A∗) of languages over the finite alphabet A such that the following conditions hold: 1. V(A∗ ) is a Boolean subalgebra of Reg(A∗ ); 2. if L ∈ V(A∗ ) and a ∈ A, then the following languages also belong to V(A∗ ): a−1 L = {w ∈ A∗ : aw ∈ L} La−1 = {w ∈ A∗ : wa ∈ L}; 3. if ϕ : A∗ → B∗ is a homomorphism and L ∈ V(B∗ ), then ϕ−1 (L) ∈ V(A∗ ). A pseudovariety of monoids is a nonempty class V of finite monoids which is closed under taking homomorphic images, submonoids, and finite direct products. 12 / 332 Theorem 2.1 (Eilenberg [Eil76]) The complete lattices of varieties of languages and of pseudovarieties of monoids are isomorphic. More precisely, the following correspondences are mutually inverse isomorphisms between the two lattices: to a variety V of languages, associate the pseudovariety V generated by all syntactic monoids M(L) with L ∈ V(A∗) for some finite alphabet A; to a pseudovariety V, associate the variety of languages V such that, for each finite alphabet A, V(A∗) consists of the languages L ⊆ A∗ such that M(L) ∈ V. 13 / 332 Thus, problems about varieties of languages admit a translation into problems about pseudovarieties of monoids. For instance, to determine if a language L ⊆ A∗ belongs to smallest variety of languages containing two given varieties of languages V and W is equivalent to determine if M(L) belongs to the pseudovariety join V ∨ W. Typically, we are given a recursively enumerable set R of finite monoids and we want to determine an algorithm to decide whether a given finite monoid M belongs to the pseudovariety V(R) generated by R. 14 / 332 Mutatis mutandis, we have languages L ⊆ A+ without the empty word 1; syntactic congruence σL of L over A+: u σL v if ∀ x, y ∈ A∗ (xuy ∈ L ⇔ xvy ∈ L). syntactic semigroup A+/σL; varieties of languages without the empty word; pseudovarieties of semigroups; Eilenberg’s correspondence in this setting. 15 / 332 Examples of pseudovarieties: S: all finite semigroups I: all singleton (trivial) semigroups G: all finite groups Gp: all finite p-groups A: all finite aperiodic semigroups Com: all finite commutative semigroups J: all finite J -trivial semigroups R: all finite R-trivial semigroups L: all finite L-trivial semigroups Sl: all finite semilattices RZ: all finite right-zero semigroups B: all finite bands N: all finite nilpotent semigroups K: all finite semigroups in which idempotents are left zeros D: all finite semigroups in which idempotents are right zeros 16 / 332 Important examples of instances of Eilenberg’s correspondence A language L ⊆ A+ is said to be star free if it admits an expression in terms of the languages {a} (a ∈ A) using only the operations: ∪ , A+ \ , and concatenation. Theorem 2.2 ([Sch65]) A language over a finite alphabet is star free if and only if its syntactic semigroup is finite and aperiodic. 17 / 332 A language L ⊆ A∗ is piecewise testable if it is a Boolean combination of languages of the form A∗a1A∗a2A∗ · · · anA∗, with the ai ∈ A. Theorem 2.3 ([Sim75]) A language over a finite alphabet is piecewise testable if and only if its syntactic semigroup is finite and J -trivial. A language L ⊆ A∗ is locally testable if it is a Boolean combination of languages of the forms A∗u, A∗vA∗, and wA∗, where u, v, w ∈ A+. Theorem 2.4 ([BS73, MP71]) A language L over a finite alphabet is locally testable if and only if its syntactic semigroup S is finite and a local semilattice (i.e., eSe is a semilattice for every idempotent e ∈ S). 18 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 19 / 332 Definition 3.1 We say that a pseudovariety V is decidable if there is an algorithm which, given a finite semigroup S as input, produces as output, in finite time, YES or NO according to whether or not S ∈ V. The semigroup S may be given in various ways: extensively, meaning the complete list of its elements together with its multiplication table; as the transformation semigroup on a finite set Q generated by a finite set A of transformations of Q ; → transition semigroup of a finite automaton (Q, A, δ, I, F); by means of a presentation. Different ways of describing S may lead to different complexity results, when such an algorithm exists. 20 / 332 Of course, not all pseudovarieties are decidable. For instance, if P is a non-recursive set of primes, then the pseudovariety AbP, generated by all groups Z/pZ with p ∈ P, contains a group Z/qZ of prime order q if and only if q ∈ P. Since there are non-recursive sets of primes P, there are pseudovarieties of the form AbP which are not decidable. Question 3.2 (Very imprecise!!) Are all“natural”pseudovarieties decidable? 21 / 332 There are many ways to construct new pseudovarieties from known ones, that is by applying operators to pseudovarieties. We proceed to introduce some natural operators. Definition 3.3 Given a pseudovariety V, consider the classes of all finite semigroups S such that, respectively: LV: eSe ∈ V for every idempotent e ∈ S; EV: E(S) ∈ V, where E(S) is the subsemigroup generated by the set E(S) of all idempotents of S; DV: the regular J -classes of S (are subsemigroups which) belong to V; V: the subgroups of S belong to V; 22 / 332 Let S be a finite semigroup and let D be one of its regular D-classes. Let ∼ be the equivalence relation on the set of group elements of D generated by the identification of elements which are either R or L-equivalent. A block of D is the Rees quotient of the subsemigroup of S generated by a ∼-class modulo the ideal consisting of the elements which do not lie in D. * * * * * * * * * * * * * * * * 1 2 3 4 5 6 * * * * * * * * 1 3 6 2 4 5 The blocks of S are the blocks of its regular D-classes. Definition 3.4 For a pseudovariety V, let BV be the class of all finite semigroups whose blocks lie in V. 23 / 332 Proposition 3.5 For a pseudovariety V, the classes BV, DV, EV, LV, V are pseudovarieties. Moreover, if V is decidable then so are those pseudovarieties. Proof. We consider only the case of LV, leaving all other cases as exercises. If ϕ : S → T is an onto homomorphism, with S ∈ S, and f ∈ E(T), then ∃e ∈ ϕ−1 (f ) ∩ E(S) and ϕ|eSe : eSe → fTf is an onto homomorphism ∴ LV is closed under taking homomorphic images. If S ≤ T and e ∈ E(S), then eSe ≤ eTe ∴ LV is closed under taking subsemigroups. If S, T are semigroups, e ∈ E(S), and f ∈ E(T), then (e, f )(S × T)(e, f ) eSe × fTf ∴ LV is closed under taking finite direct products. Given a finite semigroup, one can compute its set of idempotents E(S) and, for each e ∈ E(S), the monoid eSe. Provided V is decidable, one can then effectively check whether eSe ∈ V. Hence one can effectively check whether S ∈ LV. 24 / 332 But, the most interesting operators are defined not in structural terms but rather by describing generators: the resulting pseudovariety is given as the smallest pseudovariety containing certain semigroups which are constructed from those in the argument pseudovarieties. Definition 3.6 We say that a semigroup S divides a semigroup T, or that S is a divisor of T, and we write S T, if S is a homomorphic image of a subsemigroup of T. Proposition 3.7 Let C be a class of finite semigroups. Then the smallest pseudovariety V(C) containing C consists of all divisors of products of the form S1 × · · · × Sn with S1, . . . , Sn ∈ C. In particular, if C is closed under finite direct product, then V(C) consists of all divisors of elements of C. 25 / 332 Let S and T be semigroups and let ϕ : T1 → End S be a homomorphism of monoids, with endomorphisms acting on the left. For s ∈ S and t ∈ T1, let ts = ϕ(t)(s). The semidirect product S ∗ϕ T is the set S × T under the multiplication (s1, t1) · (s2, t2) = (s1 t1 s2, t1t2). Definition 3.8 The semidirect product V ∗ W of the pseudovarieties V and W is the smallest pseudovariety containing all semidirect products S ∗ T with S ∈ V and T ∈ W. Proposition 3.9 The pseudovariety V ∗ W consists of all divisors of semidirect products of the form S ∗ T with S ∈ V and T ∈ W. Proposition 3.10 The semidirect product of pseudovarieties is associative. 26 / 332 Definition 3.11 The Mal’cev product V m W of two pseudovarieties V and W is the smallest pseudovariety containing all finite semigroups S for which there exists a homomorphism ϕ : S → T such that T ∈ W and ϕ−1(e) ∈ V for all e ∈ E(T). Given two semigroups S and T, a relational morphism S → T is a relation µ : S → T with domain S such that µ is a subsemigroup of S × T. Proposition 3.12 The pseudovariety V m W consists of all finite semigroups S such that there is a relational morphism µ : S → T such that T ∈ W and µ−1(e) ∈ V for all e ∈ E(T). 27 / 332 For a semigroup S, denote by P(S) the semigroup of subsets of S under the product operation X · Y = {xy : x ∈ X, y ∈ Y }. Note that the empty set ∅ is a zero and P (S) = P(S) \ {∅} is a subsemigroup. Definition 3.13 For a pseudovariety V, denote by PV: the pseudovariety generated by all semigroups of the form P(S), with S ∈ V; P V: the pseudovariety generated by all semigroups of the form P (S), with S ∈ V. Proposition 3.14 The pseudovariety PV consists of all divisors of semigroups of the form P(S) with S ∈ V. Similar statement for P . 28 / 332 Some examples of results on finite semigroups formulated in terms of these operators: 1. J = N m Sl 2. DA = LI m Sl, DS = LG m Sl 3. R = Sl ∗ J [Sti73] 4. G ∨ Com = ZE (the pseudovariety of all finite semigroups in which idempotents are central) [Alm95] 5. ESl = Sl ∗ G = Sl m G = Inv (the pseudovariety generated by all finite inverse semigroups) [MP87, Ash87, Pin95], ER = R ∗ G [Eil76], EDS = DS ∗ G [AE03] 6. PG = J ∗ G = J m G = EJ = BG [MP84, HR91, Ash91, HMPR91, Pin95], PJ = PV(Y ) [PS85, Alm95] where Y = Synt(a∗bc∗) 7. S = n≥0(A ∗ G)n ∗ A [KR65] 29 / 332 S = n≥0 (A ∗ G)n ∗ A The (Krohn-Rhodes) hierarchy (A ∗ G)n ∗ A n≥0 is strict. The smallest n such that a given finite semigroup S belongs to (A ∗ G)n ∗ A is called the complexity of S, denoted c(S). Let Tn denote the full transformation semigroup of an n-element set It is known that c(Tn) = n − 1 [Eil76] and so certainly c(S) ≤ |S| (since S → TS1 ). Note 3.15 To know an algorithm to compute the complexity function is equivalent to know algorithms to decide the membership problem for each pseudovariety in the Krohn-Rhodes hierarchy. 30 / 332 This brings us to the following basic question: Question 3.16 For the operators which were defined above in terms of generators, do they preserve decidability? Theorem 3.17 (Albert, Baldinger & Rhodes’1992 [ABR92]) There exists a finite set Σ of identities such that Com ∨ [[Σ]] is undecidable. Let C2,1 = a; a2 = 0 1. Theorem 3.18 (Auinger & Steinberg’2003 [AS03]) There exists a decidable pseudovariety of groups U such that the following pseudovarieties are all undecidable: Sl ∗ U (= Sl m U), V(C2,1) ∨ U, PU (= P U). 31 / 332 The pseudovariety U is defined to be U = p∈A Gp ∗ (Gf (p) ∩ Com) ∨ p∈D (Gp ∩ Com) where: A and B constitute a computable partition of the set of primes into two infinite sets; f : A → B is an injective recursive function whose range C = f (A) is recursively enumerable but not recursive; D = B \ C is not recursively enumerable. Exercise 3.19 Show that U is decidable. 32 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 33 / 332 Let V be a pseudovariety of semigroups. For two words u, v ∈ A+, and T ∈ V, let T |= u = v if, for every homomorphism ϕ : A+ → T, ϕ(u) = ϕ(v), rV(u, v) = min{|S| : S ∈ V and S |= u = v}, dV(u, v) = 2−rV(u,v) where we take min ∅ = ∞ and 2−∞ = 0. Note 4.1 The following hold for u, v, w, t ∈ A+ and a positive integer n: (1) rV(u, v) ≥ n if and only if, for every S ∈ V with |S| < n, S |= u = v; (2) dV(u, v) ≤ 2−n if and only if, for every S ∈ V with |S| < n, S |= u = v; (3) dV(u, v) = 0 if and only if, for every S ∈ V, S |= u = v; (4) min{rV(u, v), rV(v, w)} ≤ rV(u, w); (5) min{rV(u, v), rV(w, z)} ≤ rV(uw, vz). 34 / 332 Definition 4.2 A function d : X × X → R≥0 is said to be a pseudo-ultrametric on the set X if the following properties hold for all u, v, w ∈ X: 1. d(u, u) = 0; 2. d(u, v) = d(v, u); 3. d(u, w) ≤ max{d(u, v), d(v, w)}. We then also say that X is a pseudo-ultrametric space. If instead of Condition 3, the following weaker condition holds 4. d(u, w) ≤ d(u, v) + d(v, w) (triangle inequality). then d is said to be a pseudo-metric on X, and X is said to be a pseudo-metric space. If the following condition holds 5. d(u, v) = 0 if and only if u = v, then we drop the prefix“pseudo”. 35 / 332 A function f : X → Y between two pseudo-metric spaces is said to be uniformly continuous if the following condition holds: ∀ > 0 ∃ δ > 0 ∀ x1, x2 ∈ X d(x1, x2) < δ ⇒ d(f (u), f (v)) < . Proposition 4.3 1. The function dV is a pseudo-ultrametric on A+ . 2. The multiplication is contractive: dV(u1u2, v1v2) ≤ max{dV(u1, v1), dV(u2, v2)}. In particular, the multiplication on A+ is uniformly continuous. For a (pseudo-ultra)metric d, u ∈ X, and a positive real number , consider the open ball B (u) = {v ∈ X : d(u, v) < }. The point u is the center and is the radius of the ball. A metric space that can be covered by a finite number of balls of any given positive radius is said to be totally bounded. 36 / 332 Proposition 4.4 The metric space (A+, dV) is totally bounded. Proof. Let n be a positive integer such that 2−n < . Note that, up to isomorphism, there are only finitely many semigroups of cardinality at most n in V. For such a semigroup Si consider all possible homomorphisms ϕi,j : A+ → Si , let S = i,j Si and ϕ : A+ → S u → (ϕi,j (u))i,j . Then S ∈ V and dV(u, v) < 2−n if and only if ϕ(u) = ϕ(v). For each s ∈ S, choose us ∈ A+ such that ϕ(us) = s. For v ∈ A+ and s = ϕ(v), we have ϕ(v) = ϕ(us), and so v ∈ B (us). We have thus shown that A+ = s∈S B (us). 37 / 332 A sequence (un)n in a (pseudo-ultra)metric space X is said to be a Cauchy sequence if ∀ > 0 ∃ N m, n ≥ N ⇒ d(um, un) < . Note that every convergent sequence is a Cauchy sequence. The space X is complete if every Cauchy sequence in X converges in X. 38 / 332 Theorem 4.5 Let X be a pseudo-(ultra)metric space. Then there exists a complete metric space ˆX and a uniformly continuous function ι : X → ˆX with the following universal property: for every uniformly continuous function f : X → Y into a complete metric space Y , there exists a unique uniformly continuous function ˆf : ˆX → Y such that ˆf ◦ ι = f . X ι GG f 22 ˆX ˆf  Y . X ι  γ 11 ˆX ˆγ CC Z. ˆι kk In particular, if γ : X → Z is another uniformly continuous function into another complete metric space with the above universal property then the induced unique uniformly continuous mappings ˆι : ˆX → Z and ˆγ : Z → ˆX are mutually inverse. The“unique”space ˆX of Theorem 4.5 is called the Hausdorff completion of X. 39 / 332 It may be constructed in the same way that the real numbers are obtained by completion of the rational numbers. Here is a sketch: (a) consider the set C ⊆ XN of all Cauchy sequences of elements of X; (b) note that, for s = (un)n and t(vn)n in C, the sequence of real numbers d(un, vn) n is a Cauchy sequence and, therefore, it converges; its limit is denoted d(s, t); |d(un, vn) − d(um, vm)| ≤ |d(un, vn) − d(un, vm)| + |d(un, vm) − d(um, vm)| ≤ d(un, um) + d(vn, vm) (c) Step (B) defines a pseudo-(ultra)metric on C; (d) for s = (un)n and t(vn)n in C, let s ∼ t if d(s, t) = 0; this is an equivalence relation on C; the class of s is denoted s/∼; (e) let ˆX = C/∼ and put d(s/∼, t/∼) = d(s, t), which can be easily checked to be defined; (f) finally, let ι : X → ˆX map each u ∈ X to the ∼-class of the constant sequence (u)n, and check that this mapping is uniformly continuous and has the appropriate universal property. 40 / 332 Note that ι(X) is dense in ˆX. In particular, we may consider the Hausdorff completion of the pseudo-ultrametric space (A+, dV), which is denoted ΩAV. Since the multiplication of A+ is uniformly continuous with respect to dV, it induces a uniformly continuous multiplication in ΩAV: A+ × A+ µ (mult.) GG ι×ι  A+ ι  ΩAV × ΩAV ˆµ GG ΩAV 41 / 332 We endow each finite semigroup S with the discrete metric: d(s, t) = 0 if s = t 1 otherwise Since ι(A+) is dense in ΩAV, multiplication in ΩAV is associative, and thus ΩAV is naturally a semigroup. From hereon, we write d for dV. The context should leave clear which pseudovariety is involved. 42 / 332 Note that, for S ∈ V, every homomorphism ϕ : A+ → S is uniformly continuous with respect to d. d(u, v) < 2−|S| ⇒ d(ϕ(u), ϕ(v)) = 0. Thus, ϕ induces a unique uniformly continuous mapping ˆϕ : ΩAV → S such that the following diagram commutes: A+ ι GG ϕ 44 ΩAV ˆϕ  S. One can easily check that ˆϕ is a homomorphism: ˆϕ(uv) = lim ϕ(ι(unvn)) = lim ϕ(ι(un))ϕ(ι(vn)) = lim ϕ(ι(un)) · lim ϕ(ι(vn)) = ˆϕ(u) ˆϕ(v). 43 / 332 Given u, v ∈ ΩAV and S ∈ V, we write S |= u = v if, for every homomorphism ϕ : A+ → S (which is determined by ϕ|A), the equality ˆϕ(u) = ˆϕ(v) holds. We call the formal equality u = v a V-pseudoidentity. Note that, if u = lim un, v = lim vn, and S ∈ V, then S |= u = v if and only if S |= un = vn for all sufficiently large n. Given distinct elements u, v ∈ ΩAV, there exists a positive integer m such that d(u, v) ≥ 2−m . Consider sequences of words (un)n and (vn)n such that u = lim ι(un) and v = lim ι(vn). Then, for sufficiently large n, d(u, ι(un)) < 2−m and d(v, ι(vn)) < 2−m . Hence d(un, vn) = d(ι(un), ι(vn)) ≥ 2−m for all sufficiently large n. It follows that every S ∈ V with |S| < m fails the identity un = vn and, therefore, also the pseudoidentity u = v. 44 / 332 Proposition 4.6 For u, v ∈ ΩAV, we have d(u, v) = 2−r(u,v), where r(u, v) = min{|S| : S ∈ V and S |= u = v}. Proof. We have already shown that d(u, v) ≥ 2−m implies r(u, v) ≤ m. The converse, as well as how the equivalence gives the proposition are left as an exercise. 45 / 332 Recall that a metric space is compact if every sequence admits some convergent subsequence. Equivalently, every covering by open subsets contains a finite covering. Proposition 4.7 1. If X is a totally bounded pseudo-metric space, then ˆX is also totally bounded. 2. If X is a totally bounded complete metric space, then X is compact. 46 / 332 Proof. 1. Given > 0, let u1, . . . , um ∈ X be such that X = m i=1 B /2(ui ). Then ˆX = m i=1 B (ι(ui )) since every element of ˆX is at distance at most /2 of some element of ι(X). 2. For each positive integer m, let Fm be a finite subset of X such that X = x∈Fm B2−m (x) and consider an arbitrary sequence (un)n in X. For infinitely many indices n, the un belong to the same B2−1 (x1). Let k1 be the first of these indices. Similarly, among the remaining such indices, there are infinitely many n such that the un belong to the same B2−2 (x2). We let k2 be the first of them. And so on. We thus construct a subsequence (ukn )n with the property that d(ukm , ukn ) ≤ 2− min{m,n}+1 , if p = min{m, n}, then ukm , ukn ∈ B2−p (xp), which yields d(ukm , ukn ) ≤ d(ukm , xp) + d(xp, ukn ) ≤ 2−p + 2−p whence a Cauchy sequence and, therefore, convergent. 47 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 48 / 332 By a pro-V semigroup we mean a semigroup S endowed with a metric such that the following properties hold: 1. S is compact; 2. the multiplication is uniformly continuous (metric semigroup); 3. for every pair u, v of distinct elements of S, there is a uniform continuous homomorphism ϕ : S → T into a semigroup from V such that ϕ(u) = ϕ(v) (S residually in V). By a profinite semigroup we mean a pro-S semigroup. 49 / 332 Proposition 5.1 Let S be a pro-V semigroup. Then there is a sequence (Sn)n∈N of semigroups from V and an injective homomorphism ϕ : S → n∈N Sn such that, for each component projection πm : n∈N Sn → Sm, the homomorphism πm ◦ ϕ is uniformly continuous. We may define in n∈N Sn a metric structure by letting d(u, v) = n∈N 2−n dn(πn(u), πn(v)) where dn is the discrete metric on Sn. Then ϕ is uniformly continuous. In particular, the image T of ϕ is closed in n∈N Sn, being a compact subset. 50 / 332 Note that the sequence (Sn)n∈N may be chosen so that there is a finite bound on the number of generators of the Sn if and only if S is finitely generated in the sense that there is a finite subset which generates a dense subsemigroup. On the other hand, if there is no such bound, one can show that S cannot have a countable dense subset, while it is easy to see that a compact metric space always admits a countable dense subset. Proposition 5.2 Every pro-V metric semigroup is finitely generated. 51 / 332 For a finite set A, we say that the pro-V semigroup S is freely generated by A if there is a mapping γ : A → S such that γ(A) generates a dense subsemigroup of S and the following universal property is satisfied, where ϕ : A → T is an arbitrary mapping into a pro-V semigroup T, and ˆϕ is a unique continuous homomorphism: A γ GG ϕ 11 S ˆϕ  T Theorem 5.3 For a pseudovariety of semigroups V and a finite set A, the metric semigroup ΩAV is a pro-V semigroup freely generated by A via the mapping ι|A. 52 / 332 Proof. Let S be a pro-V semigroup and let (Sn)n∈N be a countable family of semigroups from V as given by Proposition 5.1, so that there is an embedding ϕ : S → n∈N Sn with each composite function πn ◦ ϕ : S → Sn uniformly continuous. Given a mapping ψ : A → S, let ψn = πn ◦ ψ. A ι|A GG ψ  ψn 22 ΩAV ˆψn  S πn GG Sn The family ( ˆψn)n∈N induces a homomorphism ˆψ : ΩAV → n∈N Sn. Its image lies in the closed subsemigroup T, whence it lifts to the required continuous homomorphism ΩAV → S. It is uniformly continuous because every continuous mapping from a compact metric space into another metric space is uniformly continuous. 53 / 332 A subset of a metric space is said to be clopen if it is both closed and open. A metric space is said to be zero-dimensional if every open set is a union of clopen subsets. Proposition 5.4 Every pro-V semigroup is zero-dimensional. Proof. Let u be an element of the pro-V semigroup S. It suffices to show that the open ball B (u) contains some clopen set which contains u. For each v ∈ S \ B (u), let ϕv : S → Tv be a uniformly continuous homomorphism into a semigroup from V such that ϕv (u) = ϕv (v). Then Kv = ϕ−1 v ϕv (v) is a clopen set which contains v but not u. In particular, the Kv form a clopen covering of the closed set S \ B (u), from which a finite covering F can be extracted. The union of the clopen sets in F is itself a clopen set K. Note that S \ K is also clopen, contains u, and is contained in B (u). 54 / 332 For a mapping ϕ : S → T, let ker ϕ = {(u, v) : ϕ(u) = ϕ(v)} be the kernel of ϕ. Theorem 5.5 An A-generated profinite semigroup S is a continuous homomorphic image of ΩAV if and only if it is a pro-V semigroup. Corollary 5.6 Let S be a pro-V semigroup and suppose that ϕ : S → T is a continuous homomorphism onto a finite semigroup. Then T ∈ V. 55 / 332 Proof of Theorem 5.5. (⇐) Apply Theorem 5.3. (⇒) Let ϕ : ΩAV → S be an onto continuous homomorphism. We need to show that S is residually in V. Given distinct points s1, s2 ∈ S, since S is residually in S, there is an onto uniformly continuous homomorphism ψ : S → T such that T ∈ S and ψ(s1) = ψ(s2). Note that T is a finite continuous homomorphic image of ΩAV. If we can show that S ∈ V, we will be done. In other words, it suffices to consider the case where S is finite. Since ϕ is continuous and ΩAV is compact, ϕ is uniformly continuous. Hence, there is a positive integer n such that, for all u, v ∈ ΩAV, d(u, v) < 2−n ⇒ ϕ(u) = ϕ(v). In view of Proposition 4.6, it follows that the intersection ρ of the kernels of the uniformly continuous homomorphisms ΩAV → V with V ∈ V and |V | ≤ n is contained in ker ϕ. Hence, ϕ factors through the natural homomorphism ΩAV → ΩAV/ρ. Since ΩAV/ρ belongs to V, so does S. 56 / 332 Lemma 5.7 ([Num57, Hun88]) Let K be a clopen subset of a compact zero-dimensional metric semigroup S. Then there is a continuous homomorphism ϕ : S → T into a finite semigroup T such that K = ϕ−1 ϕ(K). Proof. We may define on S a syntactic congruence of K by u σK v if ∀ x, y ∈ S1 (xuy ∈ K ⇔ xvy ∈ K). It suffices to show that the classes of this congruence are open: then there are only finitely many of them, so that S/σK is a finite semigroup, and the natural mapping S → S/σK is a continuous homomorphism. We show that, if lim un = u, then all but finitely many terms in the sequence are σK -equivalent to u. Arguing by contradiction, otherwise, there is a subsequence consisting of terms which fail this property. We may as well assume that so does the original sequence. For each n there are xn, yn ∈ S1 such that one, but not both, of the products xnunyn and xnuyn lies in K. Again, by taking subsequences we may assume that lim xn = x, lim yn = y (in S1 ), and xnuyn /∈ K. Then xuy = lim xnunyn = lim xnuyn must belong to both K and its complement. 