Available via license: CC BY 4.0
Content may be subject to copyright.
The Alternation Hierarchy of First-Order Logic on
Words is Decidable
Corentin Barloy
Ruhr-Universität Bochum
Bochum, Germany
Michaël Cadilhac
DePaul University
Chicago, USA
Charles Paperman
Université de Lille
Lille, France
Howard Straubing
Boston College
Boston, USA
Abstract—We show that for any
i > 0
, it is decidable, given a
regular language, whether it is expressible in the
Σi[<]
fragment
of first-order logic
FO[<]
. This settles a question open since 1971.
Our main technical result relies on the notion of polynomial
closure of a class of languages
V
, that is, the set of languages
L0a1L1· · · anLn
where each
ai
is a letter and each
Li
a language
of
V
. We show that if a class
V
of regular languages with some
closure properties (namely, a positive variety) has a decidable
separation problem, then so does its polynomial closure
Pol(V)
.
The resulting algorithm for
Pol(V)
has time complexity that is
exponential in the time complexity for
V
and we propose a natural
conjecture that would lead to a polynomial time blowup instead.
Corollaries include the decidability of half levels of the dot-
depth hierarchy and the group-based concatenation hierarchy.
Index Terms—regular languages, first-order logic, alternation
hierarchy, concatenation hierarchy, profinite theory
I. INTRODUCTION
First-order formulas can be used to express properties of
strings. In this setting, one quantifies over positions in a word
(
∃x, ∀y
), can test whether a position appears before another
(
x < y
), and whether a position contains a given letter (e.g.,
a(x)
or
b(x)
). For instance, over the alphabet
A={a, b, c}
,
the following sentence says that there is a position with an
a
,
such that after it, every position with a
b
is eventually followed
by a position with a c:
(∃x)ha(x)∧(∀y)(y > x ∧b(y)) →(∃z)[z > y ∧c(z)]i.
As a regular expression, the language defined by this sentence
— that is, the set of strings having the stated property — can be
written as
A∗a(a+c+bA∗c)∗
. A key measure of the complexity
of such a formula is the number of times it alternates between
existential and universal quantifications. In this example, the
sentence is logically equivalent to the following sentence in
prefix form:
(∃x)(∀y)(∃z)ha(x)∧(y > x ∧b(y)) →[z > y ∧c(z)]i.
The resulting sentence consists of three blocks of alternating
existential and universal quantifiers — in this instance the
‘blocks’ each consist of a single quantifier — beginning with
an existential quantifier. So the language is defined by a
Σ3[<]
sentence:
Σ
to indicate that it starts with an existential quantifier
block,
3
for the number of quantifier blocks,
<
as this is the
only predicate on positions used beside the letter predicates.
We will also say that the language itself is in Σ3[<].
Can our previous formula be “improved” under that metric?
The astute reader may have realized that the language defined
can be reworded as “the string contains an
a
and ends either
with aor c;” this can be written as:
(∃x)(∃y)(∀z)a(x)∧(a(y)∨c(y)) ∧[¬(y < z)].
The prefix contains only two blocks of alternating quantifiers,
beginning with an existential quantifier: thus the language is
in
Σ2[<]
. We note that this complexity measure is conjectured
to be closely related to the minimal depth of an equivalent
Boolean circuit and that depth is tied to the speed at which the
circuit can be evaluated [32] — this conjecture is known to
hold up to
Σ2[<]
[4]. It is thus of crucial importance to find
what is the minimal number of alternations required to define
a given language.
Such an investigation ought to start with a simpler question:
What are the languages that can be defined by first-order
sentences? This was answered by Schützenberger [29] and
McNaughton and Papert [12] in the 1960’s: they are the so-
called star-free languages, and one can decide, given a regular
language, if it is star-free. This left open the finer-grained
question: can the minimal number of alternations be computed?
We provide a positive answer to that question.
To state our main technical theorem, we will work with an
equivalent formulation of the languages in each
Σi[<]
, that is,
the languages definable by a sentence in prefix form starting
with
∃
and alternating between quantifier types at most
i−1
times. The only quantifier-free sentences are equivalent to
⊤
(always true) and
⊥
(always false), hence
Σ0[<]
consists of
the two languages
∅
and
A∗
. To go higher in the hierarchy,
it turns out that a simple operation suffices: the polynomial
closure. For
V
a class of languages, let
Pol(V)
be the set
of languages that are finite unions of languages of the form
L0a1L1· · · anLn
, where each
ai
is a letter of
A
and each
Li
a language of
V
. We write
Πi[<]
for the dual of
Σi[<]
, that
is, the set of complements of languages in
Σi[<]
. Pin and
Weil [19], building on Perrin and Pin [13], showed that:
Theorem 1. For any i > 0,Σi[<] = Pol(Πi−1[<]).
Our previous example language can be seen to be in
Σ2[<]
using this language-theoretic construction as follows. First,
{ε}
, the language that only contains the empty word, is the
complement of
A∗aA∗∪A∗bA∗∪A∗cA∗
, so it is in
Π1[<]
.
Second, the language of words that contain an
a
and end either
in aor cis A∗a{ε} ∪ A∗aA∗c{ε}, hence it is in Σ2[<].
Beside polynomial closure, the precise statement of our main
theorem requires two more ingredients:
arXiv:2501.14899v1 [cs.FL] 24 Jan 2025
•
Some closure properties of the classes of regular languages
that we study: they have to be “positive varieties of
languages” (formally defined in Section II), which entails
closure under union and intersection, but not complement;
notably, these closure properties are preserved by taking
the dual or the polynomial closure of a class;
•
The problem of
V-
separation for a class of regular
languages
V
: it asks, given a pair
(L, L′)
of regular
languages, whether there is a language
K∈ V
with
L⊆K
and
L′∩K=∅
. Note that this relation is not
symmetric. Naturally, a language
L
is in
V
if, and only
if,
L
is
V-
separable from its complement; thus if we
can decide the separation problem for a class of regular
languages, we can decide whether a given regular language
is a member of the class.
Theorem 2 (Main theorem).Let
V
be a positive variety of
regular languages. If
V-
separation is decidable, then so is
Pol(V)-separation.
Since
Σ0[<]-
separation is trivially decidable and the decid-
ability of
Πi[<]-
separation is equivalent to that of
Σi[<]-
sepa-
ration, Theorem 1 and Theorem 2 readily imply the decidability
of separation and membership for each level
Σi[<]
and
Πi[<]
of the alternation hierarchy of first-order logic, and for their
intersections ∆i[<].
Other applications and complexity
•
If
V
is a class of regular languages with a decidable
separation problem, then we show that separation is
decidable for the levels of the hierarchy built by starting
with
V
and applying Boolean then polynomial closure.
These hierarchies are called concatenation hierarchies,
the most prominent of which we just saw. Another well-
studied example is the group concatenation hierarchy (see,
e.g., [16, 27, 23]) which starts from group languages.
•
For time complexity, we consider that the input regular
languages are provided as morphisms — this can be
exponentially larger than the equivalent automaton. In
that setting, we obtain that if
V-
separation is decidable
in time
c(n)
, then
Pol(V)-
separation is decidable in time
O(c(2O(n)))
. This implies that
Σi[<]-
separation can be
decided in a tower of exponentials of height
i
. We propose
a conjecture that would entail that our algorithm only adds
a polynomial blow-up and, consequently, that
Σi[<]-
sep-
aration and FO[<]-separation are both in PTIME.
Historical background
In 1960, McNaughton [11] raised the question of what
properties of finite-state machines could be expressed in first-
order logic. Schützenberger [29] discovered the solution to one
of the problems raised by McNaughton, by showing that the
regular languages definable by sentences of first-order logic are
precisely those whose syntactic monoids are aperiodic—that
is, contain no nontrivial groups. Since the syntactic monoid
of a regular language can be computed from any of the
usual representations of the language (e.g., an automaton
that recognizes the language, or a regular expression), this
provides an effective test for whether a regular language
is first-order definable. Schützenberger wrote in terms of
operations on languages (Boolean operations and concatenation)
rather than first-order sentences; thus the class of languages
definable in this way became known as the star-free languages,
since they can be defined by extended regular expressions
in which complementation is allowed along with union and
concatenation, but in which the star operation is not used.
The book by McNaughton and Papert [12] gave an integrated
presentation in terms of logic, extended regular expressions,
automata and finite monoids.
Cohen and Brzozowski [7] introduced a complexity measure
on the star-free languages, which they called dot-depth: The
0-level of the hierarchy consists of the finite and cofinite
languages over a finite alphabet
A.
