321
Views
0
CrossRef citations to date
0
Altmetric
Research Article

On the word problem for weakly compressible monoids

Pages 4731-4745 | Received 23 Aug 2022, Accepted 25 Apr 2023, Published online: 30 Jun 2023

Abstract

We study the language-theoretic properties of the word problem, in the sense of Duncan & Gilman, of weakly compressible monoids, as defined by Adian & Oganesian. We show that if C is a reversal-closed super-AFL, as defined by Greibach, then M has word problem in C if and only if its compressed left monoid L(M) has word problem in C. As a special case, we may take C to be the class of context-free or indexed languages. As a corollary, we find many new classes of monoids with decidable rational subset membership problem. Finally, we show that it is decidable whether a one-relation monoid containing a non-trivial idempotent has context-free word problem. This answers a generalization of a question first asked by Zhang in 1992.

2020 Mathematics Subject Classification:

1 Introduction

The word problem for finitely presented monoids was introduced in 1914 by Thue [Citation39]. Though this problem would eventually turn out to be undecidable in general, via the remarkable work by Markov [Citation31] and Post [Citation38], there are many tractable cases. Particularly the word problem for one-relation monoids has garnered a great deal of attention. Though it remains a tantalizing open problem whether this problem is decidable, much can be said (see especially the recent survey [Citation35]). Thue himself studied this problem, and solved the problem for some cases when the single defining relation is of the form w = 1. Such monoids are known as special one-relation monoids. Adian [Citation1, Citation2] studied special monoids in-depth, and proved that the word problem is decidable for all special one-relation monoids, via a reduction to the word problem for one-relator groups; this latter problem is decidable by Magnus’ famous theorem [Citation26]. Adian’s student Makanin [Citation28, Citation29, Citation33] later extended this to show that the word problem for any special k-relation monoid reduces to the word problem for some k-relator group. The author has recently proved similar language-theoretic reductions for special monoids [Citation36].

The work by Adian on special monoids would later be extended by Adian and his student Oganesian in the joint work [Citation3]. They defined a form of “compression,” which is called weak compression by the author in [Citation35]. A monoid M with defining relations ui = vi (for 1ik) is called weakly compressible if there is some non-empty word α such that every ui and vi begins and ends with α. Associated to any weakly compressible monoid is a “compressed” left monoid L(M), in which the sum of the lengths of the defining relations are shorter than in the original monoid. The main idea is that the word problem for M can be reduced to the word problem for L(M). In particular, by combining this with Adian’s result, this solves the word problem in the class of subspecial one-relation monoids, being the class of monoids which can be compressed to a special one-relation monoid. Subspecial monoids were studied already by Lallement [Citation24], and are (essentially) the one-relation monoids containing a non-trivial idempotent.

Weak compression of subspecial monoids has received some further attention in the past few decades. Zhang [Citation41] extended the above results on the word problem to solve the conjugacy problem in certain subspecial monoids. Kobayashi [Citation22] used compression to prove that all one-relation monoids satisfy the topological finiteness condition FDT, for which the subspecial case was the only outstanding case. Building on this, Gray & Steinberg showed that all one-relation monoids satisfy the homological finiteness property FP, at the same time studying the algebraic properties of subspecial monoids [Citation14].

In another direction, a highly successful approach to studying the word problem for groups came through the methods of formal language theory. Such methods were introduced to group theory by An i¯s i¯mov [Citation5], and focus on the language-theoretic properties of the set of all words representing the identity element, calling this set the word problem of the group (its properties are generally independent of the finite generating set chosen). An i¯s i¯mov showed that the word problem of a group is a regular language if and only if the group is finite, and that the class of groups with context-free word problem is closed under free products. A little over ten years later, Muller & Schupp (supplemented by [Citation12]) gave a remarkable algebraic charactersiation, generally referred to as the Muller–Schupp theorem: a finitely generated group has context-free word problem if and only if it is virtually free [Citation32].

The language of words representing the identity element in a monoid is generally not very enlightening (though for special monoids it is [Citation36], as in the case of groups). To remedy this, Duncan & Gilman introduced a new language which encodes equality in a monoid [Citation11]. This set, which is also termed the word problem for the monoid, has many of the attractive properties of the analogous set for groups, such as invariance under generating set chosen. Furthermore, it is easy to show that a monoid has regular word problem if and only if it is finite, mimicking An i¯s i¯mov’s result. The word problem also generally behaves well under taking free products [Citation9, Citation37]. This notion of the word problem has also been studied in e.g. [Citation17, Citation18].

This paper seeks to study the language-theoretic properties of Adian & Oganesian’s weak compression. We will paint a fairly complete picture. The main theorem of this article (Theorem 1.1) will involve classes of languages C which are super- AFLs, as introduced by Greibach [Citation15]. Such classes generalize the context-free languages and the indexed languages; roughly speaking, a full AFL is a super-AFL if it is closed under “recursion” (see Section 2.2 for details). We will prove the following main theorem:

Theorem 1.1.

Let M be a weakly compressible monoid, and let L(M) be its left monoid. Let C be a reversal-closed super- AFL. Then M has word problem in C if and only if L(M) has word problem in C.

An overview of the article is as follows. In Section 2 we give some notation, useful concepts, and define two language-theoretic operations from [Citation37]. In Section 3, we define weak compression, amalgamating several authors’ definitions and notations. In Section 4, we prove Theorem 1.1. Finally, in Section 5 we give some corollaries of Theorem 1.1, particularly focused on context-free monoids. In particular, we will obtain many monoids for which the rational subset membership problem is decidable (Corollary 5.1); and that it is decidable whether a one-relation monoid containing a non-trivial idempotent has context-free word problem (Theorem 5.4).

This article is a condensed form of Chapter 4 of the author’s PhD thesis [Citation34].

2 Notation and auxiliary results

We assume the reader is familiar with the fundamentals of formal language theory. In particular, a full AFL (abstract family of languages) is a class of languages closed under homomorphism, inverse homomorphism, intersection with regular languages, union, concatenation, and the Kleene star. For some background on this, and other topics in formal language theory, we refer the reader to standard books on the subject [Citation6, Citation16, Citation19]. The paper also assumes familiarity with the basics of the theory of semigroup, monoid, and group presentations, which will be written as SgpA|R,MonA|R, and GpA|R, respectively. For further background see e.g. [Citation2, Citation10, Citation25, Citation27]. We also refer the reader to [Citation36, Citation37].