57 / 332 A useful application of Lemma 5.7 is the following result, which completes that of Proposition 5.4. Theorem 5.8 A compact metric semigroup is profinite if and only if it is zero-dimensional. Proof. (⇒) This follows from Proposition 5.4. (⇐) Let S be a compact metric semigroup which is zero-dimensional. We need to show that it is residually in S, that is that, for every pair s, t of distinct points of S, there is a continuous homomorphism ϕ : S → T in to a finite semigroup T such that ϕ(s) = ϕ(t). Since S is a zero-dimensional metric space, there is some clopen subset K such that s ∈ K and t /∈ K. By Lemma 5.7, there is a continuous homomorphism ϕ : S → T into a finite semigroup T such that K = ϕ−1 ϕ(K). In particular, we have ϕ(s) = ϕ(t), as required. 58 / 332 A language L ⊆ A+ is V-recognizable if its syntactic semigroup belongs to V. Theorem 5.9 A language L ⊆ A+ is V-recognizable if and only if the closure K = ι(L) is open in ΩAV and ι−1 (K) = L. The latter condition is superfluous if ι is injective and ι(A+ ) is a discrete subset of ΩAV. Proof. (⇒) Use the universal property of ΩAV (Theorem 5.3). (⇐) By Lemma 5.7, there is a continuous homomorphism ϕ : ΩAV → S such that S ∈ V and K = ϕ−1 ϕ(K). Then ψ = ϕ ◦ ι is a homomorphism A+ → S such that ψ−1 ϕ(K) = ι−1 (K) = L and so L is V-recognizable. Theorem 5.9 implies that, as a topological space, ΩAV is the Stone dual of the Boolean algebra of V-recognizable languages of A+ . 59 / 332 Theorem 5.10 A set S of V-recognizable languages over a finite alphabet A generates the Boolean algebra of all such languages if and only if the clopen sets of the form ι(L) (L ∈ S) suffice to separate points of ΩAV. Proof. (⇒) Let u, v ∈ ΩAV be distinct points. Then = d(u, v) is positive. Since ΩAV is zero-dimensional (Proposition 5.4), there is a clopen subset K containing u and contained in B (u), whence not containing v. By Theorem 5.9, L = ι−1 (K) is V-recognizable. From the hypothesis, it follows that L is a Boolean combination f (L1, . . . , Ln) of languages Li from S. By Theorem 5.9 again, each set ι(Li ) is clopen. Since ι(X1 ∪ X2) = ι(X1) ∪ ι(X2) and ΩAV \ ι(X) = ι(A+ \ X) for V-recognizable languages X, X1, X2 ⊆ A+ , we have K = ι(L) = f (ι(L1), . . . , ι(Ln)). Hence at least one of the sets ι(Li ) must contain exactly one of the points u and v. (⇐) By Theorem 5.9, it suffices to show that the clopen sets of the form ι(L), with L ⊆ A+ V-recognizable, generate the Boolean algebra of all clopen subsets of ΩAV. This is a nice exercise on compactness. 60 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 61 / 332 Recall that a V-pseudoidentity is a formal equality u = v with u, v ∈ ΩAV for some finite set A. Recall also that, for S ∈ V, we write S |= u = v if ϕ(u) = ϕ(v) for every continuous homomorphism ϕ : ΩAV → S. In this case, we also say that u = v holds in S. For a set Σ of V-pseudoidentities, let [[Σ]] denote the class of all S ∈ V such that S |= u = v for every pseudoidentity u = v from Σ. For a subpseudovariety W of V, let pW : ΩAV → ΩAW be the natural continuous homomorphism: A ιV GG ιW 33 ΩAV pW:=ˆιW  ΩAW 62 / 332 Lemma 6.1 A pseudoidentity u = v, with u, v ∈ ΩAV, holds in every member of a subpseudovariety W of V if and only if pW(u) = pW(v). Theorem 6.2 ([Rei82]) A subclass W of V is a subpseudovariety if and only if it is of the form [[Σ]] for some set Σ of V-pseudoidentities. Usually, one takes V = S. 63 / 332 Proof of Theorem 6.2. (⇐) This amounts to verifying that the property S |= u = v is preserved under taking homomorphic images, subsemigroups and finite direct products, which follows easily from the definitions. (⇒) Fix a countably infinite set X and let Σ be the set of all pseudoidentities u = v such that u, v ∈ ΩAV for some finite subset A of X and S |= u = v for all S ∈ W. Then U = [[Σ]] is a subpseudovariety of V by the first part of the proof, and it clearly contains W. We claim that U = W. Let S ∈ U and choose an onto continuous homomorphism ϕ : ΩAU → S for some finite subset A of X (cf. Theorem 5.3). Consider the natural continuous homomorphisms pU and pW. By Lemma 6.1 and the choice of Σ, we have ker pW ⊆ ker pU and so there is a factorization pU = ψ ◦ pW for some onto continuous homomorphism ψ : ΩAW → ΩAU. Hence ϕ ◦ ψ : ΩAW → S is an onto continuous homomorphism. Corollary 5.6 then implies that S ∈ W since ΩAW is a pro-W semigroup by Theorem 5.3. ΩAV pU GG pW  ΩAU ϕ  ΩAW ψ `` S 64 / 332 To write pseudoidentities that are not identities, one needs to construct some elements of ΩAS \ A+. Lemma 6.3 Let S be a profinite semigroup, let s be an element of S, and let k ∈ Z. Then the sequence of powers (sn!+k)n≥|k| converges. For k = 0 the limit is an idempotent. Proof. Using Proposition 5.1, it suffices to consider the case where S is finite, which is left as an exercise. The limit lim sn!+k is denoted sω+k. Note that sω+ksω+ = sω+k+ . In particular, sω := sω+0 is an idempotent and sω−k and sω+k are mutual inverses in the maximal subgroup containing the idempotent sω. 65 / 332 Examples I S = [[x = x]] I = [[x = y]] G = [[xω = 1]] Gp =? A = [[xω+1 = xω ]] Com = [[xy = yx]] J = [[(xy)ω = (yx)ω , xω+1 = xω ]] R = [[(xy)ω x = (xy)ω ]] L = [[y(xy)ω = (xy)ω ]] Sl = [[xy = yx, x2 = x]] RZ = [[xy = y]] B = [[x2 = x]] N = [[xω = 0]] K = [[xω y = xω ]] D = [[yxω = xω ]] 66 / 332 Examples II Since there are uncountably many pseudovarieties of the form AbP, where P is a set of primes, and one can show that all of them admit a description of the form [[xy = yx, u = 1]] [Alm95, Corollary 3.7.8], for some u ∈ Ω{x}S, we conclude that Ω{x}S is uncountable. Let P be an infinite set of primes and let p1, p2, . . . be an enumeration of its elements, without repetitions. Let uP be an accumulation point in Ω{x}S of the sequence (xp1···pn )n. AbP = [[xy = yx, uP = 1]]. Does the sequence (xp1···pn )n converge? 67 / 332 Examples III To describe the pseudovariety Gp of all finite p-groups, we use the following result, whose proof is similar to that of Lemma 6.3. Lemma 6.4 Let S be a profinite semigroup and s ∈ S. Then the sequence (spn! )n converges. We let spω = lim spn! . Gp = [[xpω = 1]]. 68 / 332 Examples IV Exercise 6.5 (For those that know some group theory) Find, for each of the following pseudovarieties of groups, a single pseudoidentity defining them: (1) the pseudovariety Gp of all finite groups which have no elements of order p (p being a fixed prime number); (2) the pseudovariety Gnil of all finite nilpotent groups; (3) the pseudovariety Gsol of all finite solvable groups. 69 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 70 / 332 N: finite nilpotent semigroups Recall that N = [[xω = 0]] = n≥1[[x1 · · · xn = 0]]. The proof depends on the following key result. Lemma 7.1 Let S be a finite semigroup with n elements. Then, for every choice of elements s1, . . . , sn ∈ S, there exist indices i, j such that 0 ≤ i < j ≤ n and the following equality holds for all k ≥ 1: s1 · · · sn = s1 · · · si (si+1 · · · sj )k sj+1 · · · sn. Proof. Consider the n products pr = s1 · · · sr (r = 1, . . . , n). If they are all distinct, then at least one of them, say pr , is idempotent and we may take i = 0, j = r. Otherwise, there are indices i, j such that 1 ≤ i < j ≤ n and pi = pj , in which case pi = pj = pi si+1 · · · sj = pi (si+1 · · · sj )k . 71 / 332 Let ϕ : A+ → S be a homomorphism into a semigroup S ∈ N, say satisfying x1 · · · xn = 0. Then, all words of length at least n belong to ϕ−1 (0) and for s ∈ S \ {0}, the words in the language L = ϕ−1 (s) have length less than n, and so L is a finite set. Thus, every N-recognizable language is either finite or cofinite. To show that these are precisely the N-recognizable languages, it suffices to show that every singleton language {w} ⊆ A+ is N-recognizable. Let n = |w| be the length of the word w. Consider the semigroup S consisting of the words of A+ of length at most n together with a zero element 0. The product of two words is the word resulting from their concatenation if that word has length at most n and is 0 otherwise.1 Then S satisfies the identity x1 · · · xn = 0, for the natural homomorphism ϕ : A+ → S, that sends each letter to itself, we have ϕ−1 (w) = {w}. 1 This amounts to“killing”the ideal of the semigroup A+ consisting of the words of length greater than n, identifying all the elements in the ideal to a zero. In semigroup theory, such a construction is called a Rees quotient. 72 / 332 Proposition 7.2 A language over a finite alphabet A is N-recognizable if and only if it is finite or its complement in A+ is finite. In view of Theorem 5.9, we deduce the following result: Proposition 7.3 Let V be a pseudovariety of semigroups containing N. Then the completion homomorphism ι : A+ → ΩAV is injective and A+ is a discrete subspace of ΩAV. In particular, a language L ⊆ A+ is V-recognizable if and only if its closure L in ΩAV is a clopen subset. 73 / 332 Proof. The injectivity of ι amounts to V satisfying no identity u = v with u, v ∈ A+ distinct words. Indeed, Synt({u}) is nilpotent, whence it belongs to V. Since 1u1 ∈ {u} while 1v1 /∈ {u}, we deduce that u and v are not σ{u}-equivalent and so Synt({u}) |= u = v. We may therefore identify each w ∈ A+ with ι(w) ∈ ΩAV. For w ∈ A+, we have {w} = {w}, because ΩAV is a metric space. Since {w} is V-recognizable, its closure {w} is an open subset of ΩAV by Theorem 5.9. Hence A+ is a discrete subset of ΩAV. 74 / 332 Proposition 7.4 The semigroup ΩAN is obtained by adding to A+ a zero element. The open sets containing zero consist of zero together with a cofinite subset of A+.2 Proof. It suffices to observe that a non-eventually constant sequence (wn)n of words of A+ is a Cauchy sequence with respect to the metric dN if and only if lim |wn| = ∞. In the affirmative case, for every homomorphism ϕ : A+ → S into S ∈ N, we have lim ϕ(wn) = 0. Thus, all non-eventually constant Cauchy sequences converge to the same point of ΩAN, which is a zero. The open subsets of ΩAN containing 0 have complement which is a closed, whence compact, subset of A+. Since A+ is a discrete subset of ΩAN, that complement must be finite. The converse is clear. 2 This is known as the Alexandroff or one-point compactification, which in general is obtained by adding one point and declaring the open sets containing it to consist also of the complement of a compact subset of the original space. 75 / 332 K: finite semigroups satisfying es = e Recall that K = [[xω y = xω ]]. Note that K = n≥1 Kn where Kn = [[x1 · · · xny = x1 · · · xn]]. Let AN denote the set of all right infinite words over A, i.e., sequences of letters. Endow the set S = A+ ∪ AN with the operation u · v = uv if u ∈ A+ u otherwise and the function d : S × S → R≥0 defined by d(u, v) = 2−r(u,v) , where r(u, v) is the length of the longest common prefix of u and v. 76 / 332 Proposition 7.5 The set S is a pro-K semigroup for the above operation and distance function d. The unique continuous homomorphism ΩAK → S that sends each letter a ∈ A to itself is an isomorphism. Proof. It is easy to check that the multiplication defined on S is associative and that d is an totally bounded complete ultrametric. Consider the set Sn consisting of all words of A+ of length at most n, endowed with the operation u · v = uv if |uv| ≤ n in(u) otherwise where in(w) denotes the longest prefix of length at most n of the word w. This operation is associative and Sn ∈ Kn. Moreover, every n-generated semigroup from Kn is a homomorphic image of Sn. Hence Sn ΩAKn. Note also that the mapping ϕn : S → Sn which sends each w ∈ S to in(w) is a continuous homomorphism. Hence, given two distinct points u and v from S, for n = r(u, v) + 1, the mapping ϕn is a continuous homomorphism into a semigroup from K which distinguishes u from v. Thus, S is a pro-K semigroup. 77 / 332 (. . . ) Consider next the unique continuous homomorphism ψ : ΩAK → S which maps each letter a ∈ A to itself. Since K = n≥1 Kn, given distinct u, v ∈ ΩAK, there exists a continuous homomorphism θ : ΩAK → Sn such that θ(u) = θ(v). ΩAK ψ GG θ  S ϕn  Sn ΩAKn µ oo The fact that the above diagram can always be completed by a homomorphism µ shows that ψ(u) = ψ(v). Hence ψ is injective. 78 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 79 / 332 Implicit operations Let n be a positive integer. An n-ary implicit operation on pro-V semigroups is a correspondence π associating to each pro-V semigroup S an n-ary operation πS : Sn → S such that , for every continuous homomorphism ϕ : S → T between pro-V semigroups, the following diagram commutes: Sn πS GG ϕn  S ϕ  Tn πT GG T, i.e., ϕ πS (s1, . . . , sn) = πT ϕ(s1), . . . , ϕ(sn) for all s1, . . . , sn ∈ S. Examples: the binary multiplication (s1, s2) → s1s2 and the component projections (s1, . . . , sn) → si are implicit operations. Composing implicit operations we also obtain implicit operations. 80 / 332 If A and B are finite sets with the same cardinality n, then ΩAV ΩBV. We denote by ΩnV any of them. Usually, we identify ΩnV with ΩXn V, where Xn = {x1, . . . , xn} has cardinality n. To each w ∈ ΩnV, we may associate an n-ary implicit operation πw on pro-V semigroups as follows: for a pro-V semigroup S, given s1, . . . , sn ∈ S, let f : Xn → S be the function defined by f (xi ) = si (i = 1, . . . , n); let (πw )S (s1, . . . , sn) = ˆf (w) where ˆf is a continuous homomorphism completing the following diagram: Xn ι GG f 33 ΩnV ˆf  S. 81 / 332 Proposition 8.1 1. For each w ∈ ΩnV, πw is indeed an n-ary implicit operation on pro-V semigroups. 2. The correspondence w ∈ ΩnV → πw is injective and in fact πw is completely characterized by the operations (πw )S with S ∈ V. 82 / 332 Proof. 1. Let ϕ : S → T be a continuous homomorphism between two pro-V semigroups and let s1, . . . , sn be elements of S. Let f : Xn → S be defined by f (xi ) = si (i = 1, . . . , n). Then we have the following commutative diagram: Xn f  ϕ◦f ι  ΩnV ˆf }} ϕ◦f 33 S ϕ GG T which shows that ϕ (πw )S (f (s1), . . . , f (sn)) = ϕ ˆf (w) = ϕ ◦ f (w) = (πw )T ϕ(f (s1)), . . . , ϕ(f (sn)) . 83 / 332 (. . . ) 2. Let u, v ∈ ΩnV be two distinct elements. Then there exists a continuous homomorphism ϕ : ΩnV → S into a semigroup S ∈ V such that ϕ(u) = ϕ(v). Let si = ϕ(xi ) (i = 1, . . . , n). For w ∈ ΩnV, by definition of πw we have (πw )S (s1, . . . , sn) = ϕ(w). Since ϕ(u) = ϕ(v), we deduce that (πu)S (s1, . . . , sn) = (πv )S (s1, . . . , sn) and so, certainly πu = πv . We identify w with πw . Note that S ∈ V satisfies the V-pseudoidentity u = v if and only if uS = vS . We say that a pro-V semigroup S satisfies the V-pseudoidentity u = v if uS = vS . 84 / 332 Outline Language recognition devices Eilenberg’s correspondence Decidable pseudovarieties Metrics associated with pseudovarieties Pro-V semigroups Reiterman’s Theorem Examples of relatively free profinite semigroups Pseudowords as operations Implicit signatures 85 / 332 By an implicit signature we mean a set σ of implicit operations (on S) which includes the binary operation of multiplication. Example: κ = { · , ω−1}. Given an implicit signature σ, each profinite semigroup S becomes a natural σ-algebra in which each operation w ∈ σ is interpreted as wS . In particular, each ΩAV becomes a σ-algebra. The σ-subalgebra generated by ι(A) is denoted Ωσ AV. For the minimum implicit signature σ, consisting only of multiplication, we denote Ωσ AV simply by ΩAV.3 A formal term constructed from the letters a ∈ A using the operations from the implicit signature σ is called a σ-term over A. Such a σ-term w determines an element wV of Ωσ AV by evaluating the operations within Ωσ AV. 3 The bar in the notation ΩAV comes from the fact ΩAV = ι(A+ ) is dense in ΩAV. This notation (without reference to V) was introduced by Reiterman [Rei82]. 86 / 332 The following result is an immediate consequence of Theorem 5.3. Proposition 9.1 The σ-algebra Ωσ AV is a V-free σ-algebra freely generated by A in the sense of the following universal property: for every mapping ϕ : A → S into a semigroup S ∈ V, there is a unique homomorphism ˆϕ of σ-algebras such that the following diagram commutes: A ι GG ϕ 33 Ωσ AV ˆϕ  S. 87 / 332 Examples: Ωκ AN = ΩAN; for |A| ≥ 2, since ΩAK is uncountable, we have Ωσ AK ΩAK for every countable implicit signature σ; Ωκ AJ = ΩAJ [Alm95, Section 8.1]; Ωκ AG is the free group freely generated by ι(A) = A; Ωκ ACR is the free completely regular (union of groups) semigroup freely generated by ι(A) = A. 88 / 332 A key problem for the applications is to be able to solve the word problem in the free σ-algebra Ωσ AV: to find an algorithm, if one exists, that given two σ-terms over A, determines whether uV = vV. If such an algorithm exists, then we say that the word problem is decidable; otherwise, we say that it is undecidable. 89 / 332 Examples: The word problem for Ωκ AN: two κ-terms coincide in Ωκ AN if and only if they are equal or they both involve the operation ω−1 . The word problem for Ωκ AG is well known: the operation ω−1 is inversion in profinite groups, so all κ-terms can be effectively reduced (over G) to κ-terms in which that operation is only applied to letters; then use, in any order, the reduction rules aaω−1 → 1 and aω−1 a → 1 (a ∈ A) to obtain a canonical form for κ-terms over G; two κ-terms are equal over G if and only if they have the same canonical form. Word problem for Ωκ AK: exercise. The solution of the word problem for Ωκ AJ = ΩAJ gives the structure of ΩAJ [Alm95, Section 8.1]. The word problem for Ωκ ACR has been solved by Kad’ourek and Pol´ak [KP86]. The word problem for Ωκ AA has been solved by McCammond [McC01]. 90 / 332 Part II Separating words and regular languages 91 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 92 / 332 A separation problem Let V be a pseudovariety of semigroups. Suppose that a regular language L ⊆ A+ and a word w ∈ A+ are given. How do we find out whether a proof that w /∈ L exists using V-recognizable languages? More precisely, we wish to decide whether, given such L and w, there exists a V-recognizable language K ⊆ A+ such that L ⊆ K and w /∈ K. For instance, how do we determine whether there exists a finite permutation automaton such that no word from L ends in the same state as w does? Another example of the same type of problem: is there some integer n such that no word from L has the same subwords of length at most n as w does? 93 / 332 Our problem sounds like a topological separation problem, and indeed it admits such a formulation in the profinite world. Proposition 10.1 Let V be a pseudovariety of semigroups, L ⊆ A+ a regular language and w a word in A+. Then there is a V-recognizable language K ⊆ A+ such that L ⊆ K and w /∈ K if and only if ιV(w) does not belong to the closure of ιV(L) in ΩAV. Proof. By Proposition 5.4, the condition ιV(w) belongs to the closure ιV(L) in ΩAV holds if and only if every clopen subset of ΩAV which contains ιV(w) has nontrivial intersection with ιV(L). By Theorem 5.9, such clopen subsets are precisely the sets of the form ιV(K) where K is a V-recognizable subset of A+ . It remains to observe that, ιV(w) ∈ ιV(K) and ιV(K) ∩ ιV(L) = ∅ if and only if w ∈ K and K ∩ L = ∅, which follows from the facts that K = ι−1 V (ιV(K)) and L ⊆ ι−1 V (ιV(L)). 94 / 332 Corollary 10.2 Let V be a pseudovariety containing N. If w is a word in A+ and L ⊆ A+ is a regular language, then w and L can be separated by a V-recognizable language if and only if w /∈ L. Proof. By Proposition 7.3, ι embeds A+ in ΩAV as a discrete subspace, meaning that every subset is open. Hence, if w belongs to the closure L of L in ΩAV, then w ∈ L. More directly, in case N ⊆ V, the language {w} is V-recognizable and therefore there is a V-recognizable language containing w and disjoint from L if and only if w /∈ L. Note that the condition N ⊆ V is equivalent to V satisfying some pseudoidentity of the form xω+n = xn, where n is a positive integer. 95 / 332 A refined separation separation problem Even for pseudovarieties V containing N, the separation problem becomes nontrivial if we wish instead to separate two regular languages by a V-recognizable language. The following result can be proved basically by the same argument as presented for the proof of Proposition 10.1, taking additionally into account that ΩAV is compact. The details are left as an exercise. Proposition 10.3 Let V be a pseudovariety of semigroups and L1, L2 ⊆ A+ regular languages. Then there is a V-recognizable language K ⊆ A+ such that L1 ⊆ K and L2 ∩ K = ∅ if and only if the closures of ιV(L1) and ιV(L2) in ΩAV are disjoint sets. 96 / 332 In the case of A, Henckell has constructed an algorithm, working directly in a finite semigroup which recognizes simultaneously two given regular languages, which decides whether they may be separated by a star-free language [Hen88]. An idea due to Pin and Reutenauer [PR91] in the case of the pseudovariety G of all finite groups is to somehow“compute” the closure of ιV(L) not in ΩAG but in the free group Ωκ AG, or even in A+. Under the assumption of a conjectured property for the pseudovariety G, they produced an algorithm for computing the required closure, which solves our problem for G. We proceed to introduce the required property in general, returning later to their algorithm. 97 / 332 For a subset L of A+, denote by clσ,V(L) and clV(L) respectively the closure of ιV(L) in Ωσ AV and in ΩAV. Note that clσ,V(L) = clV(L) ∩ Ωσ AV. In particular, for w ∈ A+, we have ι(w) ∈ clV(L) if and only if ι(w) ∈ clσ,V(L). Denote by pV the natural continuous homomorphism ΩAS → ΩAV. Since ΩAS is compact and pV is an onto continuous mapping, we always have the equality clV(L) = pV(clS(L)). In general, for a continuous function f : S → T, and a subset X of S, we have f (X) ⊆ f (X). The reverse inclusion also holds if f is onto and S is compact. 98 / 332 σ-fullness We say that the pseudovariety V is σ-full if, for every regular language L ⊆ A+, the following equality holds: clσ,V(L) = pV(clσ,S(L)). In other words, membership of w ∈ Ωσ AV in clσ,V(L) is witnessed by some w ∈ clσ,S(L) such that pV(w ) = w. 99 / 332 Theorem 10.4 ([AS00a]) Let V be a pseudovariety, A a finite alphabet and σ an implicit signature such that the following conditions hold: (1) V is σ-full; (2) the word problem for Ωσ AV is decidable; (3) σ is recursively enumerable; (4) V is recursively enumerable; (5) for each operation in σ, there is an algorithm to compute it in any given finite semigroup. Then it is decidable whether, given a regular language L ⊆ A+ and a pseudoword w ∈ Ωσ AV, we have w ∈ clσ,V(L). Proof sketch. Enumerate the pairs (w, L) for which w ∈ clσ,V(L) holds using the assumptions (1)–(3). To enumerate those for which the condition fails, enumerate 4-tuples (ϕ, ψ, X, w) where ϕ : ΩAS → S and ψ : ΩAS → T are onto continuous homomorphisms with S ∈ S and T ∈ V, X ⊆ S, and w ∈ Ωσ AS are such that X × {ψ(w)} ∩ Im(ϕ × ψ) = ∅ and output the corresponding pairs (w, ϕ−1 (L) ∩ A+ ). This requires the assumptions (4) and (5). 