If
k≥0
then the
k+1
2
level consists of products
L1· · · Lr,
where
r≥1
, and each
Li
is a language in level
k,
and the
k+ 1
level is the Boolean
closure of the level
k+1
2.
Cohen and Brzozowski posed the
problem of effectively determining whether a given star-free
language belongs to a particular level of the hierarchy. In the
present paper, we settle this problem for the
k+1
2
levels. The
question for the Boolean-closed integer levels remains open.
Brzozowski and Knast [6] proved that the hierarchy is
strict. (See also Thomas [35] who gave a proof of this fact
using model-theoretic games.) Straubing [33] and Thérien [34]
studied a slightly different hierarchy in the star-free languages
over a given alphabet
A
. Set
V0={∅, A∗}
, then, inductively:
BVi=Boolean closure of Vi,and Vi+1 = Pol(BVi).
The class of star-free languages is then the union of the
Vi.
The class
Vi
is precisely the class of
Σi[<]
languages that we
consider in the present paper [19, 13]. This is closely related
to the original formulation of the dot-depth hierarchy: From an
algebraic perspective it is the simpler and more fundamental of
the two. A solution of the membership problem for any of the
levels of this hierarchy implies one for the corresponding level
of the dot-depth hierarchy [33]. The membership problem for
V1
is fairly simple. It was solved for
BV1
in a famous theorem
of Simon [31], for
V2
by Arfi [3], and, with a simpler proof,
by Pin and Weil [19]. More recently, in a major breakthrough,
Place and Zeitoun [22, 25] found solutions for
BV2,V3
and
V4.
Pin also contributed a thorough historical account in 2017,
covering more than 45 years of research on that topic [18].
Here, we provide a solution for all Vi,i≥0.
Organization of the paper
In Section II, we introduce all the concepts required to
go through our proof. We strive to keep the mathematical
objects as simple as possible, sometimes at the cost of spelling
out arguments that would have been more elegantly proved
using more powerful topological or algebraic tools. Sections III
to IX present the proof of our main theorem, starting with
an overview. In Section X, we provide some corollaries to
our main theorem. In Section XI, we close with some open
questions. All missing proofs appear in the appendix.
II. PRELIMINARIES
We assume familiarity with basic automata theory, including
the definitions of alphabet, word, language, and regular
language. Every statement singled out as a Lemma, Theorem,
or Corollary in this section has an elementary proof that is
provided in appendix.
A. Sets, languages
We will manipulate partially ordered sets
(M, ≤)
and, with
E⊆M
, we will write
↑E
for the upset
{y|(∃x∈E)[x≤y]}
,
and ↓Efor the downset {y|(∃x∈E)[y≤x]}.
We use
A, B
for finite alphabets and write
A∗
for the set
of words over
A
. We write
ε
for the empty word and
A+
for
A∗\ {ε}
. Classes of regular languages are usually written
V
or W.
A language
K
separates a language
L
from another language
L′
if
L⊆K
and
L′∩K=∅
. With
V
a class of regular
languages, a language
L
is
V-
separable from
L′
if there is
K∈ V
that separates
L
from
L′
. We say that
V-
separation
is decidable if there is an algorithm that decides, given two
regular languages L, L′, whether Lis V-separable from L′.
For
V
a class of regular languages, a
V-
polynomial is a finite
union of languages of the form
L0a1L1· · · anLn
, where each
ai
is a letter of
A
and each
Li
a language of
V
. The polynomial
closure of V, written Pol(V), is the set of all V-polynomials.
B. Algebraic structures
In the algebraic approach to regular languages, languages live
alongside algebraic objects. Monoids and semigroups describe
equivalence classes of words within a language, similarly to
the way in which an automaton considers two words to be
equivalent if they induce the same map from states to states.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Semigroups, monoids, morphisms.
Asemigroup is a set
equipped with an associative product. For a subset
E⊆M
,
we write
⟨E⟩
for the subsemigroup of
M
generated by
E
,
i.e., the closure of
E
under the product of
M
. A monoid is
a semigroup that has a neutral element; for a monoid
M
, we
write that element 1M.
A semigroup element
e
is idempotent if
e2=e
. If
M
is a
finite semigroup, for any m∈Mthere is a value r > 0such
that
mr
is idempotent, and this is the unique idempotent among
the positive powers of
m
(this is reminiscent of the notion of
order in finite groups). We write
mω
for the idempotent power
of m.
Throughout this article, all semigroups and monoids are
finite, with the exceptions of the free monoid
A∗
over a finite
alphabet
A
and the monoid
c
A∗
, which will be defined in
Section II-C.
Asemigroup morphism is a function
µ:M→N
with
M, N
semigroups such that
µ(mm′) = µ(m)µ(m′)
for every
m, m′∈M
. It is a monoid morphism if additionally
M, N
are
monoids and µ(1M) = 1N.
A language
L⊆A∗
is recognized by a monoid
M
if there
is a morphism
µ:A∗→M
and a subset
E⊆M
such that
L=µ−1(E)
. A language is regular if, and only if, it is
recognized by some finite monoid. This is effective: Given a
finite automaton, one can compute a morphism
µ:A∗→M
and a set
E
such that
µ−1(E)
is the language of the automaton.
However, the monoid
M
may be exponentially larger than the
automaton: we stress that all our complexity results assume
that the morphism (and hence the monoid) is provided as input.
We will often view a monoid
M
as an alphabet and consider
words in
M∗
. As a prime example, we will rely on the surjective
morphism
πM:M∗→M
, for any monoid
M
(or
πM:M+→
M
when
M
is a semigroup): this is the canonical morphism
of
M
, which simply evaluates the product in
M
. For instance,
if
m1m2
is a 2-letter word in
M∗
and the product of
m1
and
m2in Mis m, then πM(m1m2) = m.
An ordered semigroup is a semigroup
M
equipped with a
partial order ≤that is compatible with the product: for every
m, m′, n ∈M
such that
m≤m′
, we have that
mn ≤m′n
and
nm ≤nm′
. An ordered monoid is an ordered semigroup that
is a monoid. A morphism
µ:M→N
of ordered semigroups
or monoids is a morphism in the sense defined above, with the
additional property
m≤m′
implies
µ(m)≤µ(m′)
for every
m, m′∈M.
We say that an ordered monoid
M
recognizes a language
L⊆A∗
if there is a morphism
µ:A∗→M
and a subset
E⊆Msuch that L=µ−1(↑E).
Remark 3 (Recognition by ordered vs. unordered monoids).
Note that this notion of recognition is a stronger requirement
than in the unordered case: For example, if
A={a, b}
and
M
is the ordered monoid
{0,1}
with the ordering
0≤1
, then
M
recognizes
A∗aA∗
in the unordered sense by the morphism
mapping
a
to
0
and
b
to
1
, and taking
E={0}
. However
E
is not upwardly closed, so this
M
does not recognize this
language in the ordered sense. Observe that the monoid
{0,1}
does recognize
A∗aA∗
in the ordered sense if we change the
order to
1≤0
, or to the trivial order in which 0 and 1 are
incomparable. Note additionally that every finite monoid can
be seen as an ordered monoid with the trivial order
m≤m′
if,
and only if,
m
and
m′
are equal. This shows that a language
is regular if, and only if, it is recognized by some ordered
monoid.
Finally, we introduce a generalization of morphisms (due
to Tilson, in Chapter XII of Eilenberg [10]). A relational
morphism
τ:M△
−→ N
between two semigroups
M, N
is a
function from
M
to
2N
that satisfies, for every
m, m′∈M
,
τ(m)=∅
and
τ(m)τ(m′)⊆τ(mm′)
— the product of two
elements
E, F ∈2N
is defined as
{ef |e∈E∧f∈F}
. For
monoids, we require additionally that 1N∈τ(1M).
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
Varieties.
A class of regular languages is a positive variety
1
if it is closed under union, intersection, inverse morphisms
(from morphisms
µ:A∗→B∗
), and quotient (the operation
that maps a language
L
and a word
u
to
{v|uv ∈L}
, and
1
Formally, a variety maps alphabets to sets of languages over that alphabet.
As is standard, we elect to sidestep the resulting notational burden by
considering varieties themselves to be sets of languages. This has no impact
on our results and their proofs.
symmetrically, the operation that maps to
{v|vu ∈L}
).
We note that if
V
is a positive variety, then so are its dual
{L|L∈ V} and its polynomial closure Pol(V)[15].
A class of ordered finite monoids is a variety (of ordered
monoids) if it is closed under taking ordered submonoids (i.e.,
subsets of a monoid closed under the product), quotients (i.e.,
images of ordered morphisms) and finite direct products.