2.1 Monoids, words, rewriting

Let A be a finite alphabet, and let A denote the free monoid on A, with identity element denoted ε or 1, depending on the context. Let A+ denote the free semigroup on A, i.e. A+=A{ε}. For u,vA, by uv we mean that u and v are the same word. For wA, we let |w| denote the length of w, i.e. the number of letters in w. We have |ε|=0. If wa1a2an for aiA, then we let wrev denote the reverse of w, i.e. the word anan1a1. For a language L, we define Lrev as the collection of all words wrev where wL. We say that a class of languages C is reversal-closed if for every LC, we have LrevC. We say that the word wA is self-overlap free if it is empty, or else if it is non-empty and none of the proper non-empty prefixes of w is also a proper non-empty suffix of w. Thus xyxyy is self-overlap free, but xyxyx is not. If the words u,vA are equal in the monoid M=MonA|R, then we denote this u=Mv. Finally, when we say that a monoid M is generated by a finite set A, we mean that there exists a surjective homomorphism π:AM. In this case, u=Mv will be used synonymously with π(u)=π(v).

We give some notation for rewriting systems. For an in-depth treatment and further explanations of the terminology, see e.g. [Citation8, Citation21]. A rewriting system R on A is a subset of A×A. An element of R is called a rule. The system R induces several relations on A. We will write uRv if there exist x,yA and a rule (l,r)R such that uxly and vxry. We let R denote the reflexive and transitive closure of R. We denote by *R the symmetric, reflexive, and transitive closure of R. The relation *R defines the least congruence on A containing R. For XA, we let XR denote the set of ancestors of X, i.e. XR={wA|xX such that wRx}. The monoid MonA|R is identified with the quotient A/*R. For a rewriting system TA×A and a monoid M=MonA|R, we say that T is M-invariant if for every rule (u,v)T, we have u=Mv. That is, T is M-invariant if and only if *T*R.

A rewriting system RA×A is said to be monadic if (u,v)R implies |u||v| and vA{ε}. We say that R is special if (u,v)R implies vε. Every special system is monadic. Let C be a class of languages. A monadic rewriting system R is said to be C if for every aA{ε}, the language {u|(u,a)R} is in C. Thus, we may speak of e.g. monadic C-rewriting systems or monadic context-free rewriting systems.

Let M be a monoid with a finite generating set A. One way of encoding the structure of M as a language is by using the identity problem of M (with respect to A), defined as(2.1) IPAM={wA|w=M1}.(2.1)

When M is a group, this is the standard definition of the language-theoretic word problem, appearing e.g. in Muller & Schupp [Citation32] and Anisimov [Citation5]. However, for general monoids, (2.1) is generally not very useful. Instead, the central definition of this article will be the following. The (monoid) word problem of M (with respect to A) is defined as the language(2.2) WPAM:={u#vrev|u,vA,u=Mv},(2.2) where # is some fixed symbol not in A. For a class of languages C, we say that M has C-word problem if WPAM is in C. If C is closed under inverse homomorphism, then M having C-word problem does not depend on the finite generating set A chosen for M (see e.g. [Citation17]). The above definition (2.2) is due to Duncan and Gilman [Citation11, p. 522]. We will throughout this article use A# as a short-hand for the set A{#}.

2.2 Super-AFL s

The theorems in this paper involve a special type of classes of languages, called super- AFLs. These were introduced by Greibach [Citation15], using nested iterated substitutions. In [Citation37], the author gave an equivalent characterization of super-AFL s, which we shall use here. We begin with a simple definition. Recall that for a language L and a rewriting system R, the language LR denotes the language of ancestors of L under R, i.e. the set of words which can be rewritten to some word in L.

Definition 1.

Let C be a class of languages. Let RA×A be a rewriting system. Then we say that R is C-ancestry preserving if for every LA with LC, we have LRC. If every monadic C-rewriting system is C-ancestry preserving, then we say that C has the monadic ancestor property.

Example 1.

If RA×A is a monadic context-free rewriting system, and LA is a context-free language, then LR is a context-free language [Citation7, Theorem 2.2]. That is, if we let CF denote the class of context-free languages, then every monadic CF-rewriting system is CF-ancestry preserving; in other words, the class of context-free languages has the monadic ancestor property.

Definition 2.

Let C be a full AFL. Then C is said to be a super- AFL if C has the monadic ancestor property.

For example, the class CF of context-free languages forms a super-AFL by [Citation23, Theorem 1.2], and the class IND of indexed languages is also a super-AFL [Citation4, Citation13]. Both are closed under reversal. On the other hand, neither the classes REG nor DCF of regular, resp. deterministic context-free, languages are super-AFL s. Indeed, if C is a super-AFL, then CFC, by [Citation15, Theorem 2.2]. For more examples and generalizations, we refer the reader to the so-called hyper- AFLs defined by Engelfriet [Citation13], all of which are super-AFL s.

We remark that Greibach [Citation15] originally defines super-AFLs via “nested iterated substitutions”, which behave similarly to taking ancestors under monadic rewriting systems. The author [Citation37, Proposition 2.2] has proved that the above definition, using rewriting systems, is equivalent to Greibach’s, and so we will use her results about super-AFLs without restriction.

2.3 Alternating products and bipartisan ancestry

With general definitions taken care of, we will now turn to give some useful auxiliary language-theoretic results.

We first define an operation (the alternating product) on certain languages, which mimics the operation of the free product of semigroups. This operation appears in [Citation37]. Fix an alphabet A and let # be a symbol not in A. Let LA#A be any language. We say that L is concatenation-closed (with respect to #) if(2.3) u1#v1L and u2#v2Lu1u2#v2v1L,(2.3) where u1,v1,u2,v2A. The word problem of any finitely generated monoid is always a concatenation-closed language.

Let X:N{1,2} be a parametrisation such that either X(2j)=1 and X(2j+1)=2, or else X(2j)=2 and X(2j+1)=1, for all jN. A parametrisation X of this form will be called standard. Given two concatenation-closed languages in A#A, we now present an operation for combining them. Let L1,L2A#A be concatenation-closed. Then the alternating product L1L2 of L1 and L2 is defined as the language consisting of all words of the form:(2.4) u1u2uk#vkv2v1,(2.4) where for all 1ik we have ui#viLX(i), where X is a standard parametrisation. In particular, #L1L2 if and only if #L1 or #L2. Alternating products are modeled on the (semigroup) free product, as the following example shows.

Example 2.

Let L1={an#an|n0} and L2={bn#bn|n0}. Then, of course, L1=WP{a}{a} and L2=WP{b}{b}. Now both languages L1 and L2 are concatenation-closed, and it is easy to see that we haveL1L2={an1bn2ank#ankbn2an1|k0;n1,n2,,nk11;nk0}.

Thus L1L2={w#wrev|w{a,b}}=WP{a,b}{a,b}.