100 / 332 Examples: The pseudovariety N is κ-full: for a regular language L ⊆ A+ and a κ-term w, wN ∈ clκ,N(L) if and only if w is a word from L or w involves the operation ω−1 and L is infinite; in the latter case, by compactness there is some κ-term v such that vS ∈ clκ,S(L) \ A+ and so wN = 0 = pN(vS). That the pseudovariety J is κ-full follows from the structure theorem for ΩAJ [Alm95, Section 8.2]. The pseudovariety G is κ-full: the essential ingredient is a seminal theorem of Ash [Ash91]; the details follow from [AS00a] and [Del01]. The pseudovariety Ab is κ-full [Del01]. 101 / 332 The pseudovariety Gp is not κ-full: this follows from a weak version of Ash’s theorem proved by Steinberg [Ste01] for Gp together with the fact that the conjunction of this weaker property with κ-fullness implies that the pseudovariety is defined by pseudoidentities in which both sides are given by κ-terms [AS00a] (cf. Proposition 11.1); however, such a definition does not exist since, by a theorem of Baumslag [Bau65], the free group is residually a finite p-group. That the pseudovarieties A and R are κ-full has been proved by JA-JCCosta-MZeitoun using the solution of the word problems for Ωκ AA [McC01]4 and Ωκ AR [AZ07]. 4 plus refinements from an alternative proof obtained by the same authors including the fact that Ωκ AA is closed for taking factors in ΩAA. 102 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 103 / 332 Separation in σ-algebras Let σ be an implicit signature. Let V be a pseudovariety of semigroups. Note that, if two regular languages K, L ⊆ A+ have disjoint closures in ΩAV then they also have disjoint closures in Ωσ AV. If the converse also holds, then we say that V is weakly σ-reducible for the separation problem. This is a special case of a more general weak reducibility property introduced in [AS00a]. The property is considered in general for an arbitrary system of equations, the present case being that of the equation x = y. Naturally, there is also a stronger form of that property, which we are not considering in these lectures. For a σ-full pseudovariety, the two versions of the property are equivalent [AS00a, AS00b]. 104 / 332 Among others, the following pseudovarieties are known to be weakly κ-reducible for the separation problem: G [Ash91]; CR [AT01]; Gp, with p prime [Ste01]; A [JA-JCCosta-MZeitoun]; R [AS01]; J (follows from the solution of the word problem for ΩAJ [Alm95, Section 8.2]); LSl [CT04]. 105 / 332 Say that a pseudovariety V is σ-equational if it admits a definition by pseudoidentities involving only σ-operations. Such pseudoidentities are also called σ-identities. Proposition 11.1 ([AS00a]) If a pseudovariety V is weakly σ-reducible for the separation problem and σ-full, then V is σ-equational. 106 / 332 Proof. Let Σ be the set of all σ-identities which are valid in V. Clearly V ⊆ [[Σ]]. Let S be a finite semigroup S that satisfies Σ. By Reiterman’s Theorem 6.2, to show that S ∈ V, it suffices to establish that S satisfies every pseudoidentity which is valid in V. Consider such a pseudoidentity u = v, say with u, v ∈ ΩAS, and let ϕ : ΩAS → S be a continuous homomorphism. We claim that ϕ(u) = ϕ(v). Let K = ϕ−1(ϕ(u)) ∩ A+ and L = ϕ−1(ϕ(v)) ∩ A+. Note that, since u ∈ clS(K), v ∈ clS(L), and pV(u) = pV(v), we have clV(K) ∩ clV(L) = ∅. Since V is weakly σ-reducible for the separation problem, there is some w ∈ clσ,V(K) ∩ clσ,V(L). Since V is σ-full, there are w1 ∈ clσ,S(K) and w2 ∈ clσ,S(L) such that pV(w1) = w = pV(w2). Hence w1 = w2 is a pseudoidentity from Σ, which therefore is valid in S. Hence, we have ϕ(u) = ϕ(w1) = ϕ(w2) = ϕ(v), which establishes the claim. 107 / 332 Theorem 11.2 ([AS00a]) Let V be a pseudovariety, A a finite alphabet and σ an implicit signature such that the following conditions hold: (1) V is σ-full; (2) V is weakly σ-reducible for the separation problem; (3) the word problem for Ωσ AV is decidable; (4) σ is recursively enumerable; (5) V is recursively enumerable; (6) for each operation in σ, there is an algorithm to compute it in any given finite semigroup. Then it is decidable whether two given regular languages L1, L2 ⊆ A+ may be separated by a V-recognizable language K ⊆ A+. 108 / 332 Proof. Using property (5), we may enumerate the pairs (L1, L2) of regular languages over A+ that may be separated by V-recognizable languages, by simply enumerating triples (L1, L2, ϕ), with ϕ : A+ → S a homomorphism onto an arbitrary semigroup from V, and test whether ϕ separates L1 and L2. In the affirmative case, output (L1, L2). To enumerate the pairs that may not be separated, we use the assumption (2) plus Proposition 10.3, which guarantees that, if L1 and L2 cannot be separated by a V-recognizable language, then there is some w ∈ clσ,V(L1) ∩ clσ,V(L2) which witnesses the non-separability. Each condition w ∈ clσ,V(Li ) may be effectively tested by Theorem 10.4 and the candidates for witnesses may be recursively enumerated by (4). We may thus proceed as follows: enumerate all triples (L1, L2, w) with L1, L2 ∈ A+ regular languages and w ∈ Ωσ AV; for each such triple, test whether w ∈ clσ,V(Li ) (i = 1, 2); output the pairs (L1, L2) which pass the test as non-separable by V-recognizable languages. 109 / 332 The algorithms described in Theorems 10.4 and 11.2 are purely theoretical, being unusable in practice. Thus, it is worth, particularly in cases where the decidability of the separation property is already guaranteed by Theorem 11.2, to find efficient algorithms to test separability. 110 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 111 / 332 pro-V metrics The same way we defined a pseudo-ultrametric on the free semigroup A+ associated with a pseudovariety V, we may define a pseudo-ultrametric on an arbitrary semigroup S: let d(s1, s2) = 2−r(s1,s2) , where r(s1, s2) is the smallest cardinality of a semigroup T ∈ V for which there is a homomorphism ϕ : S → T such that ϕ(s1) = ϕ(s2). Similar arguments show that d is indeed a pseudo-ultrametric on S, with respect to which the multiplication in S is uniformly continuous. If S is finitely generated, then the completion ˆS is again a pro-V semigroup, but it may not be a free pro-V semigroup. The pseudo-ultrametric d is an ultrametric if and only if S is residually in V. Every homomorphism S → T into T ∈ V is uniformly continuous. 112 / 332 Let V be a pseudovariety of semigroups, and let σ be an implicit signature. Considering the members of V as σ-algebras under the natural interpretation of the operations from σ, since the homomorphisms are the same, V is still a pseudovariety of σ-algebras. We may define a pseudo-ultrametric on a σ-algebra S similarly to the semigroup case: let dσ (s1, s2) = 2−rσ(s1,s2) where rσ(s1, s2) is the smallest cardinality of a member T ∈ V for which there is a homomorphism ϕ : S → T of σ-algebras such that ϕ(s1) = ϕ(s2). 113 / 332 The delicate point here is that, while continuous semigroup homomorphisms between profinite semigroups respect implicit operations, and in particular so do semigroup homomorphisms between finite semigroups, semigroup homomorphisms between arbitrary σ-algebras may not be homomorphisms of σ-algebras. Example: let U1 be two-element semilattice, which the multiplicative subsemigroup of Z consisting of 0, 1; consider the mapping ϕ : ΩAN → U1 that maps A+ to 1 and everything else to 0. Then ϕ is a semigroup homomorphism but not a homomorphism of κ-algebras. 114 / 332 Proposition 12.1 For a pseudovariety V and an implicit signature σ, the completion of Ωσ AV with respect to the pseudo-ultrametric dσ is ΩAV, both metrically and algebraically. Proof. Since the algebraic structure is inherited by extension of uniformly continuous operations, it suffices to show that, for u, v ∈ Ωσ AV, dσ (u, v) = d(u, v), where d is the completion metric on ΩAV. By Proposition 4.6, d(u, v) = 2−r , where r is the smallest cardinality of a member T ∈ V for which there exists a continuous homomorphism ϕ : ΩAV → T such that ϕ(u) = ϕ(v). Since the restriction of ϕ to Ωσ AV is a homomorphism of σ-algebras which separates u from v, it follows that rσ (u, v) ≤ r. On the other hand, by the universal property of ΩAV and since we interpret σ-operations naturally, every homomorphism of σ-algebras Ωσ AV → T ∈ V extends uniquely to a continuous homomorphism ΩAV → T. Hence rσ (u, v) = r. 115 / 332 Pro-H metric on groups Traditionally, one denotes by H an arbitrary pseudovariety of groups. Because a group is highly symmetrical, the pro-H metric structure looks similar everywhere. Lemma 12.2 Let G be a group and consider the pro-H metric on G. Then, for every u, v, w ∈ G, the equalities d(uw, vw) = d(u, v) = d(wu, wv) hold. In particular, for > 0, we have Bε(u) = uBε(1) = Bε(1)u and a subset X is open (respectively closed) if and only if so is Xw. Moreover, for > 0, the ball B (1) is a clopen normal subgroup of G such that G/B (1) ∈ H. A subgroup H is open if and only if it contains some open ball B (1). Proof. This is a simple exercise. 116 / 332 For a subgroup H of a group G, denote by HG the largest normal subgroup of G which is contained in H. It is given by the formula HG = g∈G g−1 Hg. If we let G act on the set of right cosets of H in G by right translation, then we obtain a homomorphism ϕ : G → SG/H into the full symmetric group SG/H (of all permutations of the set G/H) such that ϕ−1(id) = HG . It follows that, if the index (G : H) of the subgroup H in G is finite, then so is (G : HG ) and (G : HG ) is a divisor of (G : H)!. 117 / 332 Lemma 12.3 A subgroup H of G is (cl)open in the pro-H metric if and only if G/HG ∈ H. Proof. Suppose first that H is open. By Lemma 12.2, H contains a normal subgroup K of G such that G/K ∈ H. Then K ⊆ HG and so G/HG (G/K)/(HG /K) belongs to H. Conversely, if G/HG ∈ H then HG is an open set, because the natural homomorphism G → G/HG is (uniformly) continuous. Since H contains HG , H is a union of cosets of HG , and so is its complement. Hence H is clopen. 118 / 332 Another natural question is whether, for a subgroup H of G, the intersection with H of an open subset of G in the pro-H metric of G is also open in the pro-H metric of H. In general, the answer is negative, but there are important situations in which it is affirmative. Example 12.4 Let G be the free group on two free generators a, b and consider the homomorphism ϕ : G → S3 defined by ϕ(a) = (12) and ϕ(b) = (13). Let K = ϕ−1(1) and let H = ϕ−1 (123) be the inverse image of the subgroup of index 2. Then H is clopen in the pro-Ab metric of G and K is clopen in the pro-Ab metric of H but K is not clopen in the pro-Ab metric of G. 119 / 332 Note that, for pseudovarieties of groups K and H, K ∗ H consists of all groups G which have a normal subgroup K such that both K ∈ K and G/K ∈ H.5 If H ∗ H = H, then we say that H is closed under extension. A condition for the answer to the above question to be affirmative is drawn from the following result. Lemma 12.5 Let H be a clopen subgroup of G in the pro-H metric of G and suppose that U is a normal subgroup of H such that H/U ∈ H. Then the normal subgroup UG of G is such that G/UG ∈ H ∗ H. 5 For those unfamiliar with semidirect products, take this as the definition of K ∗ H and show that it is a pseudovariety of groups. 120 / 332 Proof. Consider also the normal subgroup HG and let g ∈ G. By Lemma 12.3, G/HG belongs to H. For each x ∈ HG , the conjugate gxg−1 belongs to H and so the mapping ϕg : HG → H/U which sends x to gxg−1U is a group homomorphism. Moreover, for x ∈ HG , we have x ∈ UG ⇔ x ∈ g−1 Ug for all g ∈ G ⇔ gxg−1 ∈ U for all g ∈ G ⇔ ϕg (x) = 1 for all g ∈ G. It follows that HG /UG embeds in a finite power of H/U and so HG /UG ∈ H. The result now follows from the observation that G/HG (G/UG )/(HG /UG ). 121 / 332 A first application of the preceding lemma is the following answer to the above question. Proposition 12.6 Suppose that H is closed under extension. Let H be a clopen subgroup of G in the pro-H metric of G. Then a subset of H is open in the pro-H metric of H if and only if it is open in the pro-H metric of G. Proof. By Lemma 12.2, a subgroup L of H is open in the pro-H metric of H if and only if it contains a normal subgroup U of H such that H/U ∈ H. By Lemma 12.5, the normal subgroup UG of G is such that U/UG ∈ H ∗ H = H. Hence U is open in the pro-H metric of G by Lemma 12.3. Since L is a union of cosets of U, L is also open in the pro-H metric of G. 122 / 332 In terms of pro-H metrics, we obtain the following more precise result. Proposition 12.7 Suppose that H is closed under extension and G is a group residually in H. Let H be a clopen subgroup of G in the pro-H metric of G. Then the pro-H metric dH of H and the restriction to H of the pro-H metric dG of G have the same Cauchy sequences. 123 / 332 Proof. Let d be the restriction of dG to H and let r be the corresponding partial function H × H → N. Denote by d the pseudo-metric dH and by r the corresponding partial function. We start by establishing the following function inequalities: r ≤ r ≤ (G : H) · r !. (1) The first inequality in (1) follows from the observation that, if a homomorphism from G into a member of H distinguishes two elements of H then its restriction to H also distinguishes them. Suppose next that u, v ∈ H and the homomorphism ϕ : H → K with K ∈ H are such that ϕ(u) = ϕ(v). Let U = ϕ−1 (1). Then H/U embeds in K and, therefore, it belongs to H. By Lemma 12.5, UG is a normal subgroup of G of finite index such that G/UG ∈ H ∗ H = H and, by an earlier observation, (G : UG ) divides (G : U)!. If we choose above K so that |K| is minimum, then (H : U) = r (u, v) and so, since uUG = vUG , r(u, v) ≤ (G : UG ) ≤ (G : U)! = (G : H) · (H : U) ! = (G : H) · r (u, v) ! which proves (1). 124 / 332 (. . . ) From the first inequality in (1) we deduce that every Cauchy sequence with respect to d is also a Cauchy sequence with respect to d. For the converse, let f (n) = (G : H) · n !. Then f is an increasing sequence and a simple calculation shows that, for every ε > 0, d ≤ 2−f ( − log2 ε ) =⇒ d ≤ ε. This implies that Cauchy sequences for d are also Cauchy sequences for d . 125 / 332 Free products A free product in a variety V of semigroups is given by two homomorphisms ϕi : Si → F (i = 1, 2), with S1, S2, F ∈ V such that, given any other pair of homomorphisms ψi : Si → T, with T ∈ V, there exists a unique homomorphism θ : F → T such that the following diagram commutes: F θ 22 S1 ϕ1 oo ψ1  S2 ϕ2 yy ψ2 GG T By the usual argument, if the free product exists, then it is unique up to isomorphism. 126 / 332 Exercise 12.8 Show that, for every variety V and semigroups S1, S2 ∈ V, the free product of S1 and S2 in V exists. For semigroups S and T in a variety V, we say that S is a free factor of T if there exists U ∈ V such that T is a free product of S and U in V. Note that every semigroup is a free factor of itself. Exercise 12.9 Suppose that S is a free factor of T in the variety V generated by a pseudovariety V. Show that: 1. the pseudo-metric dS V and the restriction of the pseudo-metric dT V to S coincide; 2. the open sets in pro-V metric of S are the intersection with S of the open sets of T in the pro-V metric of T. 127 / 332 Theorem 12.10 Let H be a finitely generated subgroup of a finitely generated free group G in the variety generated by the extension-closed pseudovariety H and suppose that H is a free factor of a clopen subgroup of G in the pro-H metric. Then the following hold: 1. the pro-H metric dH of H has the same Cauchy sequences as the restriction of the pro-H metric dG ; 2. the completion of H with respect to dG is a free pro-H group. 128 / 332 Proof. We may as well assume that H is not the trivial pseudovariety for, otherwise, the result is obvious. Hence H contains some cyclic group of prime order Z/pZ. Since H is closed under extension, it follows that Gp ⊆ H for some prime p. Since the free group is residually a finite p-group by a result of Baumslag [Bau65], we may therefore assume that G is an absolutely free group. By the Nielsen-Schreier Theorem, the subgroup H is also an absolutely free group. By Proposition 12.1, the completion of H with respect to the metric dH is a free pro-H group. Since, by Proposition 12.7 and Exercise 12.9, the Cauchy sequences of H are the same with respect to both metrics dH and dG , the completion of H with respect to both metrics is the same metric group and so it is a free pro-H group. 129 / 332 The following corollaries will be useful later. Corollary 12.11 Let H be an extension-closed pseudovariety of groups and let G be a finitely generated free group in the variety generated by H. If the finitely generated subgroup H is a free factor of a clopen subgroup of G in the pro-H metric, then H is closed in that metric. The following result is known as the subgroup theorem in the theory of profinite groups. For simplicity, it is presented here only in the finitely generated case, although the proof could be extended to the general case. Corollary 12.12 Let H be a clopen subgroup of a finitely generated free pro-H group G, where H is an extension-closed pseudovariety of groups. Then H is itself a free pro-H group. 130 / 332 Proof of Corollary 12.11. Since the pro-H metric of a clopen subgroup is the induced metric, we may as well assume that H is a free factor of G. Let (hn)n be a sequence of elements of H and suppose that it converges in G to an element g. Suppose H is another subgroup of G such that G is the free product of H and H . Then, by considering the identity mapping on H and sending H to 1 we obtain an onto homomorphism ϕ : G → H which is obviously continuous with respect to the pro-H metrics of G and H. Hence (hn)n already converges in H, namely to ϕ(g). Since the metrics dH and dG have the same Cauchy sequences in H by Theorem 12.10, the limit must be the same in both cases, and so g ∈ H. This shows that H is closed in the pro-H metric of G. Proof of Corollary 12.12. Let A be a finite free generating set for G. Let G be the subgroup of G discretely generated by A. Then G is a free group over A in the variety generated by H and it is a dense subgroup of G. Let H = H ∩ G . Then H is a clopen subgroup of G in the pro-H topology of G . Since H is open, H is dense in H. Hence H is the completion of H with respect to the metric dG . Applying Theorem 12.10 to the free factor H of itself, it follows that H is a free pro-H group. 131 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 132 / 332 The free group and inverse automata Let now G be a free group on a finite set A. To each finite subset X of G we associate a finite inverse automaton as follows.6 For each reduced group word representative w of an element of X, consider a linear graph with directed edges labeled by the successive letters of w, the edge appearing in the direction of left-to-right reading of w if the corresponding generator appears with exponent 1 and in the opposite direction if the exponent is −1. Thus the label of the undirected path which traverses the graph in the direction of left-to-right reading of w is w. Glue these graphs together by identifying their ends to a single vertex v0 which is the unique initial and final state of the automaton. Fold edges so that the resulting automaton becomes inverse in the sense that the transformations defined by the labels are partial bijections. This can be done by applying the following procedure, in an arbitrary order, until it no longer applies: whenever we encounter two edges with the same label leaving from the same state or arriving at the same state, we identify them. 6 See [Sta83, MSW01, KM02] for details and further references. 133 / 332 Example 1 Let G be the free group on the free generators a, b, c and let X = {ab−1c−1a, a−1b−1ac−1a, bc−1a}. Then the sequence of pictures in Figure 1 describes the construction of the folded inverse automaton associated with X. ◦ ◦ ◦ ◦ ◦ ◦ a b c a a b a c a b c a ◦◦ ◦ a b c aa b a c a b c a ◦ ◦ a b caa b a c a b c a ◦ ◦ a b c b a c a b c a ◦ ◦ ab c a b a c b c ab c a b a b Figure: The folding procedure 134 / 332 The automaton resulting from the above procedure is reduced in the sense that it has a unique initial state which is simultaneously the unique final state, every state is accessible from it (through an undirected path), and there is no state of degree 1 other than possibly the initial and final state. The set of reduced group words recognized by the automaton is precisely the subgroup X generated by X. Moreover, this automaton is unique up to isomorphism and it depends only on the subgroup X and not on the specific generating set X. Thus, for a finitely generated subgroup H of the free group G on a set A, we will denote its automaton by A(H). Conversely, a given finite reduced inverse automaton A over the alphabet A recognizes a finitely generated subgroup of the free group on A whose associated automaton is A. This subgroup is the fundamental group of the underlying graph. 135 / 332 The following result summarizes the properties of the construction of the automaton associated with a finitely generated subgroup of a free group. Theorem 13.1 The correspondence which maps each finitely generated subgroup H of the free group G on the set A to its automaton A(H), constructed from a finite generating set of H, has the following properties: 1. the automaton A(H) is effectively constructible and does not depend on the generating set but only on H; 2. the correspondence defines a bijection between the set of all finitely generated subgroups of G and the set of all finite reduced inverse automata with input alphabet A. 136 / 332 Example 2 We recall the procedure to compute the fundamental group of a connected labeled directed graph with a root vertex, namely to exhibit a finite generating set for that subgroup. First, choose a generating tree, that is a maximal subgraph whose underlying undirected graph is a tree. Note that every vertex of the original graph is on the tree. To each edge e which is not on the chosen tree, associate the group word which is obtained by reading the label of the (undirected) path which goes from the root along the tree to the beginning vertex of e, then follows e, and then returns to the root along the tree. The set of all words associated in this way to the edges which are not on the tree generates a subgroup of the free group on the labels which is called the fundamental group of the labeled graph. This procedure is shown here for the inverse automaton obtained in the preceding example. The resulting generating set for the same subgroup, namely the subgroup recognized by the inverse automaton, is {a−1 ca−1 ba, ba−1 ba, ab−1 a−1 ba}. The set thus obtained freely generates the subgroup. This also gives a procedure to compute free generators for the subgroup of a free group generated by a given finite set of group words. ab a b a c b ab a b a c b ab a b a c b a−1 · c · a−1 ba ab a b a c b b · a−1 ba ab a b a c b a · b−1 a−1 ba 137 / 332 We proceed to explore some other relationships between inverse automata and finitely generated subgroups of free groups. Let A and B be two reduced inverse automata with the same input alphabet A. A morphism f : A → B is a function which maps states of A to states of B, sending the initial state of A to that of B, in such a way that the action of A is respected: if q is a state of A, a ∈ A ∪ A−1 , and qa is defined, then f (q)a is also defined and f (q)a = f (qa). Note that if such a morphism exists then it is unique. Proposition 13.2 Let G be a free group and let H and K be finitely generated subgroups of G. Then H ⊆ K if and only if there is an automaton morphism from A(H) into A(K). Moreover, if the morphism is injective then H is a free factor of K. 138 / 332 Proof. Suppose first that there exists an automaton morphism A(H) → A(K). A reduced group word w representing an element of H labels a successful path in A(H). By transforming this path by the morphism, we obtain a successful path in A(K), which shows that w also represents an element from K. Hence H ⊆ K. Conversely, suppose that H ⊆ K. We may choose finite sets X and Y of reduced words such that X = H and X ∪ Y = K. Then, the construction of A(H) and A(K) proceeds from sets of linear automata in which the first is contained in the second. This implies that every identification which is made to obtain A(H) will also occur in the construction of A(K). Hence there is a morphism A(H) → A(K). Suppose next that there is an injective morphism A(H) → A(K). Choose for A(H) a generating tree. Its image in A(K) is still a tree and so it can be expanded to a generating tree of A(K). Hence there is a free generating set for K which contains a free generating set for H and this is equivalent to H being a free factor of K. 139 / 332 Let A be an inverse automaton. A congruence on A is an equivalence relation ∼ on the state set of A which is compatible with the action of the input alphabet A: if q1, q2 are states, a ∈ A ∪ A−1, q1a, q2a are both defined, and q1 ∼ q2, then q1a ∼ q2a. We may then consider the quotient automaton A/∼, whose states are the ∼-classes of states of A and such that the action of A is given by (q/∼)a = (qa)/∼. Note that A/∼ is an inverse automaton. 