We denote varieties of ordered monoids in boldface (
V
)
and the set of languages they recognize in calligraphic style
(
V
). The correspondence
V→ V
is a bijective correspondence
between varieties of ordered monoids and positive varieties
of regular languages; this theorem is due to Eilenberg [10]
in the unordered case and to Pin [14] for varieties of ordered
monoids. Accordingly, we will write
Pol(V)
for the variety of
ordered monoids that corresponds to
Pol(V)
. We will recall in
Theorem 8 how
Pol(V)
can be defined directly from
V
; for this,
we require the formalism of profinite words and inequalities.
C. Profinite words and inequalities
Profinite words allow us to speak, in a succinct way, about
infinite sequences of words that are ultimately indistinguishable
by any fixed monoid. Here we give a self-contained sketch of
the theory that does not require any background in topology,
and that is sufficient for our purposes in this paper. For a more
thorough introduction, see the survey by Pin [17].
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
Convergence.
We say that an infinite sequence
(mi)i>0
of
elements of a finite monoid
M
is convergent if the sequence
is ultimately constant; that is, there is some
m∈M
such
that
mi=m
for all sufficiently large
i
. We write
m=
limimi
. The termwise product
(mini)i>0
of two convergent
sequences
(mi)i>0
and
(ni)i>0
is itself convergent, and clearly
limimini= (limimi)(limini).
We say that an infinite sequence
u= (wi)i>0
of words in
A∗
is convergent if for every morphism
µ:A∗→M
, where
M
is a finite monoid, the sequence
(µ(wi))i>0
is convergent.
Convergent sequences embody the idea of “a sequence of words
ultimately indistinguishable by finite monoids.”
An obvious example of a convergent sequence of words
is any ultimately constant sequence. A more interesting (and
important) example is the sequence
{wi!}i>0,
where
w∈A∗
.
If
µ:A∗→M
is any morphism into a finite monoid, and
i > 0
is large enough, then
µ(wi!) = µ(w)i!=µ(w)ω
, the unique
idempotent power of µ(w), hence limiµ(wi!) = µ(w)ω.
We say that two convergent sequences
u= (u1, u2, . . .)
and
u′= (u′
1, u′
2, . . .)
are equivalent if
limiµ(ui) = limiµ(u′
i)
for
all morphisms µfrom A∗into a finite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
The free profinite monoid.
What does a convergent sequence
of words converge to? Let us equip the set of infinite sequences
of words with an internal operation: termwise concatenation,
i.e., with
u= (u1, u2, . . .)
and
v= (v1, v2, . . .)
, the concatena-
tion
uv
is
(u1v1, u2v2, . . .)
. Clearly,
uv
is convergent if
u
and
v
are. This turns the set of convergent sequences into a monoid
˜
A∗
. If
µ:A∗→M
is a morphism into a finite monoid, then
setting, for every
u∈˜
A∗
,
˜µ(u) = limiµ(ui)
gives a morphism
from ˜
A∗into M.
Equivalence of convergent sequences is a congruence on
˜
A∗
.
The quotient of
˜
A∗
by the equivalence relation is called the
free profinite monoid of
A
and is denoted
c
A∗
. Its elements,
that is, the equivalence classes of convergent sequences, are
called profinite words. If
µ:A∗→M
is a morphism into a
finite monoid, then the morphism
˜µ
maps equivalent words to
the same value, and thus we obtain a morphism
bµ:c
A∗→M.
Thus if
u= (u1, u2, . . .)
is a convergent sequence of words,
we can define
limiui
to be the equivalence class of
u
. This
gives us the following continuity property: For every morphism
µ:A∗→M, where Mis finite, bµ(limiui) = limibµ(ui).
Everything that we do with profinite words will depend
on these morphisms
bµ
. This allows us to blur the distinction
between convergent sequences and their equivalence classes.
In particular, we will often say that a convergent sequence of
words is a profinite word.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
Some profinite words.
The monoid
A∗
embeds in
c
A∗
by
mapping each
w∈A∗
to the sequence
(w, w, . . .)
. (For any
two distinct words
w, w′
, there is a morphism into a finite
monoid that maps
w
and
w′
to different elements, so this map
is indeed injective.) As we discussed above, if
u∈A∗
and
µ:A∗→M
is a morphism into a finite monoid, then the
sequence
v= (ui!)i>0
is convergent, and maps under
bµ
to
µ(u)ω
. We thus denote
v
by
uω
, which gives us the equation
bµ(uω) = µ(u)ω.
This implies that
uω
itself is idempotent:
uω=uωuω. Abusing notation, we write uω+1 for uuω.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
Convergent sequences of profinite words.
The existence of
bµ
for a morphism
µ
naturally leads to extending the notion of
convergence to infinite sequences of profinite words. We do not
need to introduce new elements into our algebra to capture the
limits of such sequences; any convergent sequence of profinite
words behaves the same, with respect to morphisms, as a single
profinite word. Let us make this more precise. A sequence
(ui)i>0
of profinite words is convergent if for every morphism
µ:A∗→M
, with
M
a finite monoid,
(bµ(ui))i>0
is ultimately
constant. We then have
Lemma 4. For every convergent sequence
(ui)i>0
of profinite
words over
A
, there exists a unique
w∈c
A∗
such that for
every morphism
µ:A∗→M
with
M
finite, we have
bµ(w) =
limibµ(ui).
We can then write
v= limiui
. In particular, we can extend
the definition of
uω
given above to the case where
u
is a
profinite word. We will repeatedly rely on two additional
properties of convergence in
c
A∗
, given in the next two lemmas.
The first allows us to extend morphisms from
A∗
into a
free profinite monoid (over a possibly different alphabet) to
continuous morphisms on c
A∗:
Lemma 5. Let
A, B
be finite alphabets, and let
µ:A∗→b
B∗
be a morphism. There is a unique morphism
bµ:c
A∗→c
B∗
such that
bµ(w) = µ(w)
for every
w∈A∗
, and such that
bµ
is continuous; that is, if
(ui)i>0
is a convergent sequence of
profinite words over
A
, then
(bµ(ui))i>0
is convergent, and
limibµ(ui) = bµ(limiui).
The second is a sequential compactness property for se-
quences of profinite words:
Lemma 6. Every sequence of elements of
c
A∗
has a convergent
subsequence.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
Inequalities.
An ordered monoid
M
satisfies the profinite
inequality
u≤v
, for profinite words
u, v
, if for each morphism
µ:A∗→M
, one has
bµ(u)≤bµ(v)
. In particular, if a monoid
satisfies a given inequality, so do all of its quotients and
submonoids. For a set
S
of profinite inequalities, we write
JSK
for the set of ordered monoids that satisfy all the inequalities
of S.
Reiterman’s theorem (in the ordered setting) asserts that a
class
V
of ordered monoids is a variety if, and only if, it can
be defined by a set of profinite inequalities [21]. This latter set
is precisely the set of profinite inequalities that are satisfied
by all the monoids in
V
. Let us write
u≤Vv
to indicate that
every monoid of
V
satisfies
u≤v
. Another way of wording
Reiterman’s theorem is thus to say that
M∈V
if, and only if,
for every inequality
u≤Vv
and every morphism
µ:A∗→M
,
one has
bµ(u)≤bµ(v)
. Note that
≤V
is transitive, since the
order of any ordered monoid is.
We make use of a few simple properties of the relation
≤V
:
Lemma 7. Let
V
be a variety of ordered monoids, and let
A, B be finite alphabets.
(a)
(Multiplicativity) If
u1, u2, v1, v2∈b
A∗,
with
ui≤Vvi
for
i= 1,2,then u1u2≤Vv1v2.
(b)
(Continuity) If
(ui)i>0
,
(vi)i>0
are convergent sequences
of profinite words in
b
A∗,
with
ui≤Vvi
for all
i > 0,
then
limiui≤Vlimivi.
(c)
(Preservation under continuous morphisms) If
µ:A∗→
B∗
is a morphism and
u, v ∈c
A∗
with
u≤Vv,
then
bµ(u)≤bµ(v).
We are now equipped to express the polynomial closure in
purely algebraic terms. A self-contained proof is in appendix.
Theorem 8 ([19]).Let
V
be a variety of ordered monoids.
The following holds:
Pol(V) = Juω+1 ≤uωvuω,for any u≤VvK.
Finally, we note that if
V⊆W
are two varieties, then
u≤Wv
implies
u≤Vv
: an inequality that holds in a variety
holds in any subvariety.
D. V-pairs
Consider a variety of ordered monoids
V
and a finite monoid
M. The set of V-pairs of Mis:
{(dπM(u),dπM(v)) |u, v ∈d
M∗∧u≤Vv}.