Using the monadic ancestor property, one can prove the following useful statement.

Proposition 2.1.

[Citation37, Proposition 2.3] Let C be a super- AFL. Let L1,L2A#A be concatenation-closed languages. Then L1,L2CL1L2C.

We will require another operation with useful preservation properties, also introduced in [Citation37]. Let R1,R2A×A be two rewriting systems. Let LA#A. Then we define the bipartisan (R1,R2)-ancestor of L as the language(2.5) R1LR2={w1#w2|u1#u2L such that wiuiRi for i=1,2}.(2.5)

Bipartisan ancestors can, informally speaking, manipulate the left and the right side (of # in words in L) using R1 resp. R2, and these manipulations can be independent of one another. We give an example of this independence.

Example 3.

Let A={a,b}, and let L={a#a}. Let R1 be the rewriting system with the rules (bn,a) for all n1. ThenR1LR1={bn1#bn2|n1,n21}{bn#a|n1}{a#bn|n1}{a#a}.

In particular, we have bn1#bn2R1LR1 even if n1n2.

Proposition 2.2.

[Citation37, Proposition 2.5] Let C be a class of languages closed under rational transductions. Let LA#A, and let R1,R2A×A be rewriting systems. If R1,R2 are C-ancestry preserving, then LCR1LR2C.

Bipartisan ancestors are useful in describing the language theory of monoid (and group) free products. See [Citation37] for further details. Alternating products and bipartisan ancestors are the language-theoretic tools we shall require for this article. We now turn to the main subject of the article.

3 Weak compression

We present weak compression as it is defined in the survey [Citation35, Section 3.1] by the author. This is an amalgamation of other approaches [Citation3, Citation14, Citation22, Citation24, Citation41].

Let A be an alphabet. We say that a pair (u, v) of words is sealed by wA+ if u,vwAAw. If a pair is sealed by some word, then it is clearly sealed by a unique self-overlap free word α. For example, (xyxpxyx, xyxqxyx) is sealed by xyx, but also by the self-overlap free word x.

Let M be the monoid defined by the presentation(3.1) MonA|ui=vi(1ip).(3.1)

Definition 3.

We say that the monoid defined by (3.1) is weakly compressible (with respect to α) if there is some self-overlap free word αA+ such that for all 1ip, the pair (ui , vi ) is sealed by α.

If M is not weakly compressible, then we say that M is incompressible. Any special monoid is incompressible by default (in particular groups are incompressible).

Let M be a weakly compressible monoid defined by (3.1). We will assume |A|>1, for otherwise M is a finite cyclic monoid, and all is trivial. A word wA is called a left α-conjugator if wαA. Let Σ(α) be the set of left α-conjugators which contain exactly one occurrence of α, i.e. Σ(α)=α(AAαA). For ease of notation, we write Σ instead of Σ(α). Clearly, Σ is a countably infinite suffix code (as |A|>1), and Σ+ is the set of all left α-conjugators. Enumerate the words of Σ as {w1,w2,} and fix a set Γ(α)=Γ of symbols {γw1,γw2,} in bijective correspondence with Σ via the map φ:wiγwi. As Σ is a suffix code, we can extend φ to an isomorphism of the free monoids Σ and Γ.

Every defining relation ui = vi in (3.1) is sealed by α, and as α is self-overlap free, we have ui,viΣα. We factor ui , vi uniquely over the suffix code Σ, yielding(3.2) uiui,1ui,2ui,miα,andvivi,1vi,2vi,niα.(3.2)

Any word ui,j or vi,k appearing in the factorizations (3.2) for some 1ip is called a left piece of M. The set of all left pieces of M is denoted Σ(α), which will be shortened to Σ. As M is finitely presented, Σ is a finite set of words. We let Γ=Γ(α) be the set φ(Σ) of symbols from Γ.

From the factorization (3.2), we define a new presentation(3.3) MonΓ|φ(ui,1ui,2ui,mi)=φ(vi,1vi,2vi,ni)(1ip).(3.3)

Definition 4.

The monoid defined by the presentation (3.3) is called the extended left monoid associated to the monoid M (defined by (3.1)), and is denoted L(M). The submonoid of L(M) generated by Γ is the left monoid associated to M, and is denoted L(M).

It is obvious from (3.3) that L(M)=FL(M), where * denotes the monoid free product, and F is a free monoid of countably infinite rank, freely generated by ΓΓ. We emphasize that it is always decidable to compute the presentation (3.3) starting from the weakly compressible (3.1), as it only requires computing with regular languages. Of course, L(M) is finitely presented, with the same defining relations as in (3.3). Furthermore, the sum of the lengths of the defining relations in L(M) is strictly shorter than the corresponding sum for M. In particular, setting L1(M)=L(M) and Li(M)=L(Li1(M)) for i > 1, there exists some n1 such that Ln(M) is incompressible.Footnote1

We give an example. One can easily verify that if(3.4) M1=Monx,y|xyyxxxyxxyyxxxy=xy,(3.4) then the defining relation of M1 is sealed by xy. Factorizing both sides of the defining relation, we find Σ={xyx,xyyxx} and hence Γ={γxyx,γxyyxx}. Thus(3.5) L(M1)=Monγxyx,γxyyxx|γxyyxxγxyxγxyyxx=1.(3.5)

This (special) monoid is incompressible. It is isomorphic to Mona,b|aba=1, which is isomorphic to the infinite cyclic group Z by removing the redundant generator b.

We return to the general case, and present the normal form results from [Citation3, Citation24]. First, note that it is obvious that if two words u,vA are equal in M, then u contains an occurrence of the self-overlap free word α if and only if v does; and furthermore, if neither u nor v contain α, then u=Mv if and only if uv. Second, given any uAαA, there exist unique u,uAAαA and uαAAα such that uuuu. We call such a factorization the canonical form of u, and u is called the α-part of u. Write uulα, where ulΣ. The word φ(ul) is called the γ-part of u. The following theorem is fundamental to weakly compressible monoids.

Theorem 3.1

(Adian & Oganesian, Lallement). Let M be a weakly compressible monoid defined by (3.1). Let u,vAαA have canonical forms uuu and vvv, respectively. Let γu,γv be the γ-parts of u and v, respectively. Then u=Mv if and only if (1) uv and uv; and (2) γu=γv in L(M). Furthermore, (2) is equivalent to (2’) u=Mv.

Remark 1.

The above result appears as [Citation24, Lemma 3.2] and [Citation3, Theorem 3.2] only in the case of one defining relation. However, the proofs of these results use no properties of one-relation monoids, and can, mutatis mutandis, also be used to prove the above result.