140 / 332 Lemma 13.3 Let H and K be finitely generated subgroups of the free group on a finite set A and suppose that H ⊆ K. For each state p, choose a reduced word up which labels a path in A(H) from the initial state to p. Then the congruence ∼H,K on A(H) determined by the kernel of the unique morphism ϕ : A(H) → A(K) satisfies the following condition: for all states p, q, p ∼H,K q if and only if upu−1 q ∈ K. (2) Moreover, if L is the fundamental group of the quotient automaton A(H)/∼H,K , then H ⊆ L and L is a free factor of K. 141 / 332 Proof. If upu−1 q ∈ K, then up and u−1 q label paths in A(K) from the initial-final state to the states ϕ(p) and ϕ(q). Since A(K) is an inverse automaton which recognizes K, it follows that ϕ(p) = ϕ(q), that is p ∼H,K q. Conversely, if p ∼H,K q, then ϕ(p) = ϕ(q) and so the reduced form of upu−1 q labels a loop at the initial-final state in A(K), that is upu−1 q ∈ K. This proves condition (2). The last part of the statement of the lemma follows from Proposition 13.2. 142 / 332 Proposition 13.4 Let H be a finitely generated subgroup of the finitely generated free group G on a finite set A, which is endowed with the pro-V topology. Then the following hold: 1. if H is a clopen normal subgroup then A(H) is a permutation automaton whose transition monoid is isomorphic to G/H and belongs to V; 2. H is clopen if and only if A(H) is a permutation automaton whose transition monoid belongs to V. 143 / 332 Proof. By Lemma 12.3, if H is a clopen normal subgroup then G/H ∈ V. The Cayley graph of the finite group G/H with respect to A can be viewed as a reduced inverse automaton which recognizes H as the set of reduced group words which label loops at the vertex 1. Hence it is isomorphic to A(H) and A(H) is a (complete) permutation automaton. The states in the Cayley graph are precisely the cosets of H, on which the input alphabet acts by right translation. Hence the transition monoid is isomorphic to G/H and it belongs to V. Next we assume only that H is clopen. By Lemma 12.3, HG is a clopen normal subgroup. The quotient automaton A(HG )/∼HG ,H embeds in A(H). Since A(HG ) is a complete automaton by (1), so is its quotient A(HG )/∼HG ,H . Since A(H) is a reduced inverse automaton, it follows that A(H) is also a permutation automaton as there is no room in A(HG )/∼HG ,H to add vertices or edges. The transition monoid of A(H) is therefore a quotient of that of A(HG ), which is isomorphic to G/HG by (1), and therefore it belongs to V. 144 / 332 (. . . ) Conversely, suppose that A(H) is a permutation automaton whose transition monoid M belongs to V and consider the natural homomorphism ϕ : G → M and K = ϕ−1 (1). Then K consists of all reduced words which map every state of A(H) to itself while H consists of all reduced words which map the state 1 to itself. Hence K ⊆ H and H is clopen since K is clopen by Lemma 12.3. 145 / 332 A simple application of Proposition 13.4 is the following theorem due to M. Hall [Hal50] in a paper which first introduced profinite topologies on free groups. Theorem 13.5 Every finitely generated subgroup of a free group G on a finite set A is closed in the profinite metric of G. 146 / 332 Proof. Let H be a finitely generated subgroup of G. If H = G, then of course H is closed. Otherwise, it suffices to show that, for every reduced word g ∈ G \ H, there is a clopen subgroup which contains H but not g. Consider the automaton A(H) and add to it, starting from the initial-final state, a linear automaton which reads g. By applying the folding procedure, we end up with an inverse automaton, which may not be reduced, in which the end state of the added linear automaton is not identified to the initial-final state. Each generator a ∈ A determines in this automaton a partial permutation of the state set. Now, every partial permutation of a finite set may be completed to a full permutation and we choose such a completion for each generator a ∈ A. This leads to a permutation automaton B in which A(H) embeds. By Proposition 13.4, the fundamental group K of B is a clopen subgroup of G. By Proposition 13.2, K contains H. On the other hand, g /∈ K since in B the reduced word g does not label a loop at the initial-final state. 147 / 332 Note 13.6 An alternative proof of Theorem 13.5 is obtained as follows. Complete the automaton A(H) of the given finitely generated subgroup of G to a permutation automaton B, with the same initial-final state. By Proposition 13.4, the fundamental group K of B is a clopen subgroup. By Proposition 13.2, H is contained in K. By Corollary 12.11, since G is extension closed, H is closed. 148 / 332 Example 3 In Example 1 we obtained the first inverse automaton in Figure 2 1 2 3 45 ab a b a c b 1 2 3 45 ab a b a c b a a b b c c c c Figure: An inverse automaton and one of its completions which gives the partial permutations defined in the table on the left and one of their possible completions to permutations in the table on the right 1 2 3 4 5 a 2 − − 1 3 b 3 − 2 − 4 c − − − 3 − 1 2 3 4 5 a 2 4 5 1 3 b 3 1 2 5 4 c 1 2 4 3 5 and the corresponding permutation automaton in Figure 2. This permutation automaton determines a clopen subgroup K, in the profinite metric of the free group, of which the original subgroup is a free factor. 149 / 332 More generally, we say that a finitely generated subgroup H of the free group G on the finite set A is V-extendible if A(H) can be extended to a permutation automaton whose transition monoid belongs to V. In such an extension, both states and transitions may be added. Lemma 13.7 The finitely generated subgroup H of the free group G on a finite set A is V-extendible if and only if there is a clopen subgroup K of G, in the pro-V metric of G, such that the congruence ∼H,K on A(H) is the equality. Proof. If H is V-extendible, then there is a permutation automaton B containing A(H), with the same initial-final state, whose transition monoid belongs to V. By Proposition 13.4, the fundamental group K of B is clopen in the pro-V metric of G. Then A(H) embeds in A(K) and so, by definition of the congruence ∼H,K (cf. Lemma 13.3), this congruence is the equality relation on the state set of A(H). The converse is proved similarly. 150 / 332 By definition of pro-V metric, the closure of a subgroup of a group G is the intersection of the clopen subgroups that contain it. In the case of finitely generated subgroups of a finitely generated free group, we can use the previous results to obtain a more precise statement. Proposition 13.8 Let H be a finitely generated subgroup of the free group G on a finite set A, which is endowed with the pro-V metric, and let H be the closure of H in G. Then the following hold: 1. there is a clopen subgroup K of G such that ∼H,H coincides with ∼H,K ; 2. there is a smallest V-extendible subgroup containing H, namely the subgroup ˜H such that A(˜H) is the image of A(H) in A(H); 3. H ⊆ ˜H ⊆ H and ˜H is a free factor of H. 151 / 332 Proof. By Proposition 13.2, for each clopen subgroup K containing H, there is a morphism A(H) → A(K), which determines a congruence on A(H). The intersection ∼ of all such congruences is still a congruence on A(H). Since the automaton A(H) is finite, the intersection ∼ involves only finitely many congruences and so A(H)/∼ is the largest quotient of A(H) which embeds in a permutation automaton of a clopen subgroup. The fundamental group of A(H)/∼ is therefore a finitely generated subgroup L containing H which is a free factor of the clopen subgroup K, whose automaton A(K) is the smallest permutation automaton in which A(L) embeds and which has a transition monoid in V. This proves the proposition. 152 / 332 The following result reduces the computation of ˜H (and therefore also the question as to whether H is V-extendible) to the membership problem of the pro-V closure H. Proposition 13.9 Let G be a finitely generated free group, endowed with the pro-V metric, and let H be a given finitely generated subgroup of G. If the membership problem for H is decidable then one can effectively compute the smallest V-extendible subgroup ˜H of G containing H. Proof. By Proposition 13.8, it suffices to compute the congruence ∼H,H on the automaton A(H). By condition (2) of Lemma 13.3, this congruence can be computed by the choice of a label uq for a path from the initial-final state to each state q by using the solution of the membership problem for H. Thus, one may concentrate on the membership problem for the closure subgroup H. 153 / 332 Proposition 13.8 is insufficient to guarantee the existence of an algorithm to compute the closure H of a given finitely generated subgroup H. In case V is an extension-closed pseudovariety, we obtain some more precise results. Theorem 13.10 Let H be a finitely generated subgroup of a finitely generated free group, which is endowed with the pro-V metric for an extension-closed pseudovariety V. The following conditions are equivalent: 1. H is closed; 2. H is V-extendible; 3. H is a free factor of a clopen subgroup. Proof. By Proposition 13.8, (1) implies H = ˜H = H which yields (2) since ˜H is V-extendible. Condition (2) implies (3) by Proposition 13.2. Finally, (3) implies (1) by Corollary 12.11. 154 / 332 Corollary 13.11 Under the hypotheses of Theorem 13.10, ˜H = H. Exercise 13.12 Let G be a finitely generated free group and endow it with the pro-V metric. Let H be a finitely generated subgroup of G. Prove the following: 1. for every g ∈ G, the automaton A(g−1Hg) is obtained by modifying A(H) as follows: if the initial-final state of A(H) is q0 then set that of A(g−1Hg) to be the state q0g; 2. H is V-extendible if and only if all its conjugates are V-extendible; 3. H is closed (respectively open) if and only if all its conjugates have the same property. 155 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 156 / 332 Relative invertibility of endomorphisms Given a continuous endomorphism ϕ of ΩnM, we denote by M(ϕ) the n × n-matrix over the profinite ring ˆZ whose (i, j)-coordinate is the image of ϕ(xi ) of the i-th generator of ΩnM under the unique continuous homomorphism ΩnM → ˆZ which maps xj to 1 and all other generators to 0. The determinant and trace of the matrix M(ϕ) are also called, respectively, the determinant and trace of ϕ and are denoted det ϕ and tr ϕ. Exercise 14.1 Let M be a profinite monoid. Show that m ∈ M is invertible if and only if m is right invertible, if and only if m is left invertible, if and only if mω = 1. 157 / 332 For a metric semigroup S, denote by End S the monoid of continuous endomorphisms of S. For a pseudovariety V of monoids and w ∈ ΩnM, denote by wV the restriction of the implicit operation w to V. For ϕ ∈ End ΩnM, denote by ϕV the continuous endomorphism of ΩnV induced by ϕ, which maps the generator xi to (ϕ(xi ))V. Say that ϕ is V-invertible if ϕV is invertible in End ΩnV. Proposition 14.2 A continuous endomorphism ϕ of ΩnM is V-invertible if and only if ΩnV is generated by the set {ϕV(xi ) : i = 1, . . . , n}. 158 / 332 Proof. Let M be the (closed) submonoid of ΩnV generated by {ϕV(xi ) : i = 1, . . . , n}. Suppose first that ϕ is V-invertible. Then (ϕω)V = (ϕV)ω is the identity mapping on ΩnV. Given u ∈ ΩnV, letting w = ϕω−1 V (u), we conclude that u = ϕV(w) = ϕV wΩAV(x1, . . . , xn) = wΩAV ϕV(x1), . . . , ϕV(xn) , where the last equality is justified since implicit operations on V commute with continuous homomorphisms between pro-V semigroups. Since w is the limit of a sequence of words, it follows that u ∈ M, which shows that M = ΩnV. 159 / 332 (. . . ) Conversely, suppose that M = ΩnV. Given u ∈ ΩnV, there exists a sequence of words (wn)n such that u = lim n→∞ (wn)ΩnV ϕV(x1), . . . , ϕV(xn) = lim n→∞ ϕV(wn). Since ΩnM is a compact metric space, we may assume that the sequence (wn)n converges to some w ∈ ΩnM. For such an implicit operation w, we have ϕV(w) = u. In particular, for each generator xi there exists vi ∈ ΩnM such that xi = ϕV(vi ). Let ψ be the continuous endomorphism of ΩnM which maps xi to vi . Then the composite continuous endomorphism ϕV ◦ ψV fixes the generators xi and, therefore it is the identity mapping of ΩnV. Hence ϕV is invertible. 160 / 332 Let Ab denote the pseudovariety of all finite Abelian groups. Let Am denote the pseudovariety of all finite Abelian groups of exponent m. Proposition 14.3 The following conditions are equivalent for a continuous endomorphism ϕ of ΩnM and a prime integer p: 1. ϕ is Gp-invertible; 2. ϕ is Ap-invertible; 3. det ϕ is not divisible by p. 161 / 332 Proof. Since Ap ⊆ Gp, clearly (1)⇒(2). On the other hand, (2)⇔(3) follows from elementary Linear Algebra since ΩnAp is the additive reduct of the n-dimensional vector space over the field Z/pZ. We prove (2)⇒(1) by contraposition and so we assume that ϕ is not Gp-invertible. Then, by Proposition 14.2, the set {ϕGp (xi ) : i = 1, . . . , n} generates a proper closed subgroup H of ΩnGp. Since ΩnGp is a profinite group, it follows that H is contained in some proper clopen subgroup K of ΩnGp. Hence there is a continuous homomorphism ψ : ΩnGp → F onto a finite p-group F such that ψ(H) is the trivial subgroup. We may assume that ψ is the restriction mapping ΩnGp → ΩnAp, from which we deduce that ϕAp is the trivial endomorphism of ΩnAp and, therefore, it is not Ap-invertible. 162 / 332 Exercise 14.4 Show that the following conditions are equivalent for an element u of the ring ˆZ: 1. u is multiplicatively invertible (or a unit); 2. no prime integer divides u; 3. u is invertible in each p-adic completion Zp. Hence, if u ∈ Z, then u is invertible in ˆZ if and only if u = ±1. 163 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 164 / 332 Here is a first result with an algorithmic flavour towards the computation of the closure of a finitely generated subgroup of the free group. Proposition 15.1 Let V be a non-trivial extension-closed pseudovariety of groups and let G be a finitely generated free group which is endowed with the pro-V metric. Suppose that a finitely generated subgroup H of G is not dense in G and let ϕ : G → F be a homomorphism onto F ∈ V such that ϕ(H) F. Then one can compute from ϕ a closed finitely generated free factor L of some clopen subgroup K of G such that A(L) is a quotient of the automaton A(H). 165 / 332 Proof. Put K = ϕ−1 ϕ(H) . Then K is a clopen subgroup of finite index in G. Consider the congruence ∼H,K on the inverse automaton A(H) and let L be the finitely generated subgroup of G such that A(L) A(H)/∼H,K . By Lemma 13.3, L is a free factor of K. By Corollary 12.11, L is a closed subgroup of G. It remains to argue that all constructions are effective. Given the finite group F and the onto homomorphism ϕ, and the free generating set A of G, consider the right action of A on the set F/ϕ(H) of all right cosets of ϕ(H). This defines a permutation automaton which is precisely the automaton A(K). Given H by means of a finite set of generators, one may also construct its automaton A(H). The construction of the quotient automaton A(H)/∼H,K then can be made from the knowledge of both automata A(H) and A(K). The quotient automaton in turn determines the closed subgroup L, for which we may exhibit a finite set of generators. 166 / 332 Definition 15.2 We say that a pseudovariety V of groups has decidable denseness if, given a finite set A and a finite subset B of the free group G on A, it is decidable whether B generates a dense subgroup in the pro-V metric of G. The following theorem summarizes and completes some of the above results. Theorem 15.3 Let V be an extension-closed pseudovariety of groups. Let H be a finitely generated subgroup of the finitely generated free group G and let H be the closure of H in the pro-V metric of G. 1. The group H is finitely generated and a free factor of a clopen subgroup of G. 2. The automaton A(H) is a quotient of A(H). 3. If V is recursively enumerable and has decidable denseness then there is an algorithm to construct a finite set of generators of H. 167 / 332 Proof. By Corollary 13.11, the subgroup H coincides with ˜H, which in turn is finitely generated by Proposition 13.8. The remainder of the statement of (1) is a direct consequence of Theorem 13.10 while (2) now also follows from Proposition 13.8. It remains to prove (3). Let A be a set of free generators for G. Here is how the algorithm proceeds. Since V has decidable denseness, we first check whether H is dense in G. In the affirmative case, we have found H to be G. In the negative case, we know that there is some homomorphism ϕ : G → F onto some group F in V such that ϕ(H) G. Since V is recursively enumerable and G is finitely generated, we may successively enumerate candidates to such homomorphisms until we find one with this property. Once such a homomorphism is found, by Proposition 15.1 one can compute from ϕ a finite set A = {v1, . . . , vr } of free generators for a closed free factor L of some clopen subgroup K of G such that A(L) is a quotient A(H)/∼1 of A(H). 168 / 332 (. . . ) If {w1, . . . , wm} is a set of generators for H, then we may express each wi as a group word wi (v1, . . . , vr ) on the groups words vj . Consider the subgroup H of the free group G on the set A generated by {w1, . . . , wm}. The mapping which sends each free generator vj to itself, viewed as a group word in the given free generators of G, extends uniquely to a continuous homomorphism ψ : G → G which is an isomorphism with L. Since the induced metric on L is the pro-V metric of L by Theorem 12.10, ψ : G → L is a continuous isomorphism with respect to the corresponding pro-V topologies. Since ψ preserves the index of subgroups, as well as quotients for normal subgroups, ψ is an isomorphism of topological groups. Since L is closed in G by Corollary 12.11, the problem of computing the closure of H in G is thus reduced to the computation of the closure of H in G . 169 / 332 (. . . ) Now, if we apply the above procedure to H as a subgroup of G , either H is dense in G , in which case we conclude that H = L, or we obtain a free factor L of a clopen subgroup K of G such that H ⊆ L ⊆ K G and AA (L ) is a quotient of AA (H ). It follows that H ⊆ ψ(L ) ⊆ ψ(K ) L, ψ(L ) is a free factor of the clopen subgroup ψ(K ) (of G). Moreover, it is easy to see that A ψ(L ) is the quotient of A(H) by a congruence ∼2. Since H ⊆ ψ(L ) L, the congruence ∼2 is properly contained in ∼1. Thus, applying the above procedure recursively, in a finite number of steps we must obtain an affirmative answer to the question as to whether the current subgroup is a free factor of the current free group. At that stage, the closure will be computed and then it is a matter of substituting back the free generators to their expressions in the original alphabet A to obtain H. 170 / 332 Corollary 15.4 Let V be an extension-closed pseudovariety of groups with decidable denseness. Then it is decidable whether a given finitely generated subgroup of a finitely generated free group is V-extendible. Since the closure of a finitely generated subgroup is computed by successively taking closed factors of clopen subgroups, combining with Theorem 12.10, we also obtain the following result. Corollary 15.5 Let H be a finitely generated subgroup of a finitely generated free group G and suppose that H is closed with respect to the pro-V metric of G, where V is an extension-closed pseudovariety of groups. Then the completion of H with respect to the restriction to H of the pro-V metric of G is a free pro-V group. 171 / 332 We consider the special case of the extension-closed pseudovariety Gp for a prime integer p, for which the denseness test can be done efficiently. The next result may now be deduced from Proposition 14.2 and Proposition 14.3. For group words w1, . . . , wm on n letters x1, . . . , xn, let M(w1, . . . , wm) be the (integer) matrix whose (i, j)-entry is the exponent of the reduced word in xj which is obtained from wi by replacing by 1 all xk with k = j. In particular, when m = n, M(w1, . . . , wn) is the matrix of the continuous endomorphism of ΩnM which is determined by the implicit operator defined by the group terms w1, . . . , wn. 172 / 332 Proposition 15.6 Let G be a free group on n free generators and let H = w1, . . . , wn be an n-generated subgroup of G. Then H is dense in G if and only if M(w1, . . . , wn) is invertible mod p. Proof. By Baumslag’s Theorem asserting that the free group is residually in Gp, we may regard G as the subgroup of ΩnGp generated by {x1, . . . , xn}. Since the pro-Gp metric of G is the induced metric from the metric of ΩnGp by Proposition 12.1, H is dense in G if and only if it is dense in ΩnGp. On the other hand, by Proposition 14.2, if H = w1, . . . , wn , then H is dense in ΩnGp if and only if the associated implicit operator (w1, . . . , wn) is Gp-invertible. By Proposition 14.3, the latter condition is equivalent to the integer matrix M(w1, . . . , wn) being invertible mod p. 173 / 332 Finally, the next result improves the general-purpose algorithm in the proof of Theorem 15.3 by providing the means to compute efficiently a proper clopen subgroup containing a given non-dense finitely generated subgroup. Proposition 15.7 Let H be a finitely generated subgroup of the free group G on n free generators x1, . . . , xn and let {w1, . . . , wm} be generating subset of H. Let A = (Z/pZ)n and consider the homomorphism ϕ : G → A which sends xi to (0, . . . , 0, 1, 0, . . . , 0), where 1 is in the ith position. 1. The subgroup H is dense in the pro-Gp metric of G if and only if, mod p, the matrix M(w1, . . . , wm) has rank n. 2. If, mod p, the submatrix M(wr1 , . . . , wrn ) has rank n, then the subgroup wr1 , . . . , wrn is dense. 