Note that this set does not depend on whether
M
is ordered
or what its order is. Crucially, this set would be the same if
we were to use, instead of
πM
, any other surjective morphism:
Lemma 9. For any surjective morphism
µ:A∗→M
, the set
{(bµ(u),bµ(v)) |u, v ∈c
A∗∧u≤Vv}
is the set of
V-
pairs of
M
. If
µ
is not surjective, then it is a
subset of all the V-pairs.
When considering a
V-
pair
(m, m′)
, we will call witnesses
a pair of profinite words
u≤Vv
with
m=dπM(u)
and
m′=dπM(v)
. Witnesses of
V-
pairs always exist. A perhaps
surprising fact about
V-
pairs is that, because they “collapse”
infinite properties of profinite words into a finite monoid, they
are not transitive: if
(m, m′)
and
(m′, m′′)
are
V-
pairs, it does
not imply that
(m, m′′)
is a
V-
pair. Specifically,
(m, m′)
may
be witnessed by
u≤Vv
, and
(m′, m′′)
by
u′≤Vv′
, and
there is no reason why
v
and
u′
would be equal, so we cannot
deduce
u≤Vv′
, which would indeed imply that
(m, m′′)
is
aV-pair.
E. The semantics of V-pairs: separation
When
M
is ordered, our last rewording of Reiterman’s
theorem provides a salient semantics for the compatibility
of pairs and the order:
M∈V
if, and only if, for any
V-
pair
(m, n)
of
M
,
m≤n
. In general, even when
M
is not
ordered,
V-
pairs offer a fine-grained characterization of how
V-
languages can separate preimages of elements of
M
. We
include an elementary proof both for completeness and as a
warm up to profinite methods:
Lemma 10 ([1] for the unordered case).Let
V
be a variety of
ordered monoids. Let
µ:A∗→M
be a surjective morphism
and
m, n ∈M
, for a monoid
M
. The pair
(m, m′)
is a
V-
pair
of
M
if, and only if,
Lm=µ−1(m)
is not
V-
separable from
Lm′=µ−1(m′).
Proof. Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Left-to-right direction.
Let
u, v ∈c
A∗
such that
u≤V
v
and
bµ(u) = m, bµ(v) = m′
— they exist by Lemma 9. Let
K∈ V
such that
Lm⊆K
, we show that
K
contains a word
of Lm′, so that no language of Vseparates Lmfrom Lm′.
By hypothesis on
K
, there is a surjective morphism
ν:A∗→
N
, with
N∈V
, and
E⊆N
such that
K=ν−1(↑E)
. Let us
write
u, v
as convergent sequences
(ui)i>0,(vi)i>0
. For any
n
large enough,
µ(un) = m
, so that
un∈Lm⊆K
. This
means that
ν(un)∈ ↑E
and, ultimately, that
bν(u)∈ ↑E
. Since
N∈V
, and
u≤Vv
, this entails that
bν(v)∈ ↑E
. However,
for any
n
large enough, we have both that
µ(vn) = m′
, so
that vn∈Lm′, and bν(v) = ν(vn), so that vn∈K∩Lm′.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Right-to-left direction.
Assume
Lm
is not
V-
separable from
Lm′
. We need to show that there are
u, v ∈c
A∗
such that
bµ(u) = m, bµ(v) = m′
, and
u≤Vv
. To do so, we will look
at all the possible ways a language that contains
Lm
can be
recognized by a monoid in
V
and use these to construct a
convergent sequence of words from
Lm
and a convergent
sequence of words from
Lm′
. The limits of these two sequence
will be the profinite words u, v that we require.
Let
ν1, ν2, . . .
be an enumeration of all morphisms from
A∗
to a monoid of
V
and write
νi:A∗→Mi
. Construct a family
(µi)i>0
of morphisms, with
µi:A∗→M1×· · ·×Mi
, by setting
µ1=ν1
and, for
i > 1
, letting
µi(w)=(µi−1(w), νi(w))
for
any
w∈A∗
. Note
M1× · · · × Mi∈ V
and that its order is
the product of the orders on
Mi
, so that if
µi(w)≤µi(w′)
,
then for any j≤i,νj(w)≤νj(w′).
Since
Lm
is not
V-
separable from
Lm′
, for every
i
the set
↑µi(Lm)∩µi(Lm′)
is nonempty; let
ui∈Lm, vi∈Lm′
be
such that
µi(ui)≤µi(vi)
. Let
u
be a convergent subsequence
of
(ui)i>0
and similarly
v
a convergent subsequence of
(vi)i>0
,
such that the indices taken in the two sequences are the same
for
u
and
v
(we are allowed to do that as a consequence of
compactness). We clearly have that
bµ(u) = m
and
bµ(v) = m′
.
We now show that
u≤Vv
by going back to the definition.
Consider any morphism from
A∗
to a monoid in
V
: it is
νi
for some
i
. There is a
j > i
such that
uj
appears in the
sequence
u
,
vj
appears in the sequence
v
, and
bνi(u) = νi(uj)
and
bνi(v) = νi(vj)
. Since
µj(uj)≤µj(vj)
and
i < j
, we
have that
νi(uj)≤νi(vj)
; consequently,
bνi(u)≤bνi(v)
. This
concludes the proof.
We say that
V-
pairs are computable if there is an algorithm
which, given an arbitrary monoid
M
, computes its
V-
pairs.
The previous lemma implies:
Corollary 11. Let
V
be a variety of ordered monoids. The
problem of
V-
separation is decidable if, and only if,
V-
pairs
are computable. Moreover, the two problems are polytime-
reducible to one another.
Remark 12. Recall that our main goal is to show that if
V-
separation is decidable, then so is
Pol(V)-
separation. A
simpler property, already used by [25], is the following: if
V-
separation is decidable, then so is
Pol(V)-
membership. This
is easily shown:
Pol(V)-
membership can be decided by testing
that the inequalities of Theorem 8 hold, that is, for a monoid
M
,
testing
dπM(uω+1)≤dπM(uωyxω)
for all
u≤Vv
. Equivalently,
this means that
mmω≤mωm′mω
for all
V-
pairs
(m, m′)
of
M
. Since the decidability of
V-
separation implies that
V-
pairs
are computable, this last property can be tested.
III. MAIN THEOREM AND STRUCTURE OF THE PROOF
In light of Corollary 11, our main theorem can be equiva-
lently worded as the following, which is the statement we will
prove:
Theorem 13 (Main theorem, reformulated).Let
V
be a variety
of ordered monoids and assume that
V-
pairs are computable.
There is an algorithm that computes the
Pol(V)-
pairs of any
given monoid.
We now sketch the proof structure, which is split in several
sections. Let
M
be a monoid,
V
a variety of ordered monoids,
and assume that
V-
pairs are computable; we provide an
algorithm that computes the Pol(V)-pairs of M.
•
In Section IV, we introduce the notion of relational
expansion of
M
. This is a finite semigroup that contains
M
and some extra information. We can extract from a
relational expansion, by projection, a subset of
M×M
that we call the graph of the expansion.
•
In the same Section IV, we introduce two types of rela-
tional expansions: A relational expansion is Pol(V)-cor-
rect if its graph contains only
Pol(V)-
pairs of
M
,
Pol(V)-
complete if it contains all
Pol(V)-
pairs of
M
. We
also introduce a property with a more technical definition:
a relational expansion is
Pol(V)-
stable if it satisfies a
closure condition related to the inequalities in Theorem 8.
•
In Section V, we show that if a relational expansion
is
Pol(V)-
stable, then it is
Pol(V)-
complete. To com-
pute
Pol(V)-
pairs, it is thus sufficient to construct a
Pol(V)-
stable,
Pol(V)-
correct relational expansion of
M
:
its graph is exactly the set of
Pol(V)-
pairs of
M
. This is
what the algorithm for Theorem 13 does.
•
In Section VI, we provide an infinite family of relational
expansions
M0, M1, M2, . . .
of
M
. These expansions are
computable provided that
V-
pairs are computable. The
algorithm for Theorem 13 will compute each of these
expansions, one by one.
•
In Section VII, we show that this sequence is eventually
constant. We call
M∞
this final relational expansion and
note that there is a computable property that asserts that
Mi
is
M∞
: the algorithm for Theorem 13 will then stop
computing
Mi
’s. We show in that section that
M∞
is
Pol(V)-stable, implying that it is Pol(V)-complete.
•
In Section VIII, we show that
M∞
is
Pol(V)-
correct.
Its graph is thus the set of
Pol(V)-
pairs of
M
, and the
algorithm for Theorem 13 can return that set.
•
In Section IX, we recapitulate this process by explicitly
providing the algorithm for Theorem 13.