In particular, the word problem for M is decidable if and only if the word problem for L(M) is decidable. Furthermore, one can without much difficulty show that the left (right) divisibility problem for M reduces to the word and left (right) divisibility problems in L(M). In particular this solves the word and divisibility problems in the monoid M1 defined by (3.4).

4 Proof of Theorem 1.1

Let M be a weakly compressible (with respect to some self-overlap free word α) monoid defined by (3.1), and let φ,Σ,Γ,Σ, and Γ be as in Section 3. We will now prove that the language-theoretic properties of the monoid M defined by (3.1) can be reduced to the properties of the left monoid L(M). Let C be a reversal-closed super-AFL, which will remain fixed throughout this section.

To simplify some technical notation we shall sometimes consider Σ, rather than Γ, as a finite generating set for the monoid L(M), as there exists a surjective homomorphism πΣ from Σ onto L(M) given by the composition πΣ=πΓ°φ, where πΓ:ΓL(M) is the natural surjective homomorphism. However, it is important to notice that if u,vΣ, then generally πΣ(u)=πΣ(v) is very distinct from u=Mv, unlike what Theorem 3.1 may seem to suggest. Instead, using the canonical forms we find that we have(4.1) πΣ(u)=πΣ(v)uα=Mvα.(4.1)

By the first remark preceding Theorem 3.1, it follows that the language WPAM is a union WP[α]AMWα of the two disjoint languages(4.2) WP[α]AM={w1#w2rev|w1,w2AαA such that w1=Mw2},(4.2) (4.3) Wα={w#wrev|wAAαA}.(4.3)

The language Wα defined by (4.3) is the intersection of the context-free language WPAA with the regular language (AAαA)#(AAαA)rev. In particular, Wα is a context-free language. Any super-AFL contains the class of context-free languages [Citation15, Theorem 2.2]. It follows that as C is a super-AFL, we have that WP[α]AMC implies WPAMC, as C is closed under union.

Having reduced the language-theoretic properties of WPAM to those of WP[α]AM, we perform another reduction, which will yield Lemma 4.1. Let(4.4) WP[αα]AM={w1#w2rev|w1,w2αAAα such that w1=Mw2}.(4.4)

Of course, we have the strict inclusionsWP[αα]AMWP[α]AMWPAM.

Using this new language, we define the rewriting system(4.5) Rα={(w1#w2rev#)|w1#w2revWP[αα]AMWPAA}.(4.5)

This is a monadic rewriting system. As C is a super-AFL, if we now have WP[αα]AMC, then it follows that Rα is a C-rewriting system.

Lemma 4.1.

If WP[αα]AMC, then WPAMC.

Proof.

Suppose WP[αα]AMC. By the remarks following (4.2) and (4.3), it suffices to show that WP[α]AMC. We claim that(4.6) WP[α]AM=WPAARα(AαA#AαrevA).(4.6)

This suffices to establish that WP[α]AMC, as C is closed under intersection with regular languages; WPAA is a context-free language and hence in C; and C has the monadic ancestor property.

Note first that for every rule (s,t)Rα we have s,tWPAM. Hence, if wWPAARα, then it is clear by induction on the number of Rα-rewritings necessary to transform w into an element of WPAA that wWPAM. Hence the inclusion in (4.6) is proved.

Second, if wWP[α]AM, then wu#vrev for some u,vAαA with u=Mv. Let uuu and vvv be the canonical forms of u and v, respectively. By Theorem 3.1, we have uv,uv, and u=Mv. Henceu#(v)rev,u#(v)revWPAA.

Furthermore, u#(v)revWP[αα]AM. Thus from (4.5) we find(s,#)Rαfor alls{u#(v)rev,u#(v)rev,u#(v)rev}.

It follows thatwu#vrevuuu#(v)rev(v)rev(v)revRα#.

Thus w#RαWPAARα. As w was arbitrary, and as both u and v contain α, we have proved the inclusion in (4.6). This establishes the desired equality (4.6). □

Thus, to prove (the hard direction of) the main theorem, it suffices to show that the properties of WP[αα]AM reduce to the properties of L(M). This is non-trivial; a reduction to the properties of L(M) is suggested by Theorem 3.1, but L(M) is not a finitely generated monoid, being a free product of L(M) by an infinite rank free monoid F. However, the word problem of F is, informally speaking, essentially the context-free language LF=WPAA(ΣΣ)#(ΣΣ)rev. We will show that WP[αα]AM is, up to technical details, describable as an alternating product of the word problem of L(M) by this language LF, together with an appropriate amount of ancestry.Footnote2 By the preservation results for alternating products and ancestry, this yields the proof strategy for reducing WP[αα]AM to L(M).

Lemma 4.2.

If L(M) has word problem in C, then WP[αα]AMC.

Before giving the proof of this key lemma, we need some setup and auxiliary lemmas. We introduce some notation for convenience. Let Σ1=Σ, and let Σ2=ΣΣ. Obviously Σ1Σ2=. Let Γi=φ(Σi) for i = 1, 2. As Σi is a suffix code, it follows that φ(Σi)=φ(Σi)=Γi, for i = 1, 2. For further ease of notation, we write M1=Γ1L(M)=L(M), and M2=Γ2L(M)=F. Then L(M)=M1M2, where * denotes the monoid free product. In this new notation, it follows directly from Theorem 3.1 that if u,vΣ1, then(4.7) uα=Mvαφ(u)=M1φ(v),(4.7) as in this case φ(u),φ(v)Γ1, and all defining relations of the presentation (3.3) of L(M) are pairs of words over the alphabet Γ1. Analogously, if u,vΣ2, then(4.8) uα=Mvαuv.(4.8)

We define two rewriting systems on Σ1 resp. (Σ1rev). Let(4.9) Iα={(wα,α)|wΣ1+:wα=Mα},(4.9) (4.10) Iαr={((wα)rev,αrev)|wΣ1+:wα=Mα}.(4.10)

Let now u(Σ1Σ2) be any word, factorized uniquely as u0u1un with uiΣX(i)+ for all 0i<n and unΣX(n), where X is a standard parametrisation. We say that (this factorization of) u is reduced if uα or uε; or if uiαMα for all 0in. Obviously, any irreducible descendant of u modulo Iα is reduced, and any reduced word is clearly irreducible modulo Iα. Furthermore, as Iα is M-invariant, if uIαu, then u=Mu, so we conclude that every word u(Σ1Σ2) is equal in M to some reduced word u (though this is generally not unique). Given any reduced u, we can uniquely factorize it as a reduced factorization u0u1uk, where uiΣY(i) for 0ik, where Y is a standard parametrisation. This factorization u0u1uk of u is called the normal form of the reduced word u.