3. If, mod p, the matrix M(w1, . . . , wm) has rank less than n, then ϕ−1ϕ(H) is a proper clopen subgroup of G containing H. 174 / 332 Proof. Since Z/pZ is a field, a square matrix with entries in it is invertible if and only if the matrix has rank n. Moreover, an m × n matrix has rank n if and only if it contains n rows whose corresponding submatrix has rank n. On the other hand, since ϕ(H) is generated by the rows of the matrix M(w1, . . . , wm), with the entries viewed in Z/pZ, if, mod p, this matrix has rank less than n, then ϕ(H) is a proper subgroup of the group A ∈ Gp. This proves (3) and shows that H is not dense in G. The statements (1) and (2) now follow using Proposition 15.6. Thus, for the pro-Gp metric of a finitely generated free group G, combining part (3) of Proposition 15.7 and Proposition 15.1, one may compute a proper clopen subgroup containing a given finitely generated subgroup H which is not dense in G. 175 / 332 To illustrate the above algorithms, we present an example in detail. For this purpose, we consider again the subgroup H of the free group G on the free generators a, b, c of Example 1: H is generated by the group words ab−1c−1a, a−1b−1ac−1a, bc−1a. The corresponding integer matrix is M0 =   2 −1 −1 1 −1 −1 1 1 −1   and has determinant 2. Hence H is dense in G with respect to the pro-Gp metric for every odd prime p. For the pro-G2 metric of G, the rank of the matrix M0 mod 2 is 2. More precisely, if ϕ0 : G → (Z/2Z)3 is the appropriate mapping given by Proposition 15.7, then H ⊆ ϕ−1 0 ϕ0(H) G. The clopen subgroup ϕ−1 0 ϕ0(H) may be computed as the fundamental group of the automaton which represents the action of a, b, c on the right of the right cosets of ϕ0(H), which is represented in Figure 3. ϕ0(H) ϕ0(H)(0, 1, 0) b, c b, c a a Figure: The automaton of K1 The clopen subgroup K1 in question is therefore freely generated by {a, bab−1, b2, bc, cb−1}. The free factor L1 is obtained by taking the fundamental group of the image of A(H) in A(K1). Recall the automaton A(H) as given in Figure 177. 176 / 332 1 2 3 45 ab a b a c b Figure: The automaton A(H) The states 1, 2, 4 are mapped to ϕ(H) while 3, 5 are mapped to ϕ(H)(0, 1, 0) and the only missing edge from the automaton of A(K1) is the edge c from the state ϕ(H)(0, 1, 0) to ϕ(H), which corresponds to the generator bc of K1. Hence L1 is the subgroup generated by {a, bab−1, b2, cb−1}. Let b1 = bab−1, b2 = b2 and c1 = cb−1. Then the elements of H may be expressed as follows: ab−1 c−1 a = ab−1 2 c−1 1 a a−1 b−1 ac−1 a = a−1 b−1 2 b1 bc−1 a = c−1 1 a which are now viewed as generating a subgroup H1 of the free group G1 on the set a, b1, b2, c1. The matrix corresponding to the generators of H1 is M1 =   2 0 −1 −1 −1 1 −1 0 1 0 0 −1   which, mod 2, has rank 3. Hence H is not dense in L1 and we need to compute a clopen subgroup K2 of L1 and a free factor L2 of K2 containing H. 177 / 332 Since the rank of M1 mod 2 is 3, the image of H1 under the homomorphism ϕ1 : G1 → (Z/2Z)4 has index 2 and again we obtain a 2-state permutation automaton for K2, which is represented in Figure 5. ϕ1(H1) ϕ1(H1a) a, b2, c1 a, b2, c1 b1 b1 Figure: The automaton of K2 To obtain the image of A(H1) in this automaton, it suffices to read through each of the generators of H1, starting from the initial state, and keeping only the states and edges which are used. Figure 6 represents the subautomaton which is thus obtained, whose fundamental group is the required free factor L2 of K2. 178 / 332 ϕ1(H1) ϕ1(H1a) a, b2 a, c1 b1 Figure: The automaton of L2 Choosing the edge labeled b2 for the generating tree, we obtain the following free generators for L2: b1, a1 = ab−1 2 , a2 = b2a, c2 = b2c1. (3) In terms of these generators, the expressions for the generators of H1 are the following: ab−1 2 c−1 1 a = a1c−1 2 a2 a−1 b−1 2 b1 = a−1 2 b1 c−1 1 a = c−1 2 a2 The matrix of these generators with respect to the generators of L2 in the order given by (3) is M2 =   0 1 1 −1 1 0 −1 0 0 0 1 −1   which again has rank 3. 179 / 332 Proceeding as above, we get a canonical homomorphism ϕ2 : G2 → (Z/2Z)4, where G2 is the free group on the generators b1, a1, a2, c2. Let H2 be the subgroup of G2 generated by a1c−1 2 a2, a−1 2 b1, c−1 2 a2. The new clopen subgroup K3 is the fundamental group of the automaton in Figure 7. ϕ2(H2) ϕ2(H2b1) b1, a2, c2 b1, a2, c2 a1 a1 Figure: The automaton of K3 Figure 8 gives the subautomaton which is obtained by reading through the preceding automaton the generators of H2. ϕ2(H2) ϕ2(H2b1) b1, a2, c2 a1 Figure: The automaton of L3 Choosing the edge labeled c2 for the generating tree, we obtain the following free generators for L3: a1, a3 = c−1 2 a2, b3 = c−1 2 b1. (4) 180 / 332 ϕ2(H2) ϕ2(H2b1) b1, a2, c2 a1 a1, a3 = c−1 2 a2, b3 = c−1 2 b1. The corresponding expressions for the generators of H2 are: a1c−1 2 a2 = a1a3 a−1 2 b1 = a−1 3 b3 c−1 2 a2 = a3 whose matrix has rank 3 mod 2, 3 being also the rank of L3. Hence H2 is dense in L3 and the closure of H = ab−1c−1a, a−1b−1ac−1a, bc−1a in G is obtained by substituting back in terms of a, b, c the expressions for the generators of L3 given in (4): a1 = a1 = ab−1 2 = ab−2 a3 = c−1 2 a2 = c−1 1 b−1 2 b2a = bc−1 a b3 = c−1 2 b1 = c−1 1 b−1 2 b1 = bc−1 b−2 bab−1 = bc−1 b−1 ab−1 Since the generators of L3 are all recognized by A(H), we conclude that H is closed in the pro-G2 metric of G. 181 / 332 In terms of generators, we obtained the following chain of subgroups approximating the closure of H, where we also indicate on the right the corresponding congruences on A(H) as given by the partitions of the state set: G = a, b, c ∪ K1 = a, bab −1 , b 2 , bc, cb −1 ∪ L1 = a, bab −1 , b 2 , cb −1 {1, 2, 4|3, 5} ∪ K2 bab −1 , ab −2 , b 2 a, b 2 cb −1 , cb −3 , b 3 ab −3 , b 4 ∪ L2 bab −1 , ab −2 , b 2 a, b 2 cb −1 {1|2, 4|3, 5} ∪ K3 ab −2 , bc −1 a, bc −1 b −1 ab −1 , b 2 ab 2 cb −1 , babcb −1 , b 2 cbcb −1 , bc −1 b −2 acb −1 ∪ L3 ab −2 , bc −1 a, bc −1 b −1 ab −1 {1|2|3|4|5} H = ab −1 c −1 a, a −1 b −1 ac −1 a, bc −1 a Exercise 15.8 Verify that the above calculations are correct. 182 / 332 The above calculations also show that H is not Gp-extendible for every odd prime p while it is G2-extendible. An G2-extension of the automaton A(H) may be recovered from our calculations. It is the permutation automaton associated with the following clopen subgroup: K = ab−2 , bc−1 a, bc−1 b−1 ab−1 , b2 ab2 cb−1 , babcb−1 , b2 cbcb−1 , bc−1 b−2 acb−1 , cb−3 , b3 ab−3 , b4 , b2 a · cb−3 · (b2 a)−1 , b2 a · b3 ab−3 · (b2 a)−1 , b2 a · b4 · (b2 a)−1 , bc, a · bc · a−1 , b2 a · bc · (b2 a)−1 , b2 a · abca−1 · (b2 a)−1 . Exercise 15.9 Explain in general how to compute a V-extension of a V-extendible finite inverse reduced automaton, assuming the appropriate hypotheses on the extension-closed pseudovariety of groups V. 183 / 332 One may calculate the following picture for A(K) and that the transition monoid of this automaton is a group of order 128. Since it would be too tedious to do it by hand, this calculation was carried out using some adequate Mathematica routines for the symbolic computations and the package graphviz for the drawing itself, which was then converted by hand to GasTeX for readability. The numbering of states was chosen to facilitate the identification of the subautomaton A(H). New edges and states are in red. 6 7 5 8 4 2 3 1 a b c b c b a c b c a c a b c b c c b b a a a a Figure: A G2-extension of A(H) 184 / 332 Exercise 15.10 Show that the above algorithms for the pseudovariety Gp (to test denseness, to compute the closure, to exhibit a Gp-extension, if one exists) require only polynomial time in the input, which may be either a finite list of group words over a finite alphabet A, or the reduced inverse automaton of the subgroup of the free group on A that they generate. More precisely, estimate the time complexity of the algorithms. 185 / 332 Outline σ-fullness σ-reducibility of the V-separation problem pro-V-metrics Inverse automata V-invertibility of endomorphisms Computing closures of finitely generated subgroups The Pin-Reutenauer procedure 186 / 332 The Pin-Reutenauer procedure Let us consider again the problem studied by Pin and Reutenauer [PR91]: given a regular language L ⊆ A+ and a word w ∈ A+, decide whether w can be separated from L by a G-recognizable language. Since G is κ-reducible for the separation problem, the problem is equivalent to deciding whether w ∈ clκ,G(L). So, the idea is to compute the subset clκ,G(L) of the free group Ωκ AG. Since L is regular, we may assume that it is given by some regular expression. As every finite subset of Ωκ AG fullfullis closed and closure behaves well with respect to union, the key question is how the closure operator clκ,G( ) behaves well with respect to the multiplication of subsets and the Kleene star operation. 187 / 332 In general, for an arbitrary pseudovariety V and implicit signature σ, we have the inclusion clσ,V(K L) ⊇ clσ,V(K) clσ,V(L), (5) simply because the multiplication in Ωκ AV is continuous. Note: in a compact semigroup, the equality K L = K L holds. For a subset X of a σ-algebra S, denote by S σ the σ-subalgebra generated by X. The analogue of (5) for Kleene star is the inclusion clσ,V(L+ ) ⊇ clσ,V(L) σ (6) The proof of (6) is a bit more delicate and depends on the following lemma. 188 / 332 Lemma 16.1 Let V be a pseudovariety of semigroups and let S be a pro-V semigroup. Then the following evaluation mapping is continuous for every positive integer n: ΩnV × Sn → S (w, s1, . . . , sn) → wS (s1, . . . , sn). Proof. Since V-implicit operations commute with continuous homomorphisms between pro-V semigroups, in view of Proposition 5.1 it suffices to consider the case where S ∈ V. Note that ΩnV(S) is finite, say again by Proposition 5.1. Moreover, when ΩnV and S are both finite, the continuity of the evaluation mapping is trivial. Since, in terms of implicit operations, the natural projection ΩnV → ΩnV(S) is given by restriction, we deduce the general case. 189 / 332 The inclusion clσ,V(L) σ ⊆ clσ,V(L+ ) follows from the next proposition. Proposition 16.2 Let V be a pseudovariety, σ and implicit signature, and L a subset of Ωσ AV. Then clσ,V(L+ ) is a σ-subalgebra of Ωσ AV. Proof. Let u ∈ ΩmS be an implicit operation from σ, and let v1, . . . , vm be elements of clσ,V(L+ ). There are sequences with the following properties: a sequence (un)n, of words from {x1, . . . , xm}+ , converging to u; for each i ∈ {1, . . . , m}, a sequence (vi,n)n, of elements of L+ , converging to vi . For each n, the element (un)ΩAV(v1,n, . . . , vm,n) is a product of elements from L+ and, therefore, also belongs to L+ . By Lemma 16.1, we have lim(un)ΩAV(v1,n, . . . , vm,n) = uΩAV(v1, . . . , vm). Hence uΩAV(v1, . . . , vm) belongs to clσ,V(L+ ). 190 / 332 Theorem 16.3 (Pin and Reutenauer [PR91]) For the pseudovariety V = G, the signature σ = κ, and a regular language L ⊆ A+, equality holds in both formulas (5) and (6). Some comments on the proof: The proof establishes a stronger result than (6), namely that clκ,G(L+ ) = L κ. The main reason for not needing to use clκ,G(L) instead of L on the right hand side is M. Hall’s Theorem 13.5. The proof depends on several ingredients from language theory, such as a theorem of Anissimov and Seifert, stating that the rational7 subgroups of Ωκ AG are the finitely generated subgroups, and a theorem of Fliess, stating that the rational subsets of Ωκ AG form a Boolean algebra. (See [Ber79, Section III.2] — http://www-igm.univ-mlv.fr/∼berstel/LivreTransductions/LivreTransductions.html.) 7 A subset of a monoid M is said to be rational if it can be obtained from the empty set and the singleton subsets of M by using the operations of subset multiplication and taking the submonoid X∗ generated by a subset X. 191 / 332 The theorem was initially deduced from a conjectured property of free groups: that the product of finitely many finitely generated subgroups of Ωκ AG is closed in the pro-G metric. This property follows from Ash’s Theorem on inevitable graphs [Ash91] and was proved independently by Ribes and Zalesski˘ı [RZ93], using profinite group theory. The theorem was found as an approach to a conjecture of J. Rhodes (known as the Rhodes Type II Conjecture), which it implies. Ash’s Theorem on inevitable graphs was also proved to establish Rhodes’ conjecture. (See [HMPR91] for the original statement and the history of this conjecture.) The Rhodes Type II Conjecture, now theorem, may be stated as follows [AS00a]. Theorem 16.4 (Rhodes Type II) Let S be a finite semigroup and let A be a generating subset. Consider the κ-subsemigroup T of S × Ωκ AG generated by the pairs (a, a) (a ∈ A). Then, for an element s ∈ S, (s, 1) ∈ T if and only if, for every finite group G, and every subsemigroup U of S × G, which projects onto the first component, (s, 1) ∈ U. 192 / 332 Other pseudovarieties for which the Pin-Reutenauer procedure holds We say that the Pin-Reutenauer procedure holds for a pseudovariety V and an implicit signature σ if the following equalities hold for all subsets L of Ωσ AV: clσ,V(K L) = clσ,V(K) clσ,V(L) (7) clσ,V(L+ ) = clσ,V(L) σ. (8) Note that if the pseudovariety V contains N, then there is always an implicit signature for which the Pin-Reutenauer procedure holds (cf. Proposition 7.3), for example the trivial signature { · }. But, this in general too small for the pseudovariety to be full with respect to it. 193 / 332 In a sense at the other extreme, if we take σ to consist of all (finitary) implicit operations, then Ωσ AV = ΩAV and clσ,V(L) = L is the closure of L in ΩAV. So, the equality (7) certainly holds in this case, by compactness of ΩAV. On the other hand, clσ,V(L+) is the closure of a subsemigroup, whence a closed subsemigroup of ΩAV, namely the closed subsemigroup generated by L, which is therefore contained in clσ,V(L) σ. The reverse inclusion is given by Proposition 16.2. Hence, for every pseudovariety V, there is some implicit signature for which the Pin-Reutenauer procedure holds and V is σ-full. The question is whether it holds for small such signatures, such as κ. The other key question is under what conditions one can turn the Pin-Reutenauer procedure into an algorithm to test membership of a given word w ∈ L in the closure clσ,V(L) of a given regular language L ⊆ A+. 194 / 332 Here are some recent results. Theorem 16.5 (JA-JCCosta-MZeitoun) Suppose that V and W are σ-full pseudovarieties of semigroups such that V ⊆ W. If the Pin-Reutenauer procedure holds for W then it also holds for V. Theorem 16.6 (JA-JCCosta-MZeitoun) The Pin-Reutenauer procedure holds for A in the signature κ. As was already mentioned, the same authors proved that A and R are κ-full. Hence the preceding two theorems yield the following result. Corollary 16.7 The Pin-Reutenauer procedure holds for R in the signature κ. 195 / 332 Part III Relatively free profinite semigroups and Symbolic Dynamics 196 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 197 / 332 Semigroups of implicit operators Let wi ∈ ΩnS be an n-ary implicit operation for each i ∈ {1, . . . , n}. Given a profinite semigroup S, the n-tuple induces a continuous mapping [w1, . . . , wn]S : Sn → Sn (s1, . . . , sn) → (w1)S (s1, . . . , sn), . . . , (wn)S (s1, . . . , sn) . The composition of two such transformations of Sn is again a transformation of the same form, and thus they constitute a semigroup, which we call the semigroup of n-ary implicit operators of S and denote On(S). In particular, we may iterate an implicit operator f = [w1, . . . , wn]S by considering its successive powers f , f 2, f 3, . . .. 198 / 332 The following result is immediate from the commutation of implicit operations with comtinuous homomorphisms between profinite semigroups. Lemma 17.1 Let w1, . . . , wn ∈ ΩnS be n-ary implicit operations and let ϕ : S → T be a continuous homomorphism between profinite semigroups. Then the following diagram commutes: Sn [w1,...,wn]S GG ϕn  Sn ϕn  Tn [w1,...,wn]T GG Tn. 199 / 332 Lemma 17.2 Let S be a profinite semigroup and let f ∈ On(S) be an n-ary implicit operator on S. Then, for every (s1, . . . , sn) ∈ Sn, the sequence f k!(s1, . . . , sn) k of elements of Sn converges. In other words, the sequence (f k!)k of transformations of Sn converges pointwise. Moreover, the limit is an idempotent transformation of Sn. 200 / 332 Proof. By definition of profinite semigroup, S admits a complete metric d. We may endow the product space Sn with various metrics, among which the maximum metric d defined by d (s1, . . . , sn), (t1, . . . , tn) = max{d(si , ti ) : i = 1, . . . , n}. This defines a complete metric structure on Sn. Thus, to prove that the sequence f k!(s1, . . . , sn) k converges, it suffices to show that it is a Cauchy sequence. By Proposition 5.1 and Lemma 17.1, it suffices to consider the case where S is a finite semigroup, in which case the result amounts to the convergence of the corresponding sequence f k! in the finite monoid of transformations of Sn, where indeed we have observed that it converges to an idempotent. 201 / 332 We denote the limit lim[w1, . . . , wn]k! S by [w1, . . . , wn]ω S . In particular, given w1, . . . , wn ∈ ΩnS, we may consider the following n-tuple of n-ary implicit operations: (v1, . . . , vn) = [w1, . . . , wn]ω ΩnS (x1, . . . , xn). Lemma 17.3 With the above notation, if S is a profinite semigroup, then [v1, . . . , vn]S = [w1, . . . , wn]ω S . Thus, the transformation [w1, . . . , wn]ω S is again an implicit operator and the semigroup On(S) is closed under taking ω-powers. 202 / 332 Proof. Let (s1, . . . , sn) be an n-tuple of elements of S. Then there is a unique continuous homomorphism ϕ : ΩnS → S such that ϕ(xi ) = si (i = 1, . . . , n). In view of Lemma 17.1 and taking into account that ϕn is continuous, we obtain [v1, . . . , vn]S (s1, . . . , sn) = ϕn [v1, . . . , vn]ΩnS(x1, . . . , xn) = ϕn lim[w1, . . . , wn]k! ΩnS (x1, . . . , xn) = lim ϕn [w1, . . . , wn]k! ΩnS (x1, . . . , xn) = lim[w1, . . . , wn]k! S (s1, . . . , sn) = [w1, . . . , wn]ω S (s1, . . . , sn), which establishes the desired equality. 203 / 332 Since ΩnS is a profinite semigroup freely generated by the set {x1, . . . , xn}, the n-tuples of n-ary implicit operations (w1, . . . , wn) are in bijection with the continuous endomorphisms ϕ(w1,...,wn) of ΩnS, where ϕ(w1,...,wn)(xi ) = wi (i = 1, . . . , n). On the other hand, they are also in bijection with n-ary implicit operators on ΩnS, namely through the formula (w1, . . . , wn) = [w1, . . . , wn]ΩAS(x1, . . . , xn). For a profinite semigroup S, denote by End S the monoid of continuous endomorphisms of S, acting on the left. 204 / 332 Proposition 17.4 The following mapping is an anti-isomorphism of semigroups: Θ : End ΩnS → On(ΩnS) ϕ → [ϕ(x1), . . . , ϕ(xn)]ΩnS. Proof. We have already observed that the above mapping is a bijection, and described its inverse. So, it remains to verify that it is an anti-homomorphism of semigroups. Let ϕ, ψ ∈ End ΩnS and let wi = ψ(xi ) (i = 1, . . . , n). Taking into account Lemma 17.1, we obtain Θ(ϕ ◦ ψ) = [ϕ(ψ(x1)), . . . , ϕ(ψ(xn))]ΩnS = [ϕ(w1), . . . , ϕ(wn))]ΩnS = [w1(ϕ(x1), . . . ϕ(xn)), . . . , wn(ϕ(x1), . . . ϕ(xn))]ΩnS = [w1, . . . , wn]ΩnS ◦ [ϕ(x1), . . . ϕ(xn)]ΩnS = Θ(ψ) ◦ Θ(ϕ). 205 / 332 Lemma 17.5 Let ϕk, ϕ ∈ End ΩnS (k ≥ 0). Then (ϕk)k convergences pointwise to ϕ if and only if (Θ(ϕk))k converges pointwise to Θ(ϕ). Proof. Let vk,i = ϕk(xi ) and vi = ϕ(xi ) (i = 1, . . . , n; k ≥ 0). (⇒) For w1, . . . , wn ∈ ΩnS, we have Θ(ϕk)(w1, . . . , wn) = [vk,1, . . . , vk,n]ΩnS(w1, . . . , wn) = (vk,1)ΩnS(w1, . . . , wn), . . . , (vk,n)ΩnS(w1, . . . , wn) → (v1)ΩnS(w1, . . . , wn), . . . , (vn)ΩnS(w1, . . . , wn) = [v1, . . . , vn]ΩnS(w1, . . . , wn) = Θ(ϕ)(w1, . . . , wn), where the convergence step follows from Lemma 16.1. 206 / 332 (. . . ) (⇐) Let w ∈ ΩnS. Then we have ϕk(w) = ϕk wΩnS(x1, . . . , xn) = wΩnS vk,1, . . . , vk,n = wΩnS [vk,1, . . . , vk,n]ΩnS(x1, . . . , xn) → wΩnS [v1, . . . , vn]ΩnS(x1, . . . , xn) = ϕ wΩnS(x1, . . . , xn) = ϕ(w), where we use again Lemma 16.1 for the convergence step. 207 / 332 Corollary 17.6 For a continuous endomorphism ϕ of ΩnS, the sequence (ϕk!) converges pointwise to an idempotent continuous endomorphism of ΩnS. The idempotent in question is denoted ϕω and is called the ω-power or ω-iterate of ϕ. More generally, Hunter [Hun83] has shown that the monoid End S of continuous endomorphisms of a finitely generated profinite semigroup S is profinite. Since it is usually not finitely generated, this result does not fit in the realm of our metric approach to profinite semigroups. For those that know about topology of function spaces, the topology used in End S is the topology of pointwise convergence (i.e., as a subspace of the product space SS ), and turns out to coincinde with the compact-open topology (i.e., the topology of uniform convergence). See also [Alm02, AV06a, Alm05b, Ste11]. 208 / 332 Examples The ω-iteration of continuous endomorphisms of ΩnS, or of implicit operators on ΩnS can be used to construct useful implicit operations. Let ϕ ∈ End Ω1S be defined by ϕ(x1) = xp 1 . Then ϕk(x1) = xpk 1 and ϕω(x1) = xpω 1 . Recall that we used the implicit operation xpω 1 previously to define the pseudovariety Gp in terms of pseudoidentities: Gp = [[xpω 1 = 1]]. 209 / 332 Given elements u and v of a profinite semigroup S, let [u, v] = uω−1 vω−1 uv, an extension of the group commutator. Note, however, that, if S is not a group, [u, v] idempotent may not be equivalent to uv = vu. Let ϕ ∈ End Ω2S be defined by ϕ(x1) = [x1, x2] and ϕ(x2) = x2. Note that ϕk(x1) is the iterated commutator defined recursively by [x1,1 x2] = [x1, x2] together with [x1,k+1 x2] = [[x1,k x2], x2]. We let [x1,ω x2] = ϕω(x1), which is an implicit operation in Ω2S. Using a theorem of Zorn [Zor36], that states that a finite group is nilpotent (i.e., a direct product of groups of prime power order) if and only if it satisfies some identity of the form [x,n y] = 1, it is now easy to solve Exercise 6.5(2): Gnil = [[[x,ω y] = 1]]. 210 / 332 The Prouhet-Thue-Morse substitution When ϕ ∈ ΩAS extends an endomorphism of A+ it is also called a (finite) substitution since applying ϕ to a word corresponds to substituting each letter a ∈ A by the word ϕ(a). The substitution over the alphabet {a, b} defined by τ(a) = ab and τ(b) = ba was first considered by Prouhet [Pro51], and rediscovered by Thue [Thu06, Thu12] and Hedlund and Morse [HM38]. Here is an example of a result involving it: Theorem 17.7 (ˇSirˇsov [ˇS63]) The pseudovariety [[τω(a) = τω(b), xω = 1]] consists of all finite groups admitting a normal nilpotent subgroup whose corresponding factor group is a 2-group. 211 / 332 The substitution τ determines a binary implicit operator [ab, ba]S on every profinite semigroup S. How does it behave on a finite group? When iterating a transformation on a finite set, in a finite number steps one must enter into a periodic orbit. One possible way to visualize the action of our Prouhet-Thue-Morse implicit operator on a finite group G is to draw a square in which the sides stand for G, and a pixel, an elementary square, for an element of G × G. One may color each pixel so as to encode for instance how many applications of the operator it takes to reach a periodic orbit, and to distinguish between the different periodic orbits. It is somewhat striking that one obtains pictures like those in the following two pages. The ordering of the elements in the groups is that given by GAP [GAP06], a computer algebra system that was used to carry out the calculations. The groups are Z/858Z and V4 A3. 212 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 213 / 332 Finite factor complexity Given w ∈ ΩAS, denote by F(w) the set of words u ∈ A+ which are factors of w, and let Fn(w) = F(w) ∩ An. The complexity sequence qw (n) is defined by letting qw (n) = |Fn(w)|, where n is a positive integer. Lemma 18.1 Let w ∈ ΩAS \ A+ and let u ∈ F(w). If there is no a ∈ A such that ua ∈ F(w), then there is factorization of the form w = vu. 214 / 332 Proof. Let (wn)n be a sequence of finite words converging to w. For each z ∈ A+, the set Cz = (ΩAS)1z(ΩAS)1 = A∗zA∗ is a clopen set such that Cz ∩ A+ = A∗zA∗. By hypothesis, w belongs to Cu and so we may assume that so do all wn, since Cu is open. On the other hand, since Cua is closed, we may assume that no wn ∈ Cua, for any a ∈ A. This leaves only one place for u to be a factor of each wn, namely as a suffix: wn = vnu for some vn ∈ A∗. Since ΩAS is compact, some subsequence of (vn)n converges to some v ∈ ΩAS. By continuity of multiplication, it follows that w = vu. 215 / 332 Lemma 18.2 Let w ∈ ΩAS. Then the following formula holds for every positive integer n: qw (n + 1) − qw (n) = u∈Fn(w) (|{a ∈ A : ua ∈ Fn+1(w)}| − 1). (9) Proof. Each factor of w of length n + 1 is of the form ua, for some a ∈ A, which yields the following formula: qw (n + 1) = u∈Fn(w) |{a ∈ A : ua ∈ Fn+1(w)}|. Subtracting qw (n) = u∈Fn(w) 1, we obtain formula (9). 