IV. COR RE CT,COMPLETE,AND STABL E RE LATIONAL
EXPANSIONS
Definition 14. Let
M, N
be finite monoids,
η:N→M
be
a morphism, and
ρ:N△
−→ M
be a relational morphism. The
monoid
N
together with the pair
(η, ρ)
is relational expansion
of
M
if
η
is a surjective morphism and for any
x∈N
,
η(x)∈ρ(x)
. We usually say that
N
itself is a relational
expansion, with
(η, ρ)
implicitly defined, and, conversely, we
may speak of the relational expansion induced by (η, ρ).
The graph of Nis the subset of M×Mdefined by:
[
x∈N
{(η(x), m)|m∈ρ(x)}.
Let
V
be a variety of ordered semigroups. The relational
expansion Nis:
•Pol(V)-
correct if its graph contains only
Pol(V)-
pairs
of M;
•Pol(V)-
complete if its graph contains all
Pol(V)-
pairs
of M;
•Pol(V)-
stable if for any
V-
pair
(x, y)
of
N
with
x2=x
and any α, β ∈ρ(x), we have that αη(y)β∈ρ(x).
We will also use these definitions for semigroup relational
expansions, where the morphisms are adjusted to be semigroup
morphisms.
Remark 15 (On the definition of stable).Our main goal
is to construct, given a monoid
M
, a
Pol(V)-
correct and
Pol(V)-
complete relational expansion of
M
. We will show in
the next section that stability implies completeness; we first
sketch the intuition behind the definition of stability.
The definition of
Pol(V)-
stable is designed to mimic the
inequalities of Theorem 8. Let us make that more concrete.
Assume we wish to construct a relational expansion of a monoid
M
such that its graph is exactly the set of
Pol(V)-
pairs.
Theorem 8 provides a set of inequalities that generate
Pol(V)
,
so we can build our relational expansion by including those.
Specifically, consider the relational expansion of
M
induced
by letting
η:M→M
be the identity, and
ρ:M△
−→ M
be
defined by setting
ρ(m)
to
{m}
if
m2=m
and, otherwise, to:
{mm′m|(m, m′)aV-pair of M}.
Disregarding the fact that
ρ
may not be a relational morphism,
this relational expansion is
Pol(V)-
correct thanks to Theo-
rem 8. Indeed, let
u≤Vv
witness the
V-
pair
(m, m′)
with
m2=m
. Theorem 8 asserts that
uω+1 ≤Pol(V)uωvuω
, and
projecting each side of this inequality with
dπM
, we conclude
that (m, mm′m)is a Pol(V)-pair of M.
So what are the
Pol(V)-
pairs that are missing for this
expansion to be
Pol(V)-
complete? First, we may need some
extra closure properties to ensure that
ρ
is indeed a relational
morphism (a property that is crucially needed in the forthcom-
ing Lemma 18) — this is not hard to ensure. Second, and much
more importantly, the operation
JSK
, as used in Theorem 8,
induces a closure under transitivity: recall that
u≤Wv
and
v≤Ww
implies
u≤Ww
. Some profinite inequalities induced
by this closure do not appear as
uω+1 ≤uωvuω
. This is
why the definition of stability is not simply requiring that
η(x)η(y)η(x)∈ρ(x)
, but
αη(y)β∈ρ(x)
, for
α, β ∈ρ(x)
.
If the graph of the relational expansion is seen as “the
Pol(V)-
pairs already computed,” then this simple modification
ensures that a step of transitivity is applied. We encourage
the reader to keep this subtlety in mind while going through
the proof that
Pol(V)-
stability entails
Pol(V)-
completeness
(Lemma 18). We will further discuss how
Pol(V)-
stability
and
Pol(V)-
correctness conspire to make finding a relational
expansion with both properties difficult in Section VI-A.
V. Pol(V)-STABLE I MP LI ES Pol(V)-COMPLETE
To show the claimed result, we will rely on the following
lemma which appears in [2, Thm. 3.1 & Cor. 3.2] (a self-
contained proof is in the appendix):
Lemma 16. Let
W
be a variety of ordered monoids such that
Pol(W) = W
. Let
u, v ∈c
A∗
with
u≤Wv
. Suppose that there
are
u′, u′′ ∈c
A∗
and
ℓ∈A∪ {ε}
such that
u=u′ℓu′′
. Then
there exist
v′, v′′ ∈c
A∗
such that
v=v′ℓv′′
with
u′≤Wv′
and u′′ ≤Wv′′.
We will also rely on a profinite version of the classical
factorization forest theorem of Simon:
Theorem 17. Let
µ:A∗→M
be a morphism with
M
a finite
monoid. There is a function
d:c
A∗→ {0,1,...,3|M|}
called
the decomposition depth such that for any u∈c
A∗:
•If d(u)=0, then u=εor u∈A;
•
If
d(u)>0
, then there are words
u1, u2∈c
A∗
with
d(u1), d(u2)< d(u)
such that one of the following holds:
–u=u1u2
, in which case we say that
u
has binary type;
–u=u1u′u2
for some
u′∈c
A∗
with
bµ(u1) = bµ(u′) =
bµ(u2) = bµ(u)
an idempotent of
M
, in which case we
say that uhas idempotent type.
We are now ready to show the main result of this section:
Lemma 18. Any relational expansion that is
Pol(V)-
stable is
also Pol(V)-complete.
Proof.
Let
N
be a relational expansion of some monoid
M
,
and let
(η, ρ)
be the morphisms that induce
N
. Recall that
η
is a surjective morphism from
N
to
M
and
ρ
is a relational
morphism from
N
to
M
, such that for any
x∈N
,
η(x)∈ρ(x)
.
We show that for any
u≤Pol(V)v
with
u, v ∈c
N∗
, we
have that
(bη(u),bη(v))
is in the graph of
N
(recall that
η
is
surjective, so that any
Pol(V)-
pair of
M
can be obtained in
that fashion). To do so, we show that
bη(v)∈bρ(u)
. This is
shown by induction on the decomposition depth of
u
with
respect to πN, as defined in Theorem 17.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
If uhas decomposition depth 0.
In that case,
u
is either
the empty word or a single letter.
•
If
u=ε
, then
cπN(u)=1N
. Since
u≤Pol(V)v
implies
that
u≤Vv
, the pair
(1N,cπN(v))
is a
V-
pair of
N
. In
addition, since
N
is a relational expansion of
M
, we get
that
1M∈ρ(1N)
and
Pol(V)-
stability then assert that
ρ(1N)
contains
1Mbη(v)1M=bη(v)
. Since
ρ(1N) = bρ(u)
,
we are done.
•
If
u=x∈N
, then Lemma 16, seeing
x
as
εxε
, provides
ε≤Pol(V)v1, v2
with
v=v1xv2
. The previous case
(
u=ε
) shows that
bη(vb)∈bρ(ε)
,
b= 1,2
. Additionally,
since
N
is a relational expansion of
M
, we have that
bη(x)∈bρ(x), and thus:
bη(v) = bη(v1)bη(x)bη(v2)∈bρ(ε)bρ(x)bρ(ε)⊆bρ(x).
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
Inductive step: Case uhas binary type.
This implies that
u=u1u2
with
ub
,
b= 1,2
, of strictly smaller decomposition
depth. Lemma 16 provides us with a rewriting of
v
as
v1v2
with
ub≤Pol(V)vb
,
b= 1,2
. The induction hypothesis then
asserts that
bη(vb)∈bρ(ub)
, for
b= 1,2
. We obtain, as desired:
bη(v) = bη(v1v2) = bη(v1)bη(v2)∈
bρ(u1)bρ(u2)⊆bρ(u1u2) = bρ(u).
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
Inductive step: Case uhas idempotent type.
This implies
that
u=u1u′u2
with
u1, u2
of strictly smaller decomposition
depth and
cπN(u) = cπN(u1) = cπN(u′) = cπN(u2) = x
with
x∈N
idempotent. Lemma 16 shows that
v=v1v′v2
with
ub≤Pol(V)vb
,
b= 1,2
, and
u′≤Pol(V)v′
. The induction
hypothesis asserts that
bη(vb)∈bρ(ub)
, for
b= 1,2
. From
u′≤Pol(V)v′
, we obtain that
u′≤Vv′
, implying that
(cπN(u′),cπN(v′)) = (x, cπN(v′))
is a
V-
pair of
N
. From
Pol(V)-stability and since bρ(ub) = ρ(x)for b= 1,2:
bη(v) = bη(v1v′v2) = bη(v1)bη(v′)bη(v2)∈ρ(x) = bρ(u).
VI. A S EQUENCE M0, M1, M2, . . . OF RELATI ONAL
EXPANSIONS
A. The difficulty
Before going further, we identify the challenges presented
when one tries to construct a
Pol(V)-
stable,
Pol(V)-
correct
relational expansion.