Lemma 4.3.

Let u,v(Σ1Σ2). Let u, resp. v, be any reduced forms of u, resp. v, with normal forms uu0u1um and vv0v1vn respectively. Then uα=Mvα if and only if (1) n = m, and (2) ui,viΣX(i) and uiα=Mviα for all 0in, for some standard parametrisation X.

Proof.

The “if” direction follows by induction on n, and the simple observation that if ui,vi,ui+1,vi+1Σ+, then as ui+1 and vi+1 begin with α we have(uiα=Mviα and ui+1α=Mvi+1α)uiui+1α=Mvivi+1α.

For the “only if” direction, by Theorem 3.1, it follows that uα=Mvα if and only if φ(u)=M1M2φ(v). Now φ(u)φ(u0)φ(u1)φ(um). As φ(ui)=M1M21 if and only uiα=Mα, it follows from the fact that u is reduced that φ(u) is reduced with respect to the monoid free product M1M2, i.e. no non-empty subword of φ(u) equals 1 in M1M2. The analogous statement is true for φ(v). Hence, as φ(u)=M1M2φ(v), it follows from the usual normal form lemma for monoid free products (see [Citation20, Section 8.2] or [Citation37, Section 1]) that (1) n = m; and that (2) φ(ui),φ(vi)ΓX(i) and φ(ui)=MX(i)φ(vi) for 0in, where X is some standard parametrisation. As uiα=Mviα if and only φ(ui)=MX(i)φ(vi), the result follows. □

The rewriting systems Iα and Iαr defined in (4.9) and (4.10) are close to being monadic. We show that, via appropriate rational transductions, we can extend the monadic ancestor property to also apply to these rewriting systems, using the self-overlap free property of α in a non-trivial way.

Lemma 4.4.

If L(M) has word problem in C, then Iα and Iαr are C-ancestry preserving.

Proof.

Let L(Σ1Σ2) be an arbitrary language with LC. To prove that Iα is C-ancestry preserving, it suffices to show that LIαC. Let Iα+ be the set of left-hand sides of rules in Iα. We claim Iα+C. Indeed, wΣ1 is such that wα=Mα if and only if φ(w)=L(M)1 by (4.7). Recalling the definition of IPAM from (2.1), it follows that(4.11) Iα+=φ1(IPΓ1M1)α{α}=φ1(IPΓ1M1)αΣ1+α.(4.11)

As L(M) has word problem in C, it follows that IPΓ1M1C, by an easy transduction (cf. [Citation37, Lemma 3.5]). As C is a full AFL and φ is a homomorphism, it hence follows from (4.11) that Iα+C.

Let now be a new symbol. Let A=A{}, and define the homomorphism σ:AA by aa for all aA, and α. Define a new rewriting systemIα={(W)|Wσ1(Iα+)(AAαA)}.

Clearly Iα is a monadic rewriting system. The language of left-hand sides of in Iα is the intersection σ1(Iα+)(AAαA), and as Iα+C it follows from the closure of C under rational transduction that Iα is a C-rewriting system.

Now, as α is self-overlap free, Σ1 is a suffix code, so for any word uAαA we can uniquely factor u as u0αu1ααuk, where uiAAαA for 0ik. Hence there is exactly one word uA with the properties that (1) u does not contain α; and (2) σ(u)=u. The uniqueness of this word depends on the fact that α is self-overlap free.Footnote3 Of course, for any uAAαA, there is also a unique such u, namely uu. Let L be the language {w|wL}. Then σ(L)=L. Furthermore, by the above uniqueness argument, we have(4.12) L=σ1(L)(AAαA).(4.12)

Hence, from LC we conclude LC.

We now claim that LIα=σ(LIα). The right-hand side is the image under the homomorphism σ of the ancestor of L under the monadic C-rewriting system Iα. As C is a super-AFL, the right-hand side is in C, and thus we would conclude that Iα is C-ancestry preserving. The desired equality is easy to prove. Indeed, it follows directly from the fact that if wA and uL, then wIαu if and only if wIαu. This latter fact is proved by an easy induction on the number of rules applied, and we omit this proof. We conclude that Iα is C-ancestry preserving.

The case of Iαr is symmetric, bearing in mind that (1) C is reversal-closed; (2) αrev is self-overlap free if and only if α is self-overlap free, and (3) Σ1rev is a prefix code, rather than a suffix code (factorization over Σ1rev is still unique). We omit the details. □

One more step is needed before proving Lemma 4.2. Let P2={w#wrev|wΣ2}, where the P abbreviates “palindrome.” By (4.8), we have(4.13) WPAM(Σ2#(Σ2rev))=P2.(4.13)

Clearly P2 is a context-free language, as Σ2 is regular, so P2C. Let τ#:A#A# be defined by τ#(a)=a for all aA, and τ#(#)=α#αrev. We define the (rather complicated-looking) language(4.14) Lα=Iα(τ#(WPΣ1L(M)P2))Iαr.(4.14)

We shall see in the below proof of Lemma 4.2 that Lα=WP[αα]AM. We first prove that Lα is a language encoding the language-theoretic properties of L(M).

Lemma 4.5.

If L(M) has word problem in C, then LαC.

Proof.

If L(M) has word problem in C, we have WPΣ1L(M)C (cf. also the remark preceding (4.1)). This is a concatenation-closed language, as it is a word problem. Furthermore, P2 is a context-free language and is easily checked to be concatenation-closed. Hence by Proposition 2.1, the alternating product WPΣ1L(M)P2 is in C, as is its image under the homomorphism τ#. By Lemma 4.4, Iα and Iαr are C-ancestry preserving, so by Proposition 2.2, we have LαC. □

Proof of Lemma 4.2.

It suffices, by Lemma 4.5, to prove that WP[αα]AM=Lα. We prove the inclusions one at a time.