216 / 332 Because of the suffix of length n appearing nowhere else in w (cf. Lemma 18.1), there may be one negative term in the sum (9). However, the sum is always nonnegative for w ∈ ΩAS \ A+. Lemma 18.3 Let w ∈ ΩAS \ A+. Then the sequence qw (n) is increasing. 217 / 332 Proof. Suppose that qw (n + 1) < qw (n). Then, as argued above, the suffix u of length n of w must not appear elsewhere in w as a factor, and all other terms in the sum must be zero: for every v ∈ Fn(w) \ {u}, there is exactly one letter av ∈ A such that vav ∈ Fn+1(w). Now, suppose that tz is a finite suffix of w, with |t| = n. Then there is no word z with |z | > |z| such that tz is a suffix of w. For, otherwise, we may take a counterexample (t, z) with |z| minimum. Then z = 1 for, else, we must have t = u and we know that u does not appear elsewhere in w. Hence, z = atz1 and, if we write t = bt1, with b ∈ A, then (t1, z1) is still a counterexample and |z1| < |z|, which contradicts our choice of counterexample. Hence every factor of length n of w which appears as part of a finite suffix, can be found as such in only one position. Since Fn(w) ⊆ An is finite, we conclude that w cannot have arbitrarily long finite suffixes, whence w ∈ A+ , contrary to the hypothesis. 218 / 332 We say that a finite word u is primitive if it is not of the form vn for some integer n > 1. Theorem 18.4 ([AV06a]) Suppose that w ∈ ΩAS \ A+ is such that qw (n) ≤ n for some n. Then there exist finite words x, y, and z and an infinite ν ∈ ˆZ+ such that w = xyνz, |xy| ≤ n, and |yz| ≤ n. Corollary 18.5 ([AV06a]) Let w be any pseudoword which is not of the form xyνz for some finite words x, y, z and some ν ∈ ˆZ+. Then qw (n) ≥ n + 1 for every n. 219 / 332 The next result identifies maximal subgroups of ΩAS of small complexity. Theorem 18.6 ([AV06a]) Suppose that w lies in a subgroup of ΩAS and qw (n) ≤ n for some n. Then the H-class of w is a free procyclic group with generator of the form yω+1 for some word y of length at most n. In particular, for each positive n, there are only finitely many H-classes of such pseudowords w. Note that J -equivalent elements of ΩAS have the same complexity. Corollary 18.7 The complexity of any non-procyclic subgroup of ΩAS is at least q(n) ≥ n + 1. 220 / 332 Sturmian pseudowords Thus, the minimum possible complexity for a non-procyclic subgroup of ΩAS is q(n) = n + 1. Then q(1) = 2, which means that there are exactly two letters involved in the subgroup in question. We say that w ∈ ΩAS is Sturmian if qw (n) = n + 1 for every n ≥ 1. 221 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 222 / 332 Primitive substitutions Since substitutions over a one-letter alphabet are not very interesting, we assume from hereon that |A| ≥ 2. A finite substitution ϕ ∈ End ΩAS is said to be primitive if there is some positive integer N such that, for every two letters a, b ∈ A, a appears in the word ϕN(b). Proposition 19.1 Let ϕ ∈ End ΩAS be a primitive finite substitution. Then all pseudowords of the form ϕω(a), with a ∈ A, lie in the same J -class. Proof. Let N be as above and let a, b ∈ A. Then a is a factor of ϕkN(b) for every k. By compactness, it follows that a is also a factor of ϕω(b) = (ϕN)ω(b), say ϕω(b) = uav. Since ϕω is idempotent, we deduce that ϕω(b) = ϕω(u) · ϕω(a) · ϕω(v), which shows that ϕω(a) is a factor of ϕω(b). 223 / 332 We have already seen one classical example of primitive finite substitution, namely the Prouhet-Thue-Morse substitution a → ab, b → ba. Another classical example is that of the so-called Fibonacci substitution ϕ(a) = ab, ϕ(b) = a. Note that the fn = |ϕn(b)| satisfies the recurrence relation fn+2 = fn+1 + fn, with f1 = f2 = 1. ϕω(b) is an example of Sturmian pseudoword. More generally, the following is a consequence of a theorem of Mignosi and S´e´ebold [MS93] on infinite words. Theorem 19.2 Let ϕ ∈ End Ω2S be a primitive substitution. Then ϕω(x1) is Sturmian if and only if ϕ induces an automorphism of the free group Ωκ 2G. 224 / 332 By Proposition 19.1, one may consider the complexity of a primitive finite substitution ϕ ∈ End ΩAS. Theorem 19.3 ([ELR75, Pan84]) The complexity of a primitive finite substitution is at most linear. 225 / 332 Proof. Let ϕ ∈ End ΩAS be a primitive finite substitution, let q be its complexity, and let n be a positive integer. There is some p such that min a∈A |ϕp−1 (a)| ≤ n ≤ min a∈A |ϕp (a)|. (10) Then, every word in Fn(ϕω (a)) must be a factor of some ϕp (a) or of some ϕp (ab), with ab a factor of ϕω (ab). For each such two-letter word ab, there are at most |ϕp (ab)| factors of length n. Hence, if we let r = |A|2 , then we have q(n) ≤ 2r max a∈A |ϕp (a)|. (11) By the Perron-Frobenius Theorem (19.4), there are constants α, c, d > 0 such that cαp ≤ min a∈A |ϕp (a)| ≤ max a∈A |ϕp (a)| ≤ dαp . (12) Combining the inequalities (11) and (12) and taking also into account (10), we obtain q(n) ≤ 2r dαp ≤ 2r d c α n. 226 / 332 The Perron-Frobenius Theorem Let ϕ ∈ End ΩAS be a finite substitution. For each a, b ∈ A, one may consider the number |ϕ(a)|b of occurrences of the letter b in the word ϕ(a). This defines a matrix M(ϕ) with nonnegative integer entries. Note that: the sum of the entries in the row corresponding to a ∈ A is |ϕ(a)|; for another finite substitution ψ ∈ End ΩAS, M(ϕ ◦ ψ) = M(ϕ)M(ψ); thus M(ϕk ) = (M(ϕ))k for every k ≥ 1; hence ϕ is primitive if and only if tere exists k ≥ 1 such that (M(ϕ))k has all entries positive; a matrix with nonnegative real entries satisfying this property is said to be primitive; An n × n real matrix M is said to be irreducible if, for every pair of indices i, j, there is a power Mk whose i, j-entry is nonzero. 227 / 332 Theorem 19.4 (Perron-Frobenius) Let M be a nonnegative irreducible matrix. Then there is a positive simple eigenvalue α such that α ≥ |λ| for every other eigenvalue λ and α admits an eigenvector whose coordinates are all positive. In case M is primitive, α > |λ| for every other eigenvalue λ. The special, dominant eigenvalue α, is called the Perron-Frobenius eigenvalue of the matrix M. See http://en.wikipedia.org/wiki/Perron-Frobenius theorem for the significance of the theorem and several proofs. 228 / 332 For the proof of the existence of constants c, d such that the inequalities (12) hold, let M be a r × r nonnegative primitive matrix and let α be its Perron-Frobenius eigenvalue. Let v = (v1, . . . , vr ) be a corresponding positive eigenvector: Mv = αv, whence Mp v = αp v for all p ≥ 1. Let (Mp )i,j denote the i, j entry of the matrix Mp . For i = 1, . . . , r, we have r j=1 (Mp )i,j vj = αp vi . Hence, if c0 = mini vi and d0 = maxi vi , then we get c0 r j=1(Mp )i,j ≤ d0αp for every i, whence max i r j=1 (Mp )i,j ≤ d0 c0 αp . On the other hand, we have c0αp ≤ d0 r j=1(Mp )i,j for every i, which gives c0 d0 αp ≤ min i r j=1 (Mp )i,j . 229 / 332 Examples Consider the Prouhet-Thue-Morse substitution τ(a) = ab, τ(b) = ba. Then M(τ) = 1 1 1 1 has characteristic polynomial (1 − x)2 − 1 = x(x − 2) and eigenvalues 0 < 2. So, 2 is the Perron-Frobenius eigenvalue. The corresponding eigenspace is generated by (1, 1). For w = τω(a), the proof of Theorem 19.3 gives qw (n) ≤ 2 × 22 × 1 1 × 2n = 16n. It turns out that this is about double the exact value, which has been computed (see [Fog02, Proposition 5.1.9]). 230 / 332 Let ϕ be the Fibonacci substitution: ϕ(a) = ab, ϕ(b) = a. Then M(ϕ) = 1 1 1 0 has characteristic polynomial (1 − x)(−x) − 1 = x2 − x − 1 and eigenvalues 1− √ 5 2 < 1+ √ 5 2 . The Perron-Frobenius eigenvalue, 1+ √ 5 2 , has eigenspace generated by the vector (1+ √ 5 2 , 1). For w = ϕω(a), the proof of Theorem 19.3 gives qw (n) ≤ 2×22 × 1 + √ 5 2 2 × 1 + √ 5 2 n = 2(1+ √ 5)3 n 68n, while we know that qw (n) = n + 1. 231 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 232 / 332 Entropy The complexity (function) of a pseudoword has the following log-subadditive property: Lemma 20.1 Let w ∈ ΩAS. Then the inequality qw (r + s) ≤ qw (r) qw (s) holds for all positive integers r and s. Proof. A factor of w of length r + s is of the form uv, where u ∈ Fr (w) and v ∈ Fs(w). Lemma 20.2 (Fekete’s Lemma) Let (tn)n be a sequence of real numbers such that tr+s ≤ tr + ts for all r, s ≥ 1. Then we have lim tn n = inf tn n . 233 / 332 Proof. Let = inf tn n and let be a positive real number. Let K be an index such that tK K < + 2. There is also a positive integer M such that tr KM < 2 for r < K. Let n ≥ KM and write n = qK + r with r < K. Then ≤ tn n ≤ qtK qK + r + tr n ≤ qtK qK + tr KM ≤ tK K + tr KM ≤ + 2 + 2 which establishes the lemma. 234 / 332 Let A be a finite alphabet and m = |A|. For a pseudoword w ∈ ΩAS \ A+, by Lemmas 20.1 and 20.2, the following limit exists: h(w) = lim n→∞ 1 n logm qw (n). It is called the entropy of w. Since 1 ≤ qw (n) ≤ mn, the entropy h(w) is a real number in the interval [0, 1]. If u, v ∈ ΩAS \ A+ are such that u is a factor of v, then h(u) ≤ h(v). 235 / 332 Iteration of primitive finite substitutions leads to minimum entropy: Theorem 20.3 Let ϕ ∈ End ΩAS be a primitive finite substitution. Then we have h(ϕω(a)) = 0. Proof. Let w = ϕω(a). By Theorem 19.3, there is some constant C such that qw (n) ≤ Cn, for every n ≥ 1. Hence h(w) = 0. A much more general result is uncovered below in Theorem 20.8. 236 / 332 Proposition 20.4 Let w ∈ ΩAS \ A+. Then h(w) = 1 if and only if every u ∈ ΩAS is a factor if w. Proof. Since h(w) = inf 1 n logm qw (n), we have h(w) = 1 if and only if qw (n) = mn for every n, that is if and only if every finite word is a factor of w. But every element of ΩAS is the limit of a sequence of finite words. Hence every finite word is a factor of w if and only if every pseudoword is a factor of w. 237 / 332 An ideal in a semigroup S is a nonempty subset I such that SI ∪ IS ⊆ I. A minimal ideal is an ideal which is not properly contained in any other ideal. Some semigroups, such as A+, have no minimal ideals. Note that a semigroup cannot have more than one minimal ideal: given two ideals I and J, IJ is an ideal contained in both of them. When there is one minimal ideal, we call it the minimum ideal, since it is indeed contained in every ideal. The minimum ideal of a semigroup is often also called its kernel and denoted K(S). For a semigroup S, the elements of the minimum ideal, if it exists, are precisely those elements of S which admit every element of S as a factor. 238 / 332 Proposition 20.5 Every compact metric semigroup has a minimum ideal. Proof. Let S be a compact metric semigroup. For each n, by compactness S is covered by a finite number of open balls of radius 1 n . If we choose such balls and collect their centers for varying n in a set, we obtain a countable dense subset X. Let X = {x1, x2, . . . , xn, . . .} be an enumeration of the elements of X. Let s be any limit of a subsequence of the sequence (x1 · · · xn)n. Then s admits every element of X as a factor and, since X is dense in S, every element of S must be a factor of S. Hence s ∈ K(S). 239 / 332 In particular, every finite and every profinite semigroup has a minimum ideal. Proposition 20.4 means that w ∈ ΩAS has maximum entropy (one) if and only if w ∈ K(ΩAS). Theorem 20.6 ([AV06b]) Let w = w(v1, . . . , vr ) where the vi ∈ ΩAS (i = 1, . . . , r) and w ∈ Ωr S. The entropy operator satisfies the following inequality: h(w ) ≤ max{h(w) logm r, h(v1), . . . , h(vr )}. (13) An ideal I of a semigroup S is said to be prime if st ∈ I implies s ∈ I or t ∈ I. Corollary 20.7 The minimum ideal of ΩAS is prime. 240 / 332 Theorem 20.8 ([AV06b]) Let u1, . . . , un ∈ ΩnS and let (v1, . . . , vn) = [u1, . . . , un]ω ΩAS (x1, . . . , xn). Then the entropy operator satisfies the following inequality: max 1≤i≤m h(vi ) ≤ max 1≤i≤m h(ui ). The following result is a considerable strengthening of Corollary 20.9. Corollary 20.9 For m > 1, the set ΩmS \ K(ΩmS) is a subsemigroup of ΩmS which is closed under composition and iteration. 241 / 332 All the results we have presented involving complexity and entropy apply to ΩAV with V a pseudovariety containing LSl: all we require is to be able to test finite factors, that is sets of the form (ΩAV)1 w(ΩAV)1 = A∗wA∗, with w ∈ A+ , which we need to be clopen; since A∗ wA∗ is LSl-recognizable, such sets are indeed clopen by Theorem 5.9. 242 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 243 / 332 Some remarks on relatively free profinite semigroups Let V be a pseudovariety of semigroups containing K = [[xωy = xω]]. Recall that the completion mapping ι : A+ → ΩAV is an injective homomorphism (more generally, whenever N ⊆ V). We identify each word w ∈ A+ with its image ι(w). The elements of A+ in ΩAV are called finite words and the remaining elements are said to be infinite. The natural projection pK : ΩAV → ΩAK maps each finite word to itself and every infinite pseudoword w to an infinite word pK(w) ∈ AN (cf. Theorem 21.1), which is called the infinite prefix of w and is denoted −→w . Dually, if V contains the pseudovariety D = [[xyω = yω]] and w ∈ ΩAV, then pD(w) is either w, if w is a finite word, or an infinite word in AZ− otherwise. In the latter case, we call pD(w) the infinite suffix of w and denote it ←−w . 244 / 332 Hence, if V contains both K and D (whose join can be easily shown to be the pseudovariety LI), then every infinite pseudoword w has an infinite prefix −→w and an infinite suffix ←−w . Thus, every factorization w = uv in which the factor are both infinite pseudowords determines a pair (←−u , −→v ) ∈ AZ− × AN of infinite words. A first natural question is whether every infinite pseudoword w ∈ ΩAV admits such factorizations. The following theorem gives a much more precise result. It is a particular case of a more general phenomenon [Alm95, Theorem 5.6.1]. The proof presented here could be extended to the general case. 245 / 332 Theorem 21.1 Let V be an arbitrary pseudovariety and let w ∈ ΩAV be a pseudoword which does not belong to ι(A+). Then there is some factorization of the form w = xyωz, with x, y, z ∈ ΩAV. 246 / 332 Proof. We do not distinguish w ∈ A+ from ι(w) ∈ ΩAV. To prove the theorem, we claim that the set of all products of the form xyω z = xyω · yω · yω z, with x, z ∈ A∗ and y ∈ A+ , is dense in ΩAV \ A+ . The theorem follows by compactness. Let w be an element of ΩAV \ A+ . Then there is some sequence (wn)n of words in A+ such that lim wn = w and lim |wn| = +∞. Given a positive integer N, consider a complete list of representatives of the classes of isomorphism of A-generated semigroups from V with at most N elements and let M be the cardinality of their direct product S. This may be a huge number, but it is certainly finite. Taking n sufficiently large, we conclude that there is a word v = wn such that |v| ≥ M and d(w, v) < 2−N . By construction, S comes equipped with a (“diagonal”) generating mapping ϕ : A → S. The inequality d(w, v) < 2−N means that ˆϕ(w) = ˆϕ(v). Now, distributing the homomorphism ˆϕ over the individual letters that compose the word v, we obtain a product in S of M = |S| factors. By Lemma 7.1, there is a factorization v = xyz, where x, z ∈ A∗ and y ∈ A+ , such that ˆϕ(xyω z) = ˆϕ(v) = ˆϕ(w). Hence d(w, xyω z) < 2−N , which proves the claim. 247 / 332 Corollary 21.2 Let V be an arbitrary pseudovariety. Then every infinite pseudoword in ΩAV admits factorizations into the product of two infinite pseudowords. 248 / 332 Regularity in profinite semigroups An element s of a semigroup S is said to be regular if there exists t ∈ S such that sts = s. Recall that two elements of a semigroup S are R-equivalent (respectively L-equivalent, J -equivalent) if they are prefixes (respectively suffixes, factors) of each other. The equivalence relation D is the smallest such relation containing both R and L. As a relation, it is the composite R ◦ L = L ◦ R. The intersection of the relations R and L is denoted H. Since R ∪ L ⊆ J , we also have D ⊆ J . In a profinite semigroup, (xy)ω = (xy)ω x · y(xy)ω−1 = (xy)ω−1 x · y(xy)ω and so y(xy)ω L (xy)ω R (xy)ωx. 249 / 332 Proposition 21.3 Let S be a profinite semigroup. (1) If s, t ∈ S are such that st J s then st R s. (2) If s, t ∈ S are such that st J t then st L t. (3) The relations J and D coincide in S. Proof. (1) Let x, y ∈ S1 be such that s = xsty. Then the equalities s = xn s(ty)n hold for all n ≥ 1, whence so does the equality s = xω s(ty)ω . Hence, we have s = s(ty)ω = st · y(ty)ω−1 . The left-right dual statement (2) is proved by dualizing the preceding argument. (3) Let s, t ∈ S be J -equivalent elements. Then there exist x, y ∈ S1 such that t = xsy. By (1) and (2), we deduce that t = xsy R xs L s. 250 / 332 Proposition 21.4 The following conditions are equivalent for a J -class J of a profinite semigroup S: (1) there is some idempotent in J; (2) there is some regular element in J; (3) all elements of J are regular; (4) there are some elements s, t ∈ J such that st ∈ J. The statements of Propositions 21.3 and 21.4 should be compared with the example in Exercise 1.3. 251 / 332 Proof of Proposition 21.4. (1) ⇒ (4) and (3) ⇒ (2) are obvious. (2) ⇒ (1) Suppose that s ∈ J and sts = s for some t ∈ S. Then st is an idempotent R-equivalent to s, and so J contains idempotents. (4) ⇒ (3) For such elements s, t ∈ J, Proposition 21.3 entails the relations t L st R s, whence stx = s and t = yst for some x, y ∈ S1. It follows that s = stx = systx = sys, whence s is regular. For an arbitrary u ∈ J, by Proposition 21.3(3) there exist v, w, z ∈ S1 such that uv = zs and uvw = u. Hence, u = uvw = zsvw = zsysvw = u · ysvw, which shows that the product of u with another element of J belongs to J, and we have shown that such an element u must be regular. 252 / 332 Proposition 21.5 Let S be a profinite semigroup. Then the Green relations J , R, L, and H are closed subsets of S × S. In particular, the equivalence classes for each of these relations are closed sets. Proof. The proof for the first three relations is similar, while the case of H follows from H = R ∩ L. Thus, we consider only the case of R. Suppose that (sn, tn)n is a convergent sequence of pairs of R-equivalent elements of S and let s = lim sn and t = lim tn. For each n ≥ 1, there are xn, yn ∈ S1 such that tn = snxn and sn = tnyn. By compactness, we may assume that the sequences (xn)n and (yn)n converge. Let x = lim xn and y = lim yn. Then t = lim tn = lim snxn = sx and s = lim sn = lim tnyn = ty. Hence s and t are R-equivalent. Taking one of the sequences to be constant yields that R-classes are closed. 253 / 332 Maximal subgroups If a subsemigroup of a semigroup S is a group then it is also called a subgroup of S. By a maximal subgroup of a semigroup S we mean a subgroup which is not properly contained in any other subgroup. Note that all elements of a subgroup G of S are H-equivalent in G and, therefore, also in S. Hence, every subgroup of S is contained in the H-class of some idempotent. Proposition 21.6 Let S be a profinite semigroup.9 Then the maximal subgroups of S are the H-classes of S that contain idempotents. They are closed subgroups and, therefore, they are profinite groups. 9 Except for the statements involving topology, the assumption that S is profinite is superfluous but is sufficient for our purposes and considerably simplifies the argument. For the general algebraic result, see, for instance, [CP61] or [Lal79]. 254 / 332 Proof. Let e be an idempotent of S and let H be its H-class. If s ∈ H, then there are elements x, y, z, t ∈ S1 such that e = sx = ys and s = ez = te. It follows that se = es = s and e = s · exe = ete · s. Hence e is an identity element for H and every every element of H admits both a left and a right inverse with respect to e by Proposition 21.3. To show that H is a subgroup of S, it remains to prove that H is a subsemigroup. Indeed, if s is another element of H, then yss = es = s and so ss L s again by Proposition 21.3. By symmetry, we deduce that ss ∈ H. By the remarks preceding the statement of the proposition, H is a maximal subgroup and every maximal subgroup must be of this type. By Proposition 21.5, every H-class is closed, whence maximal subgroups are closed. Finally, from our definition of profinite semigroup, it is immediate that every closed subsemigroup of a profinite semigroup is also profinite. 255 / 332 Proposition 21.7 Let S be a profinite semigroup. Then all maximal subgroups in the same D-class are isomorphic.10 Proof. Let e and f be two D-equivalent idempotents in S and let He and Hf be their respective H-classes. Let x, y, z, t be such that exfye = e, exf = zf , and tzf = f . We claim that the mapping ϕ that sends each h ∈ He to fyhxf is an isomorphism He → Hf . For h1, h2 ∈ He, we have fyh1xf · fyh2xf = fyh1 · exfye · h2xf = fyh1h2xf and ex · fyhxf · ye = exfye · h · exfye = h, so that ϕ is injective. Similarly, one shows that ϕ is surjective. Since ϕ is continuous, it is indeed an isomorphism. 10 Again, the purely algebraic statement holds for arbitrary semigroups, but we treat only the case of interest in these notes. 256 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 257 / 332 Symbolic Dynamics Let A be a finite alphabet. For a function w ∈ AZ, we may sometimes denote by wi the letter w(i).11 Writing the letters wi in the usual integer order, we obtain a biinfinite word · · · w−2w−1 · w0w1w2 · · · where the · is meant to distinguish an origin for such a word. Otherwise, just knowing the letters and not the indices, we will not be able to recover the function w (unless it is constant) but only a family of such functions, namely all that can be obtained by placing the origin somewhere. 11 This is avoided when considering sequences of such w. 258 / 332 For w ∈ AZ, we let σ(w) : Z → A i → wi+1. In terms of biinfinite words, this operation corresponds to shifting the origin past one letter to the right. For this reason, σ is called the shift (function). w : · · · w−2w−1 · w0w1w2 · · · σ(w) : · · · w−2w−1w0 · w1w2 · · · For i ≤ j, we write w[i, j] to represent the word wi wi+1 · · · wj . This is often called a block of w, but we prefer to call it a finite factor of w. 259 / 332 Given two infinite words x ∈ AZ− , y ∈ AN we may consider the biinfinite word w ∈ AZ defined by w(i) = x(i) if i ∈ Z− y(i) if i ∈ N. We denote w by x · y. We already saw how to multiply a finite word u by an infinite word v ∈ AN, namely concatenating u at the beginning of v. Dually, there is a natural multiplication vu for v ∈ AZ− and u ∈ A+. Thus, for instance if 0 < i ≤ j, we have the following “factorization”for w ∈ AZ, where the notation w(−∞, n] and w[n, +∞) is self-explanatory: w = w(−∞, −1] · w[0, i − 1] w[i, j] w[j + 1, +∞). 260 / 332 The mapping Z → SAZ n → σn is a homomorphism into the permutation group of AZ, defining a natural action of the additive group of integers on the set AZ. The set AZ has a profinite structure. For u, v ∈ AZ define d(u, v) = 2−r(u,v), where r(u, v) is the smallest nonnegative integer n such that un = vn or u−n = v−n. As usual, we put min ∅ = +∞ and 2−∞ = 0. 261 / 332 Exercise 22.1 Show that: (1) the above function d is an ultrametric on the set AZ; (2) the inequalities d σ(u), σ(v) ≤ 2d(u, v) d σ−1 (u), σ−1 (v) ≤ 2d(u, v) hold for all u, v, ∈ AZ; (3) the function σ is a homeomorphism. 