It is fairly easy to construct a
Pol(V)-
stable relational
expansion of a monoid
M
: the one induced by letting
η:M→M
be the identity map and
ρ:M△
−→ M
be
ρ(m) = M
, for any
m∈M
, is
Pol(V)-
stable, but since
its graph is
M×M
, this does not give us any interesting
information on
Pol(V)-
pairs. Similarly, we have seen in
Remark 15 that it is not hard to construct a
Pol(V)-
correct
relational expansion. An even simpler
Pol(V)-
correct relational
expansion than the one of Remark 15 is the one induced by
setting
η:M→M
to the identity and
ρ
to map any
m∈M
to
{m}
. This expansion is correct for any variety, since its graph
is simply
{(m, m)|m∈M}
. Constructing a
Pol(V)-
stable,
Pol(V)-
correct relational expansion is thus a task where our
two goals seem to pull in opposite directions.
A natural strategy is to start with the trivial relational
expansion, the one with
ρ(m) = {m}
, and to add elements
to
ρ(m)
, iteratively, so that we reach
Pol(V)-
stability — we
simply need to ensure that if we add an element
m′
to
ρ(m)
,
we have that (m, m′)is a Pol(V)-pair.
The most natural inductive definition is as follows. Let
η:M→M
be the identity, and for every
m∈M
, define
ρ0(m) = {m}
and set
ρi(m)
to be
ρi−1(m)
if
m2=m
, and,
otherwise, set ρi(m)to:
ρi−1(m)∪
{αm′β|α, β ∈ρi−1(m)∧(m, m′)aV-pair of M}.
Since this follows closely the definition of
Pol(V)-
stability,
this certainly seems to converge to a
Pol(V)-
stable relational
expansion (we disregard the technicality that
ρi
may not
be a relational morphism). One would like to conclude
that the relational expansion
M
induced by
(η, limiρi)
is
Pol(V)-
correct, by inductively proving that each relational
expansion is Pol(V)-correct. Let us attempt to do so.
Assume that the relational expansion at level
i−1
is
Pol(V)-
correct. We want to show that the new elements added
to the graph of the relational expansion at level
i
are indeed
Pol(V)-pairs.
Consider
m∈M
such that
m2=m
, a
V-
pair
(m, m′)
,
and
α, β ∈ρi−1(m)
, we want to show that
(m, αm′β)
is a
Pol(V)-
pair. Since
(m, m′)
is a
V-
pair of
M
, it is witnessed
by profinite words
u≤Vv
. Similarly, by hypothesis, we have
that
(m, α)
is witnessed by
uα≤Pol(V)vα
and
(m, β)
is
witnessed by
uβ≤Pol(V)vβ
. Assume that
uω=uα=uβ
,
then we can deduce:
uω+1 ≤Pol(V)uωvuω(Theorem 8)
≤Pol(V)uαvuβ.
Taking the image under
dπM
on both sides, we do find that
(m, αm′β)
is a
V-
pair. Here lies the crux of the difficulty:
We cannot assume, in general, that
uω=uα=uβ
. We have
no information within this relational expansion about how the
computed Pol(V)-pairs can be witnessed. We will thus build
relational expansions that embed some information to help us
“lift” pre-computed Pol(V)-pairs in a simultaneous way.
Let us stress that we do not know whether the relational
expansion we just constructed is
Pol(V)-
correct. As far as we
know, it might be, but we simply do not know how to show
it. We will come back to this point in Remark 25, where we
present a property of our more complex expansions that would
entail that this construction is indeed correct.
B. Construction
Let
M
be a finite monoid. We construct a sequence of
relational expansions of
M
with the specific aim of obtaining, in
the limit, a monoid that is a
Pol(V)-
stable relational expansion
of
M
, that we will show is
Pol(V)-
correct in Section VIII.
Each expansion will be a subsemigroup of
M×2M
where the
projection on the left will be surjective and the projection on
the right will be a relational morphism (consistent with the
definition of relational expansion). In the limit, this expansion
will have a neutral element and thus be a monoid.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
Machinery on M×2M.
We write
π: (M×2M)+→M×
2M
for the canonical semigroup morphism of
M×2M
. For
any element
x= (m, E)∈M×2M
, we write
π
ℓ
(x) = m
and
πr(x) = E
and extend
π
ℓ
and
πr
to morphisms from
(M×2M)+
to, respectively,
M
and
2M
. We have, in particular,
that
π(w) = (π
ℓ
(w), πr(w))
, and that
π
ℓ
◦π=π
ℓ
and
πr◦π=
πr.
We identify a partial order
⪯
over
M×2M
by
(m, E)⪯
(m′, E′)⇔(m=m′∧E⊆E′)
. For a set
K⊆M×2M
, we
write
↓K
for the set of elements in
M×2M
that are
⪯
-smaller
than some element in
K
. It should be noted that although,
strictly speaking,
M×2M
is an ordered monoid with respect
to this partial order, and the
↓
operator has the same meaning
as before, we will see
M×2M
as an unordered monoid, and
rely on ⪯as a tool in our construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
The construction.
We construct a sequence
M0, M1, M2, . . .
of subsemigroups of M×2Mas follows:
•M0={(m, {m})|m∈M};
•
For any
i
, we define a semigroup morphism
▷i: (M×
2M)+→M×2M
as follows. Let
x= (m, E)∈M×2M
.
Let
M1
i
be equal to
Mi
if
Mi
is a monoid, and to
Mi∪
{1Mi}
otherwise, with
1Mi
a new neutral element. We
extend
π
ℓ
to a monoid morphism by setting
π
ℓ
(1Mi) =
1M. Now let E′:= Eif x2=x, and
E′:= E∪{απ
ℓ
(y)β|(x, y)aV-pair of M1
i∧α, β ∈E};
otherwise. Then we set
▷i(x)=(m, E′)
, and extend this
to a morphism on (M×2M)+.
•
For any
i
, the set
Ki
is the subset of maximal elements
of
Mi
under
⪯
, in other words, it is the smallest subset
of Misuch that Mi⊆ ↓Ki;
•
For any
i
,
Mi+1 =⟨▷i(Ki)⟩
, that is,
Mi+1
is the
subsemigroup of
M×2M
generated by
▷i(Ki)
; in other
words, Mi+1 =▷i(K+
i).
We note these two simple properties about that construction:
Fact 1. For any i≥0and u∈d
M+
i,bπ(u)⪯b
▷i(u).
Fact 2. For any
i≥0
and any
u∈d
M+
i
,
bπ
ℓ
(u)∈bπr(u)
; this
shows that every
Mi
is indeed a semigroup that is a relational
expansion of M.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The state of affairs is depicted in Figure 1 as
a commutative diagram of semigroups to which we will be
adding more edges in Section VIII-A.
VII. THE SEQUENCE EVENTUALLY RE ACHES A MONOID
M∞THAT IS A Pol(V)-STABLE RELATIONAL EXPANSION
Lemma 19. There is a value
i
such that
▷i
is the identity
over
Mi
and
Mi=Mi+1 =· · ·
. The value of
i
can be taken
to be the first index such that Ki−1=Ki.
Proof.
For any
i
, by definition,
▷i(x, E) = (x, F )
for some
F⊇E
. This implies that
↓Ki⊆ ↓▷i(Ki)⊆ ↓Ki+1
. There is
thus a value isuch that:
↓Ki−1=↓▷i−1(Ki−1) = ↓Ki=↓▷i(Ki).
Since all the
Ki
’s only contain incomparable elements,
Ki−1=
Ki.
We now show that
▷i
is the identity over
Ki
. From
↓Ki=
↓▷i(Ki)
, we deduce that these sets have the same maximal
elements
Ki
, showing that
Ki⊆▷i(Ki)
. But
|▷i(Ki)| ≤
|Ki|
, since
▷i
is a function, so
▷i(Ki) = Ki
. Now for any
x∈Ki
, Fact 1 asserts that
x⪯▷i(x)
, and since
Ki
contains
only incomparable elements and both
x
and
▷i(x)
are in
Ki
,
▷i(x) = x
. The same argument shows that
▷i−1
is the identity
over Ki−1=Ki, and so:
Mi+1 =⟨▷i(Ki)⟩=⟨▷i−1(Ki−1)⟩=Mi.
We now show that
▷i
is the identity over
Mi
. For any
x∈Mi+1
, we have that
x=▷i(x1· · · xn)
for some
xj∈Ki
,
and so:
▷i(x) = ▷i(▷i(x1· · · xn))
=▷i(▷i(x1)· · · ▷i(xn))
=▷i(x1· · · xn) = x.