(). Let wWP[αα]AM be arbitrary. Then we can write wuα#(vα)rev with u,v(Σ1Σ2) and uα=Mvα. Let u,v be reduced forms of u resp. v such that uIαu and vIαv. As Iα is M-invariant, we have u=Mu and v=Mv. Then by Lemma 4.3 we can writeuu0u1umandvv0v1vn,with n = m, ui,viΣX(i) and uiα=Mviα for all 0in=m, for some standard parametrisation X. Let X be such a parametrisation. Let(4.15) wu1u2un#(vn)rev(v2)rev(v1)rev.(4.15)

Now for every 0in, we have uiα=Mviα if and only if φ(ui)=MX(i)φ(vi). When X(i) = 1, then by (4.1) and (4.7) it follows that πΣ1(ui)=πΣ1(vi), so ui#(vi)revWPΣ1L(M). When X(i) = 2, then by (4.8) we have uivi, so ui#(vi)revP2. It follows that the right-hand side of (4.15) is an element of the alternating product of WPΣ1L(M) by P2, and hence so too is w. Let Q=WPΣ1L(M)P2, i.e. we just proved wQ. Let(4.16) wu1u2unα#αrev(vn)rev(v2)rev(v1)revu#(v)rev.(4.16)

Then wτ#(w). We have uIαu, so of course uαIαuα. As vIαv, we have vrevIαr(v)rev and so also αrevvrevIαrαrev(v)rev. Hence, by the definition of (Iα,Iαr)-ancestors, we have(4.17) uα#αrevvrevIα(τ#(Q))Iαr.(4.17)

But uα#αrevvrev is just w; and the right-hand side of (4.17) is Lα, so we are done.

() This argument is similar to the direction (), and differs only in that it uses the M-invariance of Iα. The proof is therefore only sketched here (but is detailed in the author’s Ph.D. thesis [Citation34, Lemma 4.2.13]).

Consider any word u#vLα. Two things must be proved; first, that (1) u,vrevαAAα; and second, that (2) u=Mvrev. For (1), it suffices to note that any word in τ#(WPΣ1L(M)P2) is of the form U#V with U,VrevαAAα. As Iα is M-invariant, any ancestor U of UαAAα will be equal to U in M; hence, in particular, we will also have UαAAα by Theorem 3.1. The analogous argument for V implies that (1) holds.

For (2), since u#vLα, there is some u#vWPΣ1L(M)PΣ2 such thatuIαuα,vIαrvα,as τ#(u#v)(uα)#(αrevv). By the M-invariance of Iα, we have u=Muα, and analogously also v=Mvα. Thus to show that u#v satisfies (2), it suffices to show uα=M(v)revα. But that this is the case can be easily seen by writing out u#v as an element of the alternating product WPΣ1L(M)PΣ2. □

By combining Lemmas 4.1 and 4.2, we conclude that if L(M) has word problem in C, then M has word problem in C. This is the difficult direction of the proof of Theorem 1.1. It remains to prove the converse:

Lemma 4.6.

If M has word problem in C, then L(M) has word problem in C.

Proof.

Indeed, as before let the homomorphism τ#:A#A# be defined by τ#(a)=a for all aA, and τ#(#)=α#αrev. Then it is easy to check, using (4.1), that(4.18) τ#(WPΣ1L(M))=WPAM(Σ1α#αrev(Σ1rev)).(4.18)

Now τ# is an injective homomorphism, as A{α#αrev} is a code, implying that any word in the image of τ# can be decoded by finding all the occurrences of τ#, which must be surrounded by non-overlapping occurrences of α, and the remaining letters are then decoded as themselves. In particular, τ#1°τ#(L)=L for any language L. Hence, applying τ#1 to both sides in (4.18), we find that WPΣ1L(M) is a rational transduction of WPAM. As C is closed under rational transductions, it follows that L(M) has word problem in C. □

This completes the proof of Theorem 1.1.

5 Corollaries of Theorem 1.1

In this section, we present some corollaries of Theorem 1.1. In particular, we will show that it is decidable whether a one-relation monoid containing a non-trivial idempotent has context-free word problem (Theorem 5.4). This answers a generalisation of a question first asked by Zhang in 1992, which was answered affirmatively by the author in [Citation36]. We begin by applying the result to the rational subset membership problem for monoids.

5.1 Rational subset membership problem

Throughout this section we will fix a weakly compressible monoid M, defined by the presentation (3.1), which is compressible with respect to α. As the class of context-free (resp. indexed) languages is a super-AFL closed under reversal, it of course follows from Theorem 1.1 that: M has context-free (resp. indexed) word problem if and only if L(M) does.

One of the direct consequences for a monoid having context-free word problem is decidability of its rational subset membership problem. Recall that for a finitely generated monoid M, generated by a finite set A and with associated surjective homomorphism π:AM, this decision problem asks: given a regular language RA and a word wA, can one decide whether π(w)π(R)? The rational subset membership problem clearly specializes to the submonoid membership problem, the divisibility problems, and the word problem for M. It is easy to check that any context-free monoid has decidable rational subset membership problem; indeed, if π(w)π(R) if and only if w is equal to some word in R, i.e. if and only if wWPAM/#Rrev, where / denotes the right quotient. As #Rrev is a regular language, and WPAM is a context-free language, the quotient is a context-free language; and membership in context-free languages is well-known to be (uniformly) decidable (cf. also [Citation36, Theorem 3.5]). We conclude:

Corollary 5.1.

Let M be a weakly compressible monoid. If L(M) has context-free word problem, then the rational subset membership problem for M is decidable.

Example 4.

Let n,k,β1,,βn be natural numbers such that n > 1, 1kn, and βi1 for all 1in. Let Π=Π(n,k,{βi}i=1n) be the monoid defined byMona1,a2,,an|a1β1a2β2anβn=ak.

Then as the rewriting system with the single rule (a1β1a2β2anβn,ak) is complete, monadic, context-free, and defines Π, it follows that Π has context-free word problem [Citation7, Corollary 3.8]. Hence, by Theorem 1.1 any monoid which compresses to Π has context-free word problem. Let now A be a new alphabet, let αA be self-overlap free, and let wiα(AAαA) for 1in be pairwise distinct words. Let τ be the homomorphism defined by mapping aiwi for all 1in. Let Π=Π(n,k,{βi}i=1n,{wi}i=1n) be the monoid defined by(5.1) Π=MonA|τ(a1β1a2β2anβn)α=τ(ak)α.(5.1)

Then one readily sees that L(Π)=Π. Hence Π has context-free word problem, for any choice of n,k,βi and wi (1in). As a concrete example, if A={x,y} and α=xy, then the monoid Π=Π(3,2,{2,3,4},{xy,xyx,xyy}) defined by(5.2) Monx,y|(xy)2(xyx)3(xyy)4xy=xyxxy(5.2) has context-free word problem, as it compresses to the monoid Π(3,2,{2,3,4}) defined byMona1,a2,a3|a12a23a34=a2.

Verifying directly that (5.2) has context-free word problem appears to be somewhat tedious.

We conclude that the monoid (5.2) has decidable rational subset membership problem, and more generally so too does any monoid of the form (5.1).