262 / 332 Note that, for a real number > 0, the ball of radius centered at a biinfinite word w ∈ AZ, B (w), consists of all biinfinite words v ∈ AZ such that vi = wi for all indices i satisfying the inequality 2|i| ≥ . In Symbolic Dynamics, such a set is often called a cylinder. The reason is that viewing AZ as the product of Z copies of the set A, the ball B (w) is just the set of all biinfinite words that have the same projection as w in the components corresponding to the interval [−n, n], where n = − log2 . Such sets form a basis of the product topology on the set AZ, where A is viewed with the discrete topology. Thus our metric has the same open sets as that topology. 263 / 332 In particular, since the alphabet is finite, the space AZ is totally bounded. If (wn)n is a Cauchy sequence of elements of AZ, then, given k > 0, there is Nk such that, for all n ≥ Nk, the word wn[−Nk, Nk] is constant. We may define a biinfinite word w ∈ AZ by letting w(i) = wN|i| (i). Note that w = lim wn. Proposition 22.2 The space AZ is complete, whence compact, being totally bounded. 264 / 332 By a subshift (also known as a shift space, a shift dynamical system, or simply a shift) we mean a nonempty closed subset of AZ which is stable under the action of Z. The stability condition for a subset X of AZ is equivalent to σ(X) ⊆ X and σ−1 (X) ⊆ X. The space AZ is also known as the full shift, which explains the terminology. For a subshift X ⊆ AZ , the set of all words of the form w[i, j] with w ∈ X is called the language of X and it is denoted L(X). A language L ⊆ A+ is said to be factorial if it contains all nonempty factors of its members; prolongable if it is nonempty and, for every u ∈ L there are letters a, b ∈ A such that au, ub ∈ L. Proposition 22.3 The correspondence X → L(X) is a bijection between the set of subshifts X ⊆ AZ and the set of factorial prolongable languages of A+. 265 / 332 Proof. From the definition of L(X), it is clear that L(X) is factorial and prolongable. Suppose that L ⊆ A+ is a factorial prolongable language. Let X be the set of all biinfinite words w ∈ AZ all of whose finite factors belong to L. Then X is stable under the action of Z. If w = lim wn is the limit of a sequence of elements of X, then every finite factor of w must be a factor of wn for all sufficiently large n, whence it belongs to L. Hence X is a subshift which, by definition, satisfies L(X) ⊆ L. By prolongability, any word u ∈ L can be extended arbitrarily on both sides, which allows us to define, inductively, a biinfinite word w all of whose finite factors lie in L. Moreover, w must belong to any subshift Y ⊆ AZ such that L ⊆ L(Y). Hence u ∈ L(X) and X ⊆ Y for any such Y. Thus L = L(X) and L(X) ⊆ L(Y) implies X ⊆ Y. 266 / 332 The subshift corresponding to the factorial prolongable language L ⊆ A+ is denoted XL. Given a language K ⊆ A+, the exclusion of K is the language consisting of all words that do not admit any word from K as a factor. In other words, the exclusion of K is the language A+ \ A∗KA∗. Exercise 22.4 A language is factorial if and only if it is the exclusion of some language. 267 / 332 Special types of subshifts A subshift X ⊆ AZ is of finite type if L(X) is the exclusion of some finite language, i.e., the ideal A+ \ L(X) is finitely generated. Equivalently, there is a positive integer n and a set F ⊆ An such that w ∈ AZ belongs to X if and only if all factors of w of length n belong to F. The subshift X is then also denoted by X[F]. From a set K ⊆ An one can build a graph Γ(K) with vertex set An−1 and an edge u → v if there are letters a, b ∈ A such that the word ub = av belongs to K. A word has all its factors of length n in K if and only if the sequence of its consecutive factors of length n − 1 describe a path in the graph Γ(K). A biinfinite word belongs to X[F] if and only if the biinfinite sequence of its consecutive factors of length n − 1 determines a biinfinite path in the graph Γ(An \ F). 268 / 332 The graph Γ(An+1) is known as the de Bruijn graph of order n. Given any finite directed graph Γ, considering the set of edges A as an alphabet, the set of all biinfinite paths determines a subshift XΓ ⊆ AZ, which is known as the edge subshift of Γ. Starting with the subshift of finite type X[F], every element w ∈ X[F] determines a biinfinite path in Γ(An \ F) and therefore an element Φ(w) ∈ XΓ(An\F). The correspondence Φ : X[F] → XΓ(A\F) is easily checked to be a homeomorphism which commutes with the shift σ. More generally, for subshifts X ⊆ AZ and Y ⊆ BZ, an encoding is a continuous function ϕ : X → Y such that the following diagram commutes: X σA GG ϕ  X ϕ  Y σB GG Y A bijective encoding is called a conjugation. 269 / 332 One simple way of constructing encodings is through sliding windows as above. Let A and B be two alphabets and let m, n ∈ Z be such that m + n ≥ 0. Given a function f : Am+n+1 → B we may define a function ¯f : AZ → BZ as follows: for w ∈ AZ, ¯f (w)i = f (w[i − m, i + n]). This is called a sliding block encoding with memory m and anticipation n. · · · wi−m−2wi−m−1 wi−m · · · wi−1wi wi+1 · · · wi+n f   wi+n+1 · · · · · · ¯f (w)i−2 ¯f (w)i−1 ¯f (w)i ¯f (w)i+1 · · · If m = n = 1, then the encoding (respectively conjugation) is said to be a one-block encoding (respectively one-block conjugation). 270 / 332 Theorem 22.5 (Curtis-Lyndon-Hedlund Theorem) Every encoding between subshifts is obtained by restriction of a sliding block encoding. Proof. It is an easy exercise to verify that sliding block encodings are indeed encodings. Let X ⊆ AZ and Y ⊆ BZ be subshifts over finite alphabets and suppose that ϕ : X → Y is an encoding. In particular, ϕ is a continuous function between compact metric spaces and, therefore, it is uniformly continuous. Hence, there exists a positive integer n such that d(w1, w2) < 2−n implies d ϕ(w1), ϕ(w2) < 1 2. This implies that ϕ(w)0 depends only on the interval w[−n.n] and thus allows us to define a function A2n+1 ∩ L(X) → B, which we may extend arbitrarily to a function f : A2n+1 → B. By construction, the resulting sliding block encoding ¯f extends ϕ to an encoding AZ → BZ. 271 / 332 Corollary 22.6 Every conjugation is of the form ϕ ◦ ψ−1, where both ϕ and ψ are one-block conjugations. Proof. By Theorem 22.5, every conjugation ξ : X → Y, with X ⊆ AZ and Y ⊆ BZ, is obtained by restriction of a sliding block encoding ¯f : AZ → BZ, where f : A2n+1 → B. Consider the one-block encoding ¯g : (A2n+1)Z → AZ defined by the mapping g : A2n+1 → A that sends a word of length 2n + 1 to its middle letter. Let ¯h : (A2n+1)Z → BZ be the one-block encoding defined by h = f . Then we have ξ = ¯h ◦ ¯g−1|X . A major open problem in Symbolic Dynamics is to classify the subshifts of finite type up to conjugacy. See [LM96] for a more thorough introduction to this subject and partial results. 272 / 332 A subshift X ⊆ AZ is sofic if the language L(X) is regular. If L ⊆ A+ is a factorial prolongable regular language, then L may be recognized by a finite automaton in which every state is both initial and final: starting from the minimal deterministic finite automaton recognizing L, trim it by throwing away all states that either are not accessible from the initial state or from them no final state is accessible; then, throw away all states that either are not accessible from a nontrivial (i.e., having at least one edge) strongly connected component or from them no such component is accessible; finally, make all remaining states both initial and final. Note that, for such an automaton, the subshift XL consists precisely of the biinfinite words w ∈ AZ which can be obtained by reading the labels of a biinfinite path in the automaton. 273 / 332 Examples Let A = {0, 1} be a two-letter alphabet. Consider the language F = {11}. The corresponding exclusion subshift X[F] is known as the golden mean subshift. It is conjugate to the edge subshift of the following graph (Γ(A2 \ F)), where the conjugacy (an encoding with memory 0 and anticipation 1) is given by the blue labels: 0 1 01 10 00 274 / 332 The reason for the name given to the subshift is that, if we consider the adjacency matrix of the above graph, 1 1 1 0 then, as we already observed in another context, the Perron-Frobenius eigenvalue is precisely the golden mean 1+ √ 5 2 . The significance of this eigenvalue is that its log2 is the entropy of the language L(X[F]) = A+ \ A∗11A∗ (defined, for a language L by inf 1 n log|A| |An ∩ L|).12 12 One can show that the entropy is a conjugacy invariant (see [LM96] and Section 25). 275 / 332 We may turn the same graph into the following automaton (where all states are considered both initial and final): 0 0 1 The corresponding subshift is known as the even subshift. Its language L consists of all words over the alphabet A = {0, 1} such that, in every factor of the form 10n1, n is even. Thus, A+ \ L is the ideal generated by the (minimum) set {102n+11 : n ≥ 0}, and therefore it is not finitely generated. Hence the even subshift is a sofic subshift which is not of finite type. 276 / 332 Irreducible subshifts A language L is irreducible if, for all u, v ∈ L, there exists a word w such uwv ∈ L. A subshift X is irreducible if the language L(X) is irreducible. Note that: an edge subshift is irreducible if and only if the graph that defines it is strongly connected; a factorial prolongable regular language L is irreducible if and only if itcan be recognized by a finite strongly connected automaton with all states both initial and final. 277 / 332 Minimal subshifts A subshift X ⊆ AZ is minimal if X does not properly contain another subshift. For a subshift X ⊆ AZ, and w ∈ X, the closure in AZ of the orbit O(w) = {σn(w) : n ∈ Z} is a subshift contained in X. Hence, in a minimal subshift, every orbit is dense. Thus, for a minimal subshift X, all w ∈ X have the same finite factors. With a little extra effort (exercise!), it follows that every minimal subshift is irreducible. A trivial example of minimal subshift is XL, where L consists of all factors of powers of a fixed word. Such a subshift is called periodic. By the pumping lemma, every sofic subshift contains some periodic subshift. Hence, the only sofic minimal subshifts are the periodic ones. 278 / 332 One way to produce minimal subshifts is to iterate primitive substitutions. In this context, a substitution is an endomorphism ϕ of a free semigroup A+. Primitivity has exactly the same meaning as before. A primitive substitution ϕ determines a language L(ϕ) consisting of all words which are factors of some ϕn(a), where a is a letter. Since L(ϕ) is clearly factorial and can be easily checked to be prolongable, it follows that it is the language of a subshift Xϕ. See [Fog02] for a thorough treatment of subshifts generated by substitutions and their significance in the Theory of Dynamical Systems. Before justifying that Xϕ is a minimal subshift, we introduce a few further useful notions. 279 / 332 For a factorial language L ⊆ A+ and a word u ∈ L, a return word of u in L is a word v ∈ A+ such that vu ∈ L, u is a prefix of vu, and u only appears as a factor of vu in the prefix and suffix positions. A factorial language L is recurrent if, for every u ∈ L, there is no infinite word all of whose finite factors belong to L for which u is not a factor. A factorial language L ⊆ A+ is uniformly recurrent if it is recurrent and, for every u ∈ L, there are only finitely many return words of u in L. In case L is uniformly recurrent, let n be the maximum of the lengths of the return words of u in L and let N = n + |u|. Then every word in L of length N contains u as a factor. Conversely, if there is such an N, then L is uniformly recurrent. 280 / 332 In case the language L is the language of finite factors of a biinfinite word w ∈ AZ, then we say respectively that u is a return word for u in w and that w is recurrent or uniformly recurrent. Proposition 22.7 Let w ∈ AZ be a biinfinite word. Then the subshift O(w) is minimal if and only if w is uniformly recurrent. 281 / 332 Proof. We have already observed that a subshift is minimal if and only if all its elements have the same finite factors. (⇒) Suppose that the finite factor u of w has an infinite sequence uzn of return words. Without loss of generality, we may assume that the sequence (zn)n converges in ΩAS, say to a pseudoword z. By the choice of the words zn, u is not a factor of zn for any n, and therefore it is not a factor of z. Since z is an infinite pseudoword, by Corollary 21.2, there is a factorization z = xy, with both x, y infinite. If we glue the infinite suffix of x with the infinite prefix of y, we obtain a biinfinite word w = ←−x · −→y all of whose finite factors are factors of w. Hence w belongs to O(w) and u should be a factor of w while, by construction, it is not. Thus, w can only have finitely many return words of u. (⇐) Let v ∈ O(w) and let u be a finite factor of w. We should show that u is also a factor of v. Let N be such that every factor of w of length N contains u as a factor. Since d(v, σm (w)) < 2− N 2 , for some m, we deduce that u is a factor of v. 282 / 332 Corollary 22.8 Let L ⊆ A+ be a factorial prolongable language. Then the following conditions are equivalent: (1) L is recurrent; (2) L is uniformly recurrent; (3) L is the language of finite factors of some minimal subshift. Proof. Since L is factorial and prolongable, there is a unique subshift X ⊆ AZ such that L = L(X).. (1) ⇒ (3) Since L is recurrent, every w ∈ X must have the same finite factors, namely all the words from L, whence X is minimal. (3) ⇒ (2) follows from Proposition 22.7. (2) ⇒ (1) is clear from the definitions. 283 / 332 Corollary 22.9 Let ϕ be a primitive substitution over an alphabet A. Then Xϕ is a minimal subshift. Proof. We verify that the language L(ϕ) is uniformly recurrent. Let N be such that, for all a, b ∈ A, a appears in the word ϕN(b). Let u be a word from L(ϕ). Then u is a factor of ϕn(a) for some n ≥ 1. Every v ∈ L(ϕ) of length 2 max b∈A |ϕN+n (b)| − 1 has some factor of the form ϕN+n(b) = ϕn(ϕN(b)). Since a is a factor of ϕN(b) and u is a factor of ϕn(a), it follows that u is factor of v. 284 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 285 / 332 Going profinite Suppose that L ⊆ A+ is a factorial prolongable language. The closure L of L in ΩAS provides a natural subset of ΩAS associated with L. Since A+ is a discrete subspace of ΩAS (more generally, this holds for ΩAV for every pseudovariety V containing N by Proposition 7.3), there is no new finite word in L, i.e., we have L ∩ A+ = L. Hence, symbolic dynamics can also be studied by investigating suitable closed subset of ΩAS. Exactly which subsets? 286 / 332 A subset K of a semigroup S is said to be factorial if, whenever u, v ∈ S are such that u is a factor of v, if v belongs to K, then so does u. A subset K ⊆ ΩAS is prolongable if, for every w ∈ K, there are a, b ∈ A such that aw, wb ∈ K. Proposition 23.1 ([Cos06, AC09]) Let L ⊆ A+ be a language. (1) The language L is prolongable in A+ if and only if so is L in ΩAS. (2) The language L is factorial in A+ if and only if so is L in ΩAS. Thus, the suitable closed subsets of ΩAS to study symbolic dynamics in the profinite world are precisely the closed factorial prolongable subsets. The implications (⇐) are both easy. 287 / 332 While the preservation of prolongability is a simple exercise, the preservation of factoriality is more delicate. The proof requires some preparation, and it is better to put it in the right framework, replacing S by a more general pseudovariety V with suitable properties. We say that a pseudovariety V is closed under concatenation if so is the corresponding variety of languages. A mapping between two metric spaces is said to be open if the image of every open set is open. Proposition 23.2 ([AC09, Lemma 2.3]) Let V be a pseudovariety of semigroups containing N. Then V is closed under concatenation if and only if the multiplication in ΩAV is an open mapping for every finite alphabet A. 288 / 332 Proof. By Theorem 5.9 and Proposition 7.3, a language L over a finite alphabet A is V-recognizable if and only if L is open. Hence, if the multiplication is open and K, L ⊆ A+ are V-recognizable, then KL = K L is the product of two open sets, i.e., the image under multiplication of an open subset of ΩAV × ΩAV, whence an open subset of ΩAV, so that KL is V-recognizable. Conversely, suppose that V is closed under concatenation. By Theorems 5.8 and 5.9, every open subset of ΩAV is the union of sets of the from L, where L ⊆ A+ is V-recognizable. Since multiplication distributes over union, using again Theorem 5.9, it suffices to show that KL = K L is open in ΩAV whenever K, L ⊆ A+ are V-recognizable, which follows once more from Theorem 5.9 since V is closed under concatenation. 289 / 332 Lemma 23.3 ([AC09, Lemma 2.5]) Let S be a metric semigroup and suppose that the multiplication S × S → S is an open mapping. Let u, v ∈ S and suppose that the sequence (wn)n converges to uv. Then there is a subsequence (wnk )k whose terms admit a factorization wnk = unk vnk such that lim unk = u and lim vnk = v. Proof. For each k ≥ 1, the product B1 k (u) B1 k (v) of open balls is open. Hence, it is possible to choose nk such that wnk ∈ B1 k (u) B1 k (v) and n1 < n2 < n3 < · · · . For each k, there is therefore a factorization wnk = unk vnk with unk ∈ B1 k (u) and vnk ∈ B1 k (v). By construction, we have lim unk = u and lim vnk = v. 290 / 332 Proposition 23.4 ([AC09, Proposition 2.4]) Let V be a pseudovariety closed under concatenation containing N and let L ⊆ A+ be a factorial language. then the set L is factorial in ΩAV. Proof. Suppose that L ⊆ A+ is a factorial language, w ∈ L, and w = uv is a factorization in ΩAV. Then there is a sequence (wn)n of elements of L converging to w. By Lemma 23.3, we may assume that each wn admits a factorization wn = unvn, with lim un = u and lim vn = v. By Theorem 21.1, since V contains N, un and vn must be finite words and therefore they belong to the factorial language L. Hence u, v ∈ L. Every factor of w can be obtained from w in at most a couple of such factorization steps. 291 / 332 The special cases If L is the complement of a finitely generated ideal A∗FA∗ of A+, then A∗FA∗ and L are regular languages, A∗FA∗ and L are clopen subsets of ΩAS and L = ΩAS \ A∗FA∗ = ΩAS \ (ΩAS)1 F(ΩAS)1 . Hence L is a clopen subset of ΩAS which is the complement of a finitely generated ideal. Proposition 23.5 Every clopen finitely generated ideal I of ΩAS is of the form I = A∗FA∗ for some finite subset F of A+. Proof. By Theorem 5.9, the set J = I ∩ A+ is a regular language such that J = I. Since A+ is a subsemigroup of ΩAS, J is an ideal of A+. Since no finite word admits an infinite pseudoword as a factor, J must also be finitely generated. 292 / 332 Thus, subshifts of finite type over the alphabet A correspond to the complements of clopen finitely generated ideals of ΩAS. Sofic subshifts over the alphabet A correspond to clopen factorial prolongable subsets of ΩAS. What about irreducible and minimal subshifts? A subset K of a semigroup S is said to be irreducible if, for all u, v ∈ K, there is some w ∈ S such that uwv ∈ K. Proposition 23.6 Let L ⊆ A+ be an arbitrary language. Then L is irreducible in A+ if and only if L is irreducible in ΩAS. 293 / 332 Proof. (⇒) Given two elements u, v ∈ L, let (un)n and (vn)n be sequences of elements of L converging, respectively, to u and v. Since L is irreducible, there is a sequence (wn)n of words such that unwnvn ∈ L. By compactness, we may assume that (wn)n converges to some w ∈ ΩAS. By continuity of multiplication, it follows that uwv ∈ L. (⇐) Suppose that L is irreducible in ΩAS and let u, v ∈ L. Then there exists w ∈ ΩAS such that uwv ∈ L. Let (xn)n be a sequence of words from L converging to uwv. By Lemma 23.3, we may assume that, for every n ≥ 1, there is a factorization xn = unwnvn such that lim un = u, lim wn = w, and lim vn = v. Since A+ is a discrete subsemigroup of ΩAS by Proposition 7.3, we have un = u and vn = v for all sufficiently large n. Hence uwnv ∈ L for all sufficiently large n. 294 / 332 Proposition 23.7 Let K ⊆ ΩAS be an irreducible subset. Then there is in K some element w such that every other element of K is a factor of w. In other words, there is a unique J -class containing elements of K such that every other such J -class is J -above it. Moreover, this J -class is regular. The special J -class of Proposition 23.7 is denoted J (K). In case K = L(X) for an irreducible subshift X ⊆ AZ, we also denote J (K) by J (X). 295 / 332 Proof. Let I be the collection of all closed ideals of ΩAS which have nonempty intersection with K. Given two such ideals I and J, there exist u ∈ I ∩ K and v ∈ J ∩ K. Since K is irreducible, there is some w ∈ ΩAS such that uwv ∈ K. Since I and J are ideals, uwv belongs to I ∩ J, which shows that I ∩ J ∈ I. By induction on the number of terms, it follows that the intersection of any finite subcollection of I also belongs to I. Hence, again by compactness, the intersection I = I is a closed ideal of ΩAS with nonempty intersection with K. Each element of I ∩ K must generate I, as an ideal, whence they are all J -equivalent and their J -class is the required one. Since K is irreducible, if u ∈ I ∩ K, then there is some w ∈ ΩAS such that uwu ∈ K, whence uwu ∈ I and so uwu J u. By Proposition 21.4, we conclude that the J -class of u is regular. 296 / 332 Of the special families of subshifts, it remains to consider the minimal subshifts. Theorem 23.8 ([Alm05a]) Let X ⊆ AZ be a subshift. Then X is minimal if and only if L(X) \ A+ is contained in some J -class of ΩAS. Moreover, such a J -class is a J -maximal regular J -class and every J -maximal regular J -class of ΩAS contains a set of the form L(X) \ A+ for a unique minimal subshift X ⊆ AZ. Recall that every minimal subshift is irreducible. Note that the J -class containing L(X) \ A+ associated with a minimal subshift X is precisely the J -class J (X) introduced earlier for an irreducible subshift. 297 / 332 Proof of Theorem 23.8. Suppose first that the subshift X is minimal. We claim that, if u ∈ L(X) \ A+ and v ∈ ΩAS \ A+ is such that F(v) ⊆ L(X), then u is a factor of v. It follows that, for all u, v ∈ L(X) \ A+, u and v are factors of each other, and so they are J -equivalent. The claim also entails that every u ∈ L(X) \ A+ is a factor of all its infinite factors, and so the J -class J of u is J -maximal among J -classes containing infinite pseudowords. Since, by Theorem 21.1 every infinite pseudoword has some idempotent factor (or using Proposition 23.7), the J -class J is regular. To prove the claim, let (un)n and (vn)n be sequences of elements of L(X) such that lim un = u and lim vn = v. Since A+ is a discrete subspace of ΩAS, we have lim |un| = lim |vn| = +∞. By Proposition 22.7, there is a strictly increasing sequence of indices (nk)k such that uk is a factor of vnk . Since ΩAS is a compact semigroup, it follows that u is a factor of v, as claimed. 298 / 332 (. . . ) Conversely, suppose that the set L(X) \ A+ is contained in a unique J -class J of ΩAS. To establish that X is minimal, we show that L(X) is uniformly recurrent. Indeed, otherwise, there would be a word u ∈ L(X) and a sequence (vn)n of words from L(X) such that u is not a factor of any of the vn and lim |vn| = +∞. By compactness, we may assume that (vn)n converges to some v ∈ ΩAS. Since the language A∗uA∗ is regular, u is not a factor of v. On the other hand, since the language L(X) is prolongable, there is also a sequence (un)n of elements of L(X) converging to some u ∈ ΩAS \ A+ such that u is a factor of each un. Hence u and v are both infinite pseudowords in L(X) but one has u as factor and the other does not, which contradicts the assumption that they are J -equivalent. 