Since Mi=Mi+1, we are done.
Finally, since
Mi=Mi+1
, we have that
▷i=▷i+1
and
Ki=Ki+1
, showing that
Mi+2 =⟨▷i+1(Ki+1 )⟩
is the same
as
⟨▷i(Ki)⟩=Mi+1
. Iterating this argument shows that
Mi=
Mi+1 =· · ·.
We write
M∞, K∞,▷∞, . . .
for the objects at level
i
given
by the previous lemma. An immediate consequence of the
previous lemma is the following:
Corollary 20.
M∞
is a monoid that is a
Pol(V)-
stable
relational expansion of M.
Proof.
That
M∞
is a monoid is proved in appendix. We
argue that
M∞
is
Pol(V)-
stable. Let
(x, y)
be a
V-
pair
of
M∞
with
x2=x
and
α, β ∈πr(x)
; we need to show
that
απ
ℓ
(y)β∈πr(x)
. By construction,
πr(▷∞(x))
contains
απ
ℓ
(y)β
, but since
▷∞(x) = x
,
πr(▷∞(x)) = πr(x)
and we
are done.
VIII. M∞IS Pol(V)-CORRECT
The goal of this section is to show:
Theorem 21.
M∞
is a
Pol(V)-
correct relational expansion
of M.
Proof.
In the next section, in Lemma 22, we will show that
for every
i
, and
x∈Mi
, we can find a profinite word
τi(x)
whose image under
bπ
ℓ
is the same as
x
. Furthermore, for
every
m∈πr(x)
, there is a profinite word
um
that maps to
m
through
bπ
ℓ
, with
τi(x)≤Pol(V)um
. This immediately entails
that (π
ℓ
(x), m)is a Pol(V)-pair of M, by Lemma 9.
Thus for every
w∈K+
i
, and every
m∈πr(▷i(w))
, the pair
(π
ℓ
(w), m)
is a
Pol(V)-
pair of
M
. Since
M∞=⟨K∞⟩
and
▷∞
is the identity map over
M∞
, this implies in particular that
for every
x∈M∞
and every
m∈πr(x)
, the pair
(π
ℓ
(x), m)
is a Pol(V)-pair of M.
A. Burst and Lift
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Burst.
Let
i≥0
. Let us fix a pair of witnesses for each
V-
pair of
Mi
; specifically, for
(x, y)
a
V-
pair of
Mi
, let
uxy ≤Vvxy
be profinite words in
d
M+
i
such that
bπ(uxy ) = x
and bπ(vxy ) = y.
Let the continuous morphism
bursti:d
K+
i→d
M+
i
be defined,
for any x∈Ki, by bursti(x) = xif x2=x, and otherwise:
bursti(x) = Y
(x,y)aV-pair
of Mi
uω+1
xy .
Note that the order in which the product of the
uω+1
xy
is taken is
not important and that
bursti
is indeed a continuous morphism
(by Lemma 5). We note the following elementary fact:
Fact 3. For any iand u∈d
K+
i,bπ(bursti(u)) = bπ(u).
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Lift.
Let
i≥0
. We write
lifti:
[
M+
i+1 →d
K+
i
for the
continuous morphism defined by setting, for any
x∈Mi+1
,
lifti(x)
to be an arbitrary element
w∈K+
i
such that
▷i(w) = x
. Note that it is a morphism of words extended to the
respective profinite semigroups; in particular, if
x, y ∈Mi+1
and
z=xy
is their product in
Mi+1
, it is not true that
lifti(z) =
lifti(x)lifti(y)
. However,
▷i(lifti(z)) = ▷i(lifti(x)lifti(y))
and, more generally:
Fact 4. For any
i≥0
and any
u∈
[
M+
i+1
, we have
b
▷i(lifti(u)) = bπ(u).
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Burst and lift.
Define inductively
τi:d
K+
i→d
M+
0
by setting
τ0=burst0
and
τi=τi−1◦lifti−1◦bursti
. (Here,
τ
is short
for the French word
τ
émoin.) Combining Fact 3 and Fact 4:
Fact 5. For any
i≥0
and any
u∈d
K+
i
, we have that
bπ
ℓ
(u) = bπ
ℓ
(τi(u)).
M+
0K+
0M+
1K+
1M+
2K+
2· · ·
M0M1M2· · ·
M
ππ▷0
⊇
ππ▷1
⊇
ππ
⊇
π
ℓ
π
ℓ
π
ℓ
Figure 1. Commutative diagram of semigroups for the main construction. It indicates that some compositions of morphisms are equal, by following edges
from the same source and destination; for instance, following paths from K+
1to M, the diagram expresses that π
ℓ
◦▷1=π
ℓ
◦π.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The situation.
The current state of affairs is depicted as a
commutative diagram of semigroups in Figure 2. To include
bursti,lifti,
and
τi
in this diagram, the top row is a sequence
of profinite semigroups.
B. Correctness lemma
We are now ready to prove the correctness lemma:
Lemma 22. For any
i≥0
, any
u∈d
K+
i
, and any
m∈
πr(b
▷i(u))
, there is a profinite word
um∈d
M+
0
, such that
bπ
ℓ
(um) = mand τi(u)≤Pol(V)um.
Proof.
First, we note that it is enough to prove the claim for
the case where uis a single-letter word:
Fact 6. If the claim holds for any
u∈Ki
, then it holds for
any u∈d
K+
i.
Proof.
The full proof is in appendix. It proceeds by showing
that the statement is closed under concatenation over finite
words, in the sense that if the claim is satisfied for two finite
words
u:= w1
and
u:= w2
then it will also be satisfied for
u:= w1w2
. Then, it is argued that the claim over finite words
implies that it holds over profinite words, using elementary
techniques.
Equipped with the previous fact, we need only show the
claim for
u
an element of
Ki
that we will write
x
. We proceed
by induction on
i
. Let
m∈πr(▷i(x))
. We construct the word
um∈d
M+of the claim as follows.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
Base case: i= 0.
By definition of
M0
, we have that
x=
(t, {t})for some t∈M. We distinguish several cases:
•
If
x2=x
, then
τ0(x) = burst0(x) = x
, and
▷0(x) = x
.
Since
m∈πr(▷0(x)) = πr(x)
, we have that
m=t
.
Consequently, setting um:= xsatisfies the claim.
•
Assume now that
x2=x
and that
m
can be written as
απ
ℓ
(y)β
for a
V-
pair
(x, y)
of
M0
and
α, β ∈πr(x)
.
Since
πr(x) = {t}
, it must hold that
α=β=t
. We thus
have that
m=tπ
ℓ
(y)t
. Let
uxy ≤Vvxy
be the witnesses
in d
M+
0of the V-pair (x, y). We have that:
uω+1
xy ≤Pol(V)uω
xyvxy uω
xy.
Since there are profinite words
p, s ∈d
M+
0
with
bπ(p) =
bπ(s) = xsuch that burst0(x) = puω+1
xy s, we have:
τ0(x) = burst0(x) = puω+1
xy s≤Pol(V)puω
xyvxy uω
xys
As xis idempotent, um:= puω
xyvxy uω
xysis such that
bπ
ℓ
(um) = π
ℓ
(bπ(puω
xy)bπ(vxy )bπ(uω
xys))
=π
ℓ
(xyx) = tπ
ℓ
(y)t=m,
so that umsatisfies the claim.
•
The only case left is if
x2=x
and
m∈πr(x)
. This
entails that
m=t
. However, since
(x, x)
is a
V-
pair of
M0
, this means that
m=t=ttt
can be expressed as
απ
ℓ
(y)β
for
α, β ∈πr(x)
and
(x, y)
a
V-
pair of
M0
, and
this is thus covered by the previous case.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
Inductive step.
We assume the claim to be true at level
i−1and show it holds at level i. We consider two cases:
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Inductive step: Case m∈πr(x).
Let
u∈
[
K+
i−1
be de-
fined as
u=lifti−1(bursti(x))
. Fact 4 implies that
b
▷i−1(u) =
π(x)
, and we can thus apply the induction hypothesis on
u
and
m∈πr(b
▷i−1(u))
, giving us a profinite word
um∈d
M+
0
such that
bπ
ℓ
(um) = m
and
τi−1(u)≤Pol(V)um
. But
τi−1(u) = τi(x)
, so we obtain
τi(x)≤Pol(V)um
and we
are done.
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
Inductive step: Case m /∈πr(x).