5.2 Subspecial one-relation monoids

We turn to the case when M in (3.1) is defined by a single relation u = v. Assume without loss of generality that |u||v|. We say that M is subspecial if uvAAv. Any special monoid (i.e. when the defining relation is u = 1) is obviously subspecial. An element eM is called an idempotent if e2=e. Of course, the identity element 1 is always a (trivial) idempotent. Associated to any idempotent e is a maximal subgroup, being the set of elements which are invertible with respect to this idempotent (alternatively, the group H-class of e). Lallement [Citation24] proved that a one-relation monoid M contains a non-trivial idempotent if and only if (1) M is special, and not right cancellative; (2) |u|>|v|>0 and M is subspecial. By using Adian’s overlap algorithm (see [Citation36]), it is decidable whether a one-relation special monoid is right cancellative. Hence it is decidable whether a one-relation monoid M contains a non-trivial idempotent.

As M is subspecial, it is clearly weakly compressible. Furthermore, it is not hard to see that L(M) is also subspecial (see e.g. [Citation22, Lemma 5.4]). Thus to any subspecial monoid M we can associate a special monoid Ls(M), obtained by iterating weak compression until we arrive at a special monoid. Hence M has context-free word problem if and only if Ls(M) does, by Theorem 1.1 (of course, the same is also true for any reversal-closed super-AFL). Using the structural results about maximal subgroups of subspecial monoids by Gray & Steinberg [Citation14], we can prove the following extension of the Muller–Schupp theorem:

Theorem 5.2.

Let M be a subspecial (one-relation) monoid. Then M has context-free word problem if and only if all of its maximal subgroups are virtually free.

Proof.

First, assume M is special. Then, by a result due to Malheiro [Citation30, Theorem 4.6], all maximal subgroups of M are isomorphic to the group of units U(M) of M. The main theorem of [Citation36] states that M has context-free word problem if and only if U(M) is virtually free. The result thus follows in this case.

Suppose that M is subspecial, but not special. By [Citation14, Lemma 5.2], the maximal subgroups of M are all isomorphic to the group of units of Ls(M), with the exception of the group of units of M, which is trivial (and hence virtually free). Let G be one of the non-trivial maximal subgroups of M, so that GU(Ls(M)). Now Ls(M) is a finitely generated one-relation special monoid. Hence, by classical results (cf. [Citation36]), G is a finitely generated one-relator group.

Assume now that M has context-free word problem. Then, as G is a finitely generated subsemigroup of M, it follows from [Citation17, Proposition 8(b)] that G has context-free word problem. Hence, by the Muller–Schupp theorem, G is virtually free, as desired. For the converse, assume G is virtually free. By the easy direction of the Muller–Schupp theorem, G has context-free word problem. Then, by the main theorem of [Citation36], Ls(M) has context-free word problem. Hence, by the main theorem of the present paper, M has context-free word problem. □

This gives a complete algebraic characterization of the subspecial monoids with context-free word problem, extending the Muller–Schupp theorem to this class.

Corollary 5.3.

Let M be a subspecial one-relation monoid such that all of its maximal subgroups are virtually free. Then M has decidable rational subset membership problem.

For example, one can check that the subspecial monoid(5.3) M3=Monx,y|xyxyyxyx=x(5.3) compresses (in a single step) to Ls(M3)Mona,b,c|abca=1. Now it follows from [Citation2] that the group of units of Ls(M3) is isomorphic to Gpp,q|pqp=1, which is isomorphic to Z by removing the redundant generator q. Hence the non-trivial maximal subgroups of M3 are all infinite cyclic, and in particular virtually free. Hence the monoid M3 defined by (5.3) has context-free word problem by Theorem 5.2, and it has decidable rational subset membership problem by Corollary 5.3.

We end with a final corollary of Theorem 5.2, the statement of which does not use compression or subspeciality.

Theorem 5.4.

There is an algorithm which takes as input a one-relation monoid M=MonA|u=v containing a non-trivial idempotent, and decides whether M has context-free word problem.

Proof.

Suppose we are given a one-relation monoid presentation MonA|u=v for M. Assume without loss of generality that |u||v|. Then as M contains a non-trivial idempotent, either M is special, or else M is subspecial with |u|>|v|0. In the first case, the defining relation is u = 1, and by [Citation36, Theorem B] we can hence decide (uniformly in u) whether this special one-relation monoid has context-free word problem.

In the latter case, we repeatedly compress the relation u = v until we find(5.4) Ls(M)=MonB|w=1.(5.4)

By Theorem 1.1, M has context-free word problem if and only if the special one-relation monoid (5.4) has context-free word problem; this latter problem is (uniformly) decidable by another application of [Citation36, Theorem B]. □

In 1992, Zhang [Citation40, Problem 3] asked if it is decidable whether a special one-relation monoid has context-free word problem. This was recently answered affirmatively by the author [Citation36, Theorem B]. The above Theorem 5.4 thus extends this answer from special to subspecial one-relation monoids.

Acknowledgments

The research in this article was carried out while the author was a Ph.D. student at the University of East Anglia. The author wishes to thank his supervisor Dr Robert D. Gray for many useful comments and discussions, and the anonymous referee for very thorough and careful comments, all of which significantly improved the article.

Additional information

Funding

The author gratefully acknowledges funding from the Dame Kathleen Ollerenshaw Trust, which is supporting his current research at the University of Manchester.

Notes

1 The compression defined by Lallement [24] transforms M into Ln(M) in a single step, whereas that by Adian & Oganesian [3] or indeed Kobayashi [22] corresponds to transforming M into L(M).

2 The alternating products and ancestry defined and used by the author in [37] were, in fact, first developed by the author to deal with precisely the problem of describing the word problem of L(M).