299 / 332 (. . . ) Finally, suppose that J is a J -maximal regular J -class of ΩAS and let L be the language consisting of all finite factors of the elements of J. Note that L is factorial and prolongable, whence it is the language of finite factors of a unique subshift X ⊆ AZ. By the usual compactness argument, every element of L \ A+ is a factor of the elements of J and, therefore, it must belong to J. By the above, it follows that X is minimal. Since, by the choice of L, it contains every other subshift Y ⊆ AZ such that L(Y) \ A+ ⊆ J, it is the only such minimal subshift. 300 / 332 Note that we have also proved the following result. Corollary 23.9 Let X ⊆ AZ be a minimal subshift. Then a pseudoword w ∈ ΩAS belongs to J (X) if and only if F(w) = L(X). A pseudoword w ∈ ΩAS is said to be uniformly recurrent if the language F(w) is uniformly recurrent. Corollary 23.10 The uniformly recurrent pseudowords in ΩAS are the elements of J -maximal regular J -classes. 301 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 302 / 332 Free and projective profinite groups Let H be a pseudovariety of groups. Given a pro-H group G, a diagram of the form G ϕ  θ || H ψ GG GG K (14) where ϕ is a continuous homomorphism, ψ is a homomorphism, and H ∈ H, is called an embedding problem for G.13 A solution of the embedding problem (14) is an onto continuous homomorphism θ such that the diagram commutes. A weak solution of the embedding problem (14) is a continuous homomorphism θ such that the diagram commutes. 13 The terminology comes from Galois theory, see [FJ86]. 303 / 332 A pro-H group G is said to be H-projective if every embedding problem (14) has a weak solution. A G-projective profinite group is also said to be projective. A pro-H group G is said to be H-free of countable rank if every embedding problem (14) has a solution.14 Proposition 24.1 ([RZ00, Lemma 7.6.3]) Let H be a pseudovariety of groups. (1) Every H-projective pro-H group is isomorphic to a closed subgroup of a free pro-H group. (2) Suppose that H is closed under extension. Then a pro-H group is H-projective if and only if it is isomorphic to a closed subgroup of a free pro-H group. 14 The terminology is justified by showing that this property is equivalent to G possessing an universal property similar to freeness, namely being“free on a (countable) set converging to 1”. See [RZ00] for details. 304 / 332 A pseudovariety of groups H is said to be arborescent if (H ∩ Ab) ∗ H = H. Proposition 24.2 ([RZ00, Proposition 7.6.7]) Let H be an arborescent pseudovariety of groups. Then a pro-H group is projective if and only if it is H-projective. Theorem 24.3 ([RS08]) Let V be a pseudovariety of semigroups such that (V ∩ Ab) ∗ V = V. Then every closed subgroup of a free pro-V semigroup is projective. 305 / 332 Outline Semigroups of implicit operators Complexity Complexity of pseudowords generated by iteration of substitutions Entropy Interlude on profinite semigroups Symbolic Dynamics Going profinite Interlude on profinite groups Maximal subgroups 306 / 332 Maximal subgroups Let X ⊆ AZ be an irreducible subshift. Consider the J -minimal J -class J (X) of ΩAS which contains elements of L(X) and recall that it is regular. By Proposition 21.7, all maximal subgroups of ΩAS contained in J (X) are isomorphic profinite groups. Viewed as an abstract profinite group, we denote any of them by G(X). By Theorem 24.3, G(X) is a projective profinite group associated with the subshift. By a conjugacy invariant for a class of subshifts we mean a function defined on that class whose value is the same for conjugate subshifts. Theorem 25.1 ([Cos06]) The group G(X) is a conjugacy invariant of the subshift X. 307 / 332 Can one somehow compute these groups? For the question to make sense from an algorithmic point of view, both the subshift and the group should be somehow described by finite data. There are two natural families of subshifts described by finite data, namely sofic subshifts and subshifts defined by the iteration of primitive substitutions. 308 / 332 The sofic case In the case of a periodic subshift X, the group G(X) is free procyclic by Theorem 18.6. For all other irreducible sofic subshifts, the group is in a sense much larger as stated in the following theorem. Theorem 25.2 ([CS11]) Let X be a non-periodic irreducible sofic subshift. Then the profinite group G(X) is free of countable rank. 309 / 332 Presentations of profinite groups By a presentation of a profinite group we mean a pair X; R , where X is a set and R is a binary relation on the free profinite group on the set X. We only consider here the case where the set of generators X is a finite set. Thus, R ⊆ ΩX G × ΩX G. Usually we write the pairs (u, v) of R as formal equalities u = v and call them the relations of the presentation. The presentation is finite if R is also a finite set. Let N be the closed normal subgroup of ΩAG generated by the elements of the form uv−1 with u = v a relation of the presentation. 310 / 332 A natural metric on the quotient group ΩX G/N is given by letting d(uN, vN) = 2−r(u,v), where r(u, v) is the minimum cardinality of a finite group G for which there exists a continuous homomorphism ϕ : ΩX G → G such that ϕ(N) = {1} and ϕ(u) = ϕ(v). Exercise 25.3 The quotient group ΩX G/N is profinite for the above metric. The profinite group thus constructed is called the profinite group defined by the presentation X; R and is usually also denoted by X; R . A profinite group is said to admit the presentation X; R if it isomorphic to the profinite group X; R . 311 / 332 Proposition 25.4 ([Lub01, AC10]) Every finitely generated closed subgroup of ΩAS admits a finite presentation of the form X; Φω (x) = x (x ∈ X) where Φ is a continuous endomorphism of ΩAG. We say that an X-generated profinite group G is decidable if, given any mapping X → H into a finite group H, there is an algorithm to determine whether or not ϕ extends to a continuous homomorphism G → H. We say that a continuous endomorphism of ΩX G is a finite group substitution if it induces an endomorphism of the free group Ωκ X G. 312 / 332 Proposition 25.5 Suppose that Φ is a finite group substitution over a finite alphabet X. Then the profinite group X; Φω(x) = x (x ∈ X) is decidable. We say that a primitive finite substitution ϕ is periodic if the subshift Xϕ is periodic. There is an algorithm to test if a primitive finite substitution is periodic [Pan86, HL86]. 313 / 332 For a primitive finite substitution ϕ, we also denote the J -class J (Xϕ) by J (ϕ) and the profinite group G(Xϕ) by G(ϕ). Theorem 25.6 ([AC10]) There is an algorithm that solves the following problem: Input: a non-periodic primitive finite substitution over a finite alphabet A; Output: a finite group substitution Φ over a finite alphabet X such that G(ϕ) admits the presentation X; Φω(x) = x (x ∈ X) . 314 / 332 The generating set X can be explicitly realized in a maximal subgroup contained in J (ϕ) as follows: let ba be a two-letter word in L(ϕ) such that ϕω (a) starts with a and ϕω (b) ends with b; such a word ba is called a connection for ϕ; let R(ba) be the finite set of all return words of ba in Xϕ (which may be effectively computed [Dur98, Alm05a]); let X = b−1 R(ba)b, which is a code in the sense that it generates a free subsemigroup of A+ ; by a theorem of Margolis, Sapir and Weil [MSW98], the closed subsemigroup of ΩAS generated by ϕω (X) is freely generated by this set as a profinite semigroup; by the choice of the word ba, some power ϕN of ϕ maps X+ into itself; hence ϕN induces an endomorphism of X+ which determines a continuous endomorphism Φ of ΩX G; a mapping showing that the profinite group G(ϕ) admits the presentation X; Φω (x) = x (x ∈ X) is given by x ∈ X → ϕω (x). 315 / 332 Corollary 25.7 Let ϕ be a primitive finite substitution over a finite alphabet A. Then the group G(ϕ) is decidable. A substitution ϕ is said to be proper if all ϕ(a) (a ∈ A) start with the same letter and all ϕ(a) (a ∈ A) end with the same letter. Theorem 25.8 ([AC10]) Let ϕ is a proper primitive finite substitution over a finite alphabet A. Then the group G(ϕ) admits the presentation A; ϕω (a) = a (a ∈ A) . Theorem 25.9 ([Alm05a]) Let ϕ be a primitive finite substitution over a finite alphabet A which induces an automorphism of the free group Ωκ AG. Then G(ϕ) is a finitely generated free profinite group and the number of generators can be effectively computed. 316 / 332 Examples Consider the Fibonacci substitution ϕ(a) = ab, ϕ(b) = a. Since ϕ induces an automorphism of the free group Ωκ {a,b}G, the group G(ϕ) is a free profinite group by Theorem 25.9. More precisely, the H-class of ϕω(aω) is freely generated by {ϕω(aba), ϕω(ababa)}. The reason for this is that a2 is a connection and its return words are aab = a · aba · a−1 and a2 bab = a · ababa · a−1 . Let x = aba and y = ababa. Then ϕ2 maps X+ into itself. The continuous endomorphism Φ of ΩX G is given by Φ(x) = xy and Φ(y) = xy2. Since Φ induces an automorphism of the free group Ωκ X G, Φω is the identity mapping and the relations in the presentation X; Φω(x) = x (x ∈ X) are trivial. 317 / 332 More generally, we have the following result. A subshift X over a two-letter alphabet A is said to be Sturmian if |L(X)| ∩ An = n + 1. Equivalently, every w ∈ L(X) \ A+ is Sturmian. Theorem 25.10 ([Alm05a, Corollary 6.1]) Let X be any Sturmian subshift. Then G(X) is a free profinite group on two generators. The proof of this result was obtained by a suitable approximation of an arbitrary Sturmian subshift by Sturmian subshifts generated by substitutions. 318 / 332 A non-free profinite maximal subgroup Consider the proper substitution defined by ϕ(a) = ab and ϕ(b) = a3b. The only connection for ϕ is the word ba, whose return words are ba and ba3 and so X = {ab, a3b}. Letting x = ab and y = a3b and taking N = 1, the continuous endomorphism Φ of the free group Ωκ X G is given by Φ(x) = xy and Φ(y) = x3y, that is, it looks exactly like the original substitution. The maximal subgroup H of ΩAS containing ϕω(X) is also generated by {ϕω(a), ϕω(b)}. By Theorem 25.8, the group G(ϕ) admits the presentation a, b; ϕω (a) = a, ϕω (b) = b . One can use it to prove the following result. 319 / 332 Theorem 25.11 ([AC10]) For the above proper substitution, the group G(ϕ) is not H-free for any pseudovariety of groups H. Proof. Suppose first that H is free on a pseudovariety H containing the two element additive group Z/2Z. Since {ϕω (a), ϕω (b)} generates H, it freely generates H as a pro-H group. Hence the mapping that sends both ϕω (a) and ϕω (b) to 1 extends to an onto continuous homomorphism H → Z/2Z. But it sends both generators ϕω (ab) and ϕω (a3 b) to 0, and so it should be a constant mapping. Hence, H is not free pro-H for any pseudovariety containing Z/2Z. To conclude the proof, it suffices to show that H has a finite continuous image of even order. We claim that in fact the alternating group A5 is such an image. To prove the claim, it suffices to show that the generating 3-cycles α = (1 2 3) and β = (3 4 5) verify the defining relations for our presentation, i.e., that (α, β) is a periodic point of the implicit operator [ab, a3 b]A5 . Indeed, direct calculation shows that it has period 12 for this operator. 320 / 332 The case of the Prouhet-Thue-Morse substitution To conclude, consider the Prouhet-Thue-Morse substitution τ(a) = ab, τ(b) = ba. All two-letter words are connections for τ. We choose a2. The return words of a2 in Xτ are aababb, aababbab, aabb, aabbab and the set X consists of the words x = ab2 a, y = abab2 a, z = ab2 aba, t = abab2 aba. The subsemigroup X+ is stabilized by τ2. Let α = τω(x), β = τω(y), γ = τω(z), and δ = τω(t). A simple calculation shows that βα−1γ = δ, and so the group H-class H of ϕω(a2) is generated by {α, β, γ}. Moreover the substitution on X+ associated with τ2 is given by Φ(x) = zxy, Φ(y) = ztxy, Φ(z) = zxty, Φ(t) = ztxty. 321 / 332 By Theorem 25.6 and the description of the algorithm following it, H admits the presentation x, y, z, t; Φω (u) = u (u = x, y, z, t) . Taking account the equality βα−1γ = δ, we may eliminate the generator t, expressing Φ only in terms of the three remaining generators to obtain the presentation x, y, z; Ψω (u) = u (u = x, y, z) where Ψ(x) = zxy, Ψ(y) = zyx−1zxy, and Ψ(z) = zxyx−1zy. One may use similar techniques to those described in the proof of Theorem 25.11 to prove that H cannot be generated by fewer than three elements and also the following result. Theorem 25.12 ([AC10]) The profinite group G(τ) is not H-free for any pseudovariety of groups H. 322 / 332 Final remarks Although we have an algorithm to determine if a finite group is a quotient of G(τ), we still do not know whether the pseudovariety generated by all finite quotients of this profinite group is G. We also do not know if the profinite group G(X) of a minimal subshift X is always decidable. Finally, we do not know if there is any profinite group of the form G(X) which is not free but which is H-free for some proper pseudovariety of groups H. Much of the connections with symbolic dynamics presented in Part 3 of the lectures carry over to pseudovarieties V containing LSl. Of course, when groups are involved, the results further depend on the intersection V ∩ G. For instance, if H = V ∩ G is commutative and satisfies a pseudoidentity of the form xp = 1, where p is prime, then every group in ΩAV is free pro-H, being a vector space over the field Z/pZ. 323 / 332 References [ABR92] D. Albert, R. Baldinger, and J. Rhodes, The identity problem for finite semigroups (the undecidability of), J. Symbolic Logic 57 (1992), 179–192. [AC09] J. Almeida and A. Costa, Infinite-vertex free profinite semigroupoids and symbolic dynamics, J. Pure Appl. Algebra 213 (2009), 605–631. [AC10] , Presentations of Sch¨utzenberger groups of minimal subshifts, Tech. Report CMUP 2010-1, Univ. Porto, 2010. [AE03] J. Almeida and A. Escada, Semidirect products with the pseudovariety of all finite groups, Proceedings of the International Conference Words, Languages and Combinatorics (Kyoto, March, 2000) (Singapore) (M. Ito and T. Imaoka, eds.), World Scientific, 2003, pp. 1–21. [Alm95] J. Almeida, Finite semigroups and universal algebra, World Scientific, Singapore, 1995, English translation. [Alm02] , Dynamics of implicit operations and tameness of pseudovarieties of groups, Trans. Amer. Math. Soc. 354 (2002), 387–411. [Alm05a] , Profinite groups associated with weakly primitive substitutions, Fundamentalnaya i Prikladnaya Matematika (Fundamental and Applied Mathematics) 11 (2005), no. 3, 13–48, In Russian. English version in J. Math. Sciences 144, No. 2 (2007) 3881–3903. 324 / 332 [Alm05b] , Profinite semigroups and applications, Structural Theory of Automata, Semigroups, and Universal Algebra (New York) (Valery B. Kudryavtsev and Ivo G. Rosenberg, eds.), NATO Science Series II: Mathematics, Physics and Chemistry, vol. 207, Springer, 2005, Proceedings of the NATO Advanced Study Institute on Structural Theory of Automata, Semigroups and Universal Algebra, Montr´eal, Qu´ebec, Canada, 7-18 July 2003, pp. 1–45. [AS00a] J. Almeida and B. Steinberg, On the decidability of iterated semidirect products and applications to complexity, Proc. London Math. Soc. 80 (2000), 50–74. [AS00b] , Syntactic and global semigroup theory, a synthesis approach, Algorithmic Problems in Groups and Semigroups (J. C. Birget, S. W. Margolis, J. Meakin, and M. V. Sapir, eds.), Birkh¨auser, 2000, pp. 1–23. [AS01] J. Almeida and P. V. Silva, SC-hyperdecidability of R, Theor. Comp. Sci. 255 (2001), 569–591. [AS03] K. Auinger and B. Steinberg, On the extension problem for partial permutations, Proc. Amer. Math. Soc. 131 (2003), 2693–2703. [Ash87] C. J. Ash, Finite semigroups with commuting idempotents, J. Austral. Math. Soc., Ser. A 43 (1987), 81–90. [Ash91] , Inevitable graphs: a proof of the type II conjecture and some related decision procedures, Int. J. Algebra Comput. 1 (1991), 127–146. 325 / 332 [AT01] J. Almeida and P. G. Trotter, The pseudoidentity problem and reducibility for completely regular semigroups, Bull. Austral. Math. Soc. 63 (2001), 407–433. [AV06a] J. Almeida and M. V. Volkov, Subword complexity of profinite words and subgroups of free profinite semigroups, Int. J. Algebra Comput. 16 (2006), 221–258. [AV06b] , Subword complexity of profinite words and subgroups of free profinite semigroups, Int. J. Algebra Comput. 16 (2006), 221–258. [AZ07] J. Almeida and M. Zeitoun, An automata-theoretic approach of the word problem for ω-terms over R, Theor. Comp. Sci. 370 (2007), 131–169. [Bau65] G. Baumslag, Residual nilpotence and relations in free groups, J. Algebra 2 (1965), 271–282. [Ber79] J. Berstel, Transductions and context-free languages, B. G. Teubner, Stuttgart, 1979. [BS73] J. A. Brzozowski and I. Simon, Characterizations of locally testable events, Discrete Math. 4 (1973), 243–271. [Cos06] A. Costa, Conjugacy invariants of subshifts: an approach from profinite semigroup theory, Int. J. Algebra Comput. 16 (2006), 629–655. [CP61] A. H. Clifford and G. B. Preston, The algebraic theory of semigroups, vol. I, Amer. Math. Soc., Providence, R.I., 1961. 326 / 332 [CS11] A. Costa and B. Steinberg, Profinite groups associated to sofic shifts are free., Proc. London Math. Soc. 102 (2011), 341–369. [CT04] J. C. Costa and M. L. Teixeira, Tameness of the pseudovariety LSl, Int. J. Algebra Comput. 14 (2004), 627–654. [Del01] M. Delgado, On the hyperdecidability of pseudovarieties of groups, Int. J. Algebra Comput. 11 (2001), 753–771. [Dur98] F. Durand, A characterization of substitutive sequences using return words, Discrete Math. 179 (1998), no. 1-3, 89–101. [Eil76] S. Eilenberg, Automata, languages and machines, vol. B, Academic Press, New York, 1976. [ELR75] A. Ehrenfeucht, K. P. Lee, and G. Rozenberg, Subword complexities of various classes of deterministic developmental languages without interaction, Theor. Comp. Sci. 1 (1975), 59–75. [FJ86] M. D. Fried and M. Jarden, Field arithmetic, Springer, Berlin, 1986. [Fog02] N. Pytheas Fogg, Substitutions in dynamics, arithmetics and combinatorics, Lecture Notes in Mathematics, vol. 1794, Springer-Verlag, Berlin, 2002. [GAP06] The GAP Group, GAP – Groups, Algorithms, and Programming, Version 4.4, 2006, (http://www.gap-system.org). 327 / 332 [Hal50] M. Hall, A topology for free groups and related groups, Ann. Math. 52 (1950), 127–139. [Hen88] Karsten Henckell, Pointlike sets: the finest aperiodic cover of a finite semigroup, J. Pure Appl. Algebra 55 (1988), 85–126. [HL86] T. Harju and M. Linna, On the periodicity of morphisms on free monoids, RAIRO Inf. Th´eor. et Appl. 20 (1986), 47–54. [HM38] G. A. Hedlund and M. Morse, Symbolic dynamics, Amer. J. Math. 60 (1938), 815–866. [HMPR91] K. Henckell, S. Margolis, J.-E. Pin, and J. Rhodes, Ash’s type II theorem, profinite topology and Malcev products. Part I, Int. J. Algebra Comput. 1 (1991), 411–436. [HR91] K. Henckell and J. Rhodes, The theorem of Knast, the PG=BG and Type II Conjectures, Monoids and Semigroups with Applications (Singapore) (J. Rhodes, ed.), World Scientific, 1991, pp. 453–463. [Hun83] R. P. Hunter, Some remarks on subgroups defined by the Bohr compactification, Semigroup Forum 26 (1983), 125–137. [Hun88] , Certain finitely generated compact zero-dimensional semigroups, J. Austral. Math. Soc., Ser. A 44 (1988), 265–270. [KM02] I. Kapovich and A. Myasnikov, Stallings foldings and subgroups of free groups, J. Algebra 248 (2002), 608–668. 328 / 332 [KP86] Jiˇri Kad’ourek and Libor Pol´ak, On the word problem for free completely regular semigroups, Semigroup Forum 34 (1986), 127–138. [KR65] K. Krohn and J. Rhodes, Algebraic theory of machines. I. Prime decomposition theorem for finite semigroups and machines, Trans. Amer. Math. Soc. 116 (1965), 450–464. [Lal79] G. Lallement, Semigroups and combinatorial applications, Wiley-Interscience, J. Wiley & Sons, Inc., New York, 1979. [LM96] D. Lind and B. Marcus, An introduction to symbolic dynamics and coding, Cambridge University Press, Cambridge, 1996. [Lub01] A. Lubotzky, Pro-finite presentations, J. Algebra 242 (2001), 672–690. [McC01] J. McCammond, Normal forms for free aperiodic semigroups, Int. J. Algebra Comput. 11 (2001), 581–625. [MP71] R. McNaughton and S. Papert, Counter-free automata, MIT Press, Cambridge, MA, 1971. [MP84] S. W. Margolis and J.-E. Pin, Varieties of finite monoids and topology for the free monoid, Proc. 1984 Marquette Semigroup Conference (Milwaukee), Marquette University, 1984, pp. 113–129. [MP87] , Inverse semigroups and varieties of finite semigroups, J. Algebra 110 (1987), 306–323. 329 / 332 [MS93] F. Mignosi and P. S´e´ebold, Morphismes sturmiens et r`egles de rauzy, J. Th´eor. Nombres Bordeaux 5 (1993), 221–233. [MSW98] S. Margolis, M. Sapir, and P. Weil, Irreducibility of certain pseudovarieties, Comm. Algebra 26 (1998), 779–792. [MSW01] , Closed subgroups in pro-V topologies and the extension problem for inverse automata, Int. J. Algebra Comput. 11 (2001), 405–445. [Num57] K. Numakura, Theorems on compact totally disconnetced semigroups and lattices, Proc. Amer. Math. Soc. 8 (1957), 623–626. [Pan84] J.-J. Pansiot, Complexit´e des facteurs des mots infinis engendr´es par morphismes it´er´es, Automata, Languages and Programming (Antwerp, 1984) (Berlin), Lect. Notes Comput. Sci., no. 172, Springer, 1984, pp. 380–389. [Pan86] , Decidability of periodicity for infinite words, RAIRO Inf. Th´eor. et Appl. 20 (1986), 43–46. [Pin95] J.-E. Pin, BG=PG: A success story, Semigroups, Formal Languages and Groups (Dordrecht) (J. Fountain, ed.), vol. 466, Kluwer, 1995, pp. 33–47. [PR91] J.-E. Pin and C. Reutenauer, A conjecture on the Hall topology for the free group, Bull. London Math. Soc. 23 (1991), 356–362. [Pro51] E. Prouhet, M´emoire sur quelques relations entre les puissances des nombres, C. R. Acad. Sci. Paris 33 (1851), 31. 330 / 332 [PS85] J.-E. Pin and H. Straubing, Monoids of upper triangular matrices, Semigroups: structure and universal algebraic problems (Amsterdam) (G. Poll´ak, ed.), North-Holland, 1985, pp. 259–272. [Rei82] J. Reiterman, The Birkhoff theorem for finite algebras, Algebra Universalis 14 (1982), 1–10. [RS08] J. Rhodes and B. Steinberg, Closed subgroups of free profinite monoids are projective profinite groups, Bull. London Math. Soc. 40 (2008), no. 3, 375–383. [RZ93] L. Ribes and P. A. Zalesski˘ı, On the profinite topology on a free group, Bull. London Math. Soc. 25 (1993), 37–43. [RZ00] , Profinite groups, Ergeb. Math. Grenzgebiete 3, no. 40, Springer, Berlin, 2000. [Sch65] M. P. Sch¨utzenberger, On finite monoids having only trivial subgroups, Inform. and Control 8 (1965), 190–194. [Sim75] I. Simon, Piecewise testable events, Proc. 2nd GI Conf. (Berlin), Lect. Notes in Comput. Sci., vol. 33, Springer, 1975, pp. 214–222. [Sta83] J. R. Stallings, Topology of finite graphs, Inventiones Mathematicae 71 (1983), 551–565. 331 / 332 [Ste01] B. Steinberg, Inevitable graphs and profinite topologies: some solutions to algorithmic problems in monoid and automata theory, stemming from group theory, Int. J. Algebra Comput. 11 (2001), 25–71. [Ste11] , On the endomorphism monoid of a profinite semigroup, Portugal. Math. 68 (2011), 177–183. [Sti73] P. Stiffler, Extension of the fundamental theorem of finite semigroups, Advances in Math. 11 (1973), 159–209. [Thu06] A. Thue, ¨Uber unendlichen zeichenreihen, Kra. Vidensk. Selsk. Skrifter, I. Mat. Nat. Kl. (1906), no. 7, 1–22. [Thu12] A. Thu¨e, ¨Uber die gegenseitige Lage gleicher Teile gewisser Zeichenreihen, Norske Vid. Selsk. Skr. I. Mat.-nat. Kl., Christiana 1 (1912), 1–67. [ˇS63] A. I. ˇSirˇsov, On certain near-engel groups, Algebra i Logika 2 (1963), no. 5, 5–18. [Zor36] M. Zorn, Nilpotency of finite groups (abstract), Bull. Amer. Math. Soc. 42 (1936), 485–486. 332 / 332