We have
x2=x
, by
construction, and so
m=απ
ℓ
(y)β
for some
V-
pair
(x, y)
of
Mi
, and
α, β ∈πr(x)
. Consider the pair of profinite words
uxy ≤Vvxy
of
d
M+
i
that witness the
V-
pair
(x, y)
. We can
write
bursti(x)
as
puω+1
xy s
for some profinite words
p, s
that
both map to
x
through
bπ
. Since
uxy ≤Vvxy
, we have that
uω+1
xy ≤Pol(V)uω
xyvxy uω
xy, and so:
bursti(x) = puω+1
xy s≤Pol(V)puω
xyvxy uω
xys.
In order to use our induction hypothesis, let
upxy =
lifti−1(puω
xy)
and
uxys =lifti−1(uω
xys)
, two profinite words
in
[
K+
i−1
. By Fact 4, and since
bπ(puω
xy) = bπ(uω
xys) = x
, we
have that
b
▷i−1(upxy) = b
▷i−1(uxys) = x
. This shows that
α, β ∈πr(x) = πr(b
▷i−1(upxy)) = πr(b
▷i−1(uxys)).
We can thus apply the induction hypothesis at level
i−1
twice, setting first
u:= upxy
and
m:= α
then
u:= uxys
and
m:= β
, to obtain profinite words
uα, uβ∈d
M+
0
such
d
M+
0d
K+
0d
M+
1d
K+
1d
M+
2d
K+
2· · ·
M0M1M2· · ·
M
bπbπb
▷0
=
τ0
burst0
bπ
lift0
bπb
▷1
burst1
τ1
bπ
lift1
bπ
burst2
τ2
π
ℓ
π
ℓ
π
ℓ
Figure 2. Commutative diagram of semigroups for the main construction, with additional morphisms.
that
bπ
ℓ
(uα) = α
,
bπ
ℓ
(uβ) = β
,
τi−1(upxy)≤Pol(V)uα
, and
τi−1(uxys)≤Pol(V)uβ.
We thus obtain, via several applications of Lemma 7, the
following chain, where ≤is ≤Pol(V):
τi(x) = τi−1(lifti−1(bursti(x)))
=τi−1(lifti−1(puω+1
xy s))
≤τi−1(lifti−1(puω
xyvxy uω
xys))
≤τi−1(lifti−1(puω
xy)) ·τi−1(lifti−1(vxy )) ·
τi−1(lifti−1(uω
xys))
≤τi−1(upxy)·τi−1(lifti−1(vxy )) ·τi−1(uxys )
≤uα·τi−1(lifti−1(vxy)) ·uβ.
We can now set
um:= uα·τi−1(lifti−1(vxy)) ·uβ
, and we
obtain bπ
ℓ
(um) = απ
ℓ
(y)β=m.
IX. WRAPPING UP: THE ALGORITHM
A. The algorithm
To compute the
Pol(V)-
pairs of a monoid
M
, we can
compute each relational expansion
M1
i
one after the other
(recall that
M1
i
is
Mi
with possibly a neutral element added
to ensure it is a monoid). These relational expansions can be
computed provided that
V-
pairs are computable. Once we reach
Ki−1=Ki
, we have that
M1
i=M∞
(Lemma 19). Since
M∞
is
Pol(V)-
stable (Corollary 20) — hence
Pol(V)-
complete
(Lemma 18) — and
Pol(V)-
correct (Theorem 21), its graph
is exactly the set of
Pol(V)-
pairs of
M
. As an algorithm, we
obtain the pseudocode of Algorithm 1, in which
Exp
takes the
role of the successive expansions
M1
0, M 1
1, . . .
and
MaxElts
is used for
K0, K1, . . .
An analysis of this algorithm shows
that:
Theorem 23. If
V-
pairs are computable in time
cV(n)
with
cV(n)≥n
, then
Pol(V)-
pairs are computable in time
O(cV(2O(n))).
Algorithm 1
Compute
Pol(V)-
pair of a monoid, assuming
V-pairs are computable
1: procedure POLVPAIRS(M)▷ M a finite monoid
2: Exp ← {(m, {m})|m∈M}▷ Exp =M0=M1
0
3: MaxElts ←Exp ▷ M axElts =K0
4: repeat
5: NewExp ← {(m, E)∈Exp |m2=m}
6: for all (m, E)∈Exp with m2=mdo
7: E′←E
8: for all ((m, E),(m′, F )) V-pair of Exp do
9: for all α, β ∈Edo
10: E′←E′∪ {αm′β}
11: end for
12: end for
13: NewExp ←NewExp ∪ {(m, E′)}
14: end for
15: OldM axElts ←M axElts
16: Exp ← ⟨N ewExp⟩1
17: MaxElts ← {(m, T )∈Exp |
18: (∀T′=T)[(m, T ′)∈Exp ⇒T⊆ T′]}
19: until OldM axElts =M axElts
20: return {(m, m′)|(∃E)[(m, E)∈Exp ∧m′∈E]}
21: end procedure
B. Tightness: Early exit and a conjecture
The halting condition of the previous algorithm is the one
provided by Lemma 19. A different condition is to test whether
Exp
itself is
Pol(V)-
stable; however, if
Mi
is
Pol(V)-
stable,
then
Mi=Mi+1
, hence this condition would not reduce
significantly the number of times the main loop is taken. Rather
than testing whether
Mi
is
Pol(V)-
stable, we can test whether
Mi
’s graph induces a
Pol(V)-
stable relational expansion (with
M
as base monoid). This latter property can be worded as a
property of the graph itself, which can be tested in polynomial
time:
Lemma 24. We say that a subset
G⊆M×M
is
Pol(V)-
tight
if for every
V-
pair
(m, n)
of
M
with
m2=m
and for every
(m, m′),(m, m′′)∈G, we have that (m, m′nm′′ )∈G.
Consider the sequence of relational expansions
M0, M1, . . .
of
M
defined earlier. If there is
i≥0
such that the graph
G
of
Miis Pol(V)-tight, then Gis the set of Pol(V)-pairs of M.
Proof. By Theorem 21, Gcontains only Pol(V)-pairs of M.
Note that
G
is a submonoid of
M×M
: if
(m, m′),(n, n′)∈
G
, then there are
x, y ∈Mi
such that
m=π
ℓ
(x), n =π
ℓ
(y)
,
and
m′∈πr(x), n′∈πr(y)
. Consequently,
mn =π
ℓ
(xy)
and
m′n′∈πr(x)πr(y)⊆πr(xy). Hence (mn, m′n′)∈G.
Consider now the relational expansion of
M
induced by
letting
η:M→M
be the identity and letting
ρ(m) = {m′|
(m, m′)∈G}
for any
m∈M
. Note that
ρ
is indeed a
relational morphism, since
G
is a submonoid of
M×M
, and
that, by construction, the graph of this expansion is G.
The hypothesis on
G
immediately implies that this expansion
is
Pol(V)-
stable. By Lemma 18,
G
thus contains all the
Pol(V)-pairs of M.
Remark 25. If any relational expansion
Mi
has a
Pol(V)-
tight
graph
G
, then
G
is the smallest submonoid
P
of
M×M
that is
Pol(V)-
tight. To show this, first note that
P⊆G
by definition.
Next, the argument of the previous lemma showing that
G
contains all
Pol(V)-
pairs of
M
can be repeated for
P
, hence
G⊆P, since Gis the set of all Pol(V)-pairs.
This set
P
is also the graph of the construction presented
in Section
VI-A
and it is computable in polynomial time given
the monoid
M
, provided that
V-
pairs are also computable in
polynomial time. We conjecture:
Conjecture 26. For any variety of ordered monoid
V
and
any monoid
M
, one of the expansions
M0, M1, . . .
has a
Pol(V)-
tight graph. Consequently, if
V-
pairs are computable
in polynomial time, so are Pol(V)-pairs.
X. APPLICATION TO CONCATENATION HIERARCHIES
Definition 27. The concatenation hierarchy over a positive
variety of regular languages
V
is defined by setting
ch0(V) = V
and, for any
i≥0
,
chi+1
2(V) = Pol(chi(V))
and
chi+1(V) =
Bool(chi+1
2(V))
, where
Bool
denotes the Boolean closure of
a class. We write
ch(V)
for
Sichi(V)
. We write
chi(V)
and
ch(V)
for the algebraic counterparts of these varieties, as
provided by the ordered version of Eilenberg theorem.
Remark 28. Two concatenation hierarchies are extensively
studied in the literature. The first one was mentioned in the
introduction: it is defined by setting
V:= {∅, A∗}
, and it turns
out that
chi+1
2(V)
is
Σi[<]
. The other one is defined by setting
V:= G
, the variety of group languages, that is, languages
recognized by groups. We note that separability for
V
, in both
cases, can be decided in polynomial time: this is trivial for
V={∅, A∗}
and proved by van Rooijen