3 For example, if we take the word αxyx, which has self-overlaps, then σ(xy)=σ(yx)=xyxyx.

References

  • Adian, S. I. (1960). The problem of identity in associative systems of a special form. Soviet Math. Dokl. 1:1360–1363.
  • Adian, S. I. (1966). Defining relations and algorithmic problems for groups and semigroups. Proceedings of the Steklov Institute of Mathematics, No. 85.
  • Adian, S. I., Oganesian, G. U. (1978). On problems of equality and divisibility in semigroups with a single defining relation. Math. USSR, Izv. 12:207–212. [Izv. Akad. Nauk SSSR Ser. Mat. 42(2)].
  • Aho, A. V. (1968). Indexed grammars—an extension of context-free grammars. J. Assoc. Comput. Mach. 15: 647–671. DOI: 10.1145/321479.321488.
  • An i¯s i¯mov, A. V. (1971). The group languages. Kibernetika (Kiev) 4:18–24.
  • Berstel, J., Perrin, D. (1985). Theory of Codes, volume 117 of Pure and Applied Mathematics. Orlando, FL: Academic Press, Inc.
  • Book, R. V., Jantzen, M., Wrathall, C. (1982). Monadic Thue systems. Theoret. Comput. Sci. 19(3):231–251. DOI: 10.1016/0304-3975(82)90036-6.
  • Book, R. V., Otto, F. (1993). String-Rewriting Systems. Texts and Monographs in Computer Science. New York: Springer-Verlag.
  • Brough, T., Cain, A. J., Pfeiffer, M. (2019). Context-free word problem semigroups. In: Hofman, P., Skrzypczak, M., eds. Developments in Language Theory, volume 11647 of Lecture Notes in Computer Science. Cham: Springer, pp. 292–305.
  • Campbell, C. M., Robertson, E. F., Ruškuc, N., Thomas, R. M. (1995). Semigroup and group presentations. Bull. London Math. Soc. 27(1):46–50. DOI: 10.1112/blms/27.1.46.
  • Duncan, A., Gilman, R. H. (2004). Word hyperbolic semigroups. Math. Proc. Cambridge Philos. Soc. 136(3): 513–524. DOI: 10.1017/S0305004103007497.
  • Dunwoody, M. J. (1985). The accessibility of finitely presented groups. Invent. Math. 81(3):449–457. DOI: 10.1007/BF01388581.
  • Engelfriet, J. (1985). Hierarchies of hyper-AFLs. J. Comput. System Sci. 30(1):86–115. DOI: 10.1016/0022-0000(85)90006-6.
  • Gray, R. D., Steinberg, B. (2022). A Lyndon’s identity theorem for one-relator monoids. Selecta Math. (N.S.) 28(3):Paper No. 59, 53.
  • Greibach, S. A. (1970). Full AFLs and nested iterated substitution. Inf. Control 16:7–35.
  • Harrison, M. A. (1978). Introduction to Formal Language Theory. Reading, MA: Addison-Wesley Publishing Co.
  • Hoffmann, M., Holt, D. F., Owens, M. D., Thomas, R. M. (2012). Semigroups with a context-free word problem. In: Yen, H.-C., Ibarra, O. H., eds. Developments in Language Theory, volume 7410 of Lecture Notes in Computer Science. Heidelberg: Springer, pp. 97–108.
  • Holt, D. F., Owens, M. D., Thomas, R. M. (2008). Groups and semigroups with a one-counter word problem. J. Aust. Math. Soc. 85(2):197–209. DOI: 10.1017/S1446788708000864.
  • Hopcroft, J. E., Ullman, J. D. (1979). Introduction to Automata Theory, Languages, and Computation. Addison-Wesley Series in Computer Science. Reading, MA: Addison-Wesley Publishing Co.
  • Howie, J. M. (1995). Fundamentals of Semigroup Theory, volume 12 of London Mathematical Society Monographs. New Series. New York: The Clarendon Press, Oxford University Press, Oxford Science Publications.
  • Jantzen, M. (1988). Confluent String Rewriting, volume 14 of EATCS Monographs on Theoretical Computer Science. Berlin: Springer-Verlag.
  • Kobayashi, Y. (2000). Finite homotopy bases of one-relator monoids. J. Algebra 229(2):547–569. DOI: 10.1006/jabr.1999.8251.
  • Král, J. (1970). A modification of a substitution theorem and some necessary and sufficient conditions for sets to be context-free. Math. Systems Theory 4:129–139. DOI: 10.1007/BF01691097.
  • Lallement, G. (1974). On monoids presented by a single relation. J. Algebra 32:370–388. DOI: 10.1016/0021-8693(74)90146-X.
  • Lyndon, R. C., Schupp, P. E. (1977). Combinatorial Group Theory. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 89. Berlin-New York: Springer-Verlag.
  • Magnus, W. (1932). Das Identitätsproblem für Gruppen mit einer definierenden Relation. Math. Ann. 106(1): 295–307. DOI: 10.1007/BF01455888.
  • Magnus, W., Karrass, A., Solitar, D. (1966). Combinatorial Group Theory: Presentations of Groups in Terms of Generators and Relations. New York-London-Sydney: Interscience Publishers [John Wiley & Sons, Inc.].
  • Makanin, G. S. (1966). On the identity problem for finitely presented groups and semigroups. PhD thesis, Steklov Mathematical Institute, Moscow.
  • Makanin, G. S. (1966). On the identity problem in finitely defined semigroups. Dokl. Akad. Nauk SSSR 171: 285–287.
  • Malheiro, A. (2005). Complete rewriting systems for codified submonoids. Int. J. Algebra Comput. 15(2):207–216. DOI: 10.1142/S0218196705002220.
  • Markov, A. A. (1947). On certain insoluble problems concerning matrices. Doklady Akad. Nauk SSSR (N. S.) 57: 539–542.
  • Muller, D. E., Schupp, P. E. (1983). Groups, the theory of ends, and context-free languages. J. Comput. System Sci. 26(3):295–310. DOI: 10.1016/0022-0000(83)90003-X.
  • Nyberg-Brodda, C.-F. (2021). A translation of G. S. Makanin’s 1966 Ph.D. thesis “On the Identity Problem for Finitely Presented Groups and Semigroups”. Available online at arXiv:2102.00745.
  • Nyberg-Brodda, C.-F. (2021). The word problem and combinatorial methods for groups and semigroups. PhD thesis, University of East Anglia, UK.
  • Nyberg-Brodda, C.-F. (2021). The word problem for one-relation monoids: a survey. Semigroup Forum 103(2): 297–355. DOI: 10.1007/s00233-021-10216-8.
  • Nyberg-Brodda, C.-F. (2022). On the word problem for special monoids. Semigroup Forum 105(1):295–327. DOI: 10.1007/s00233-022-10286-2.
  • Nyberg-Brodda, C.-F. (2023). On the word problem for free products of semigroups and monoids. J. Algebra 622:721–741. DOI: 10.1016/j.jalgebra.2023.02.007.
  • Post, E. L. (1947). Recursive unsolvability of a problem of Thue. J. Symbolic Logic 12:1–11. DOI: 10.2307/2267170.
  • Thue, A. (1914). Problem über Veränderungen von Zeichenreihen nach gegebenen Regeln. Christiana Videnskaps-Selskabs Skrifter, I. Math. naturv. Klasse, 10.
  • Zhang, L. (1992). Congruential languages specified by special string-rewriting systems. In: Ito, M., ed. Words, Languages and Combinatorics (Kyoto, 1990). River Edge, NJ: World Scientific Publishing, pp. 551–563.
  • Zhang, L. (1992). On the conjugacy problem for one-relator monoids with elements of finite order. Int. J. Algebra Comput. 2(2):209–220. DOI: 10.1142/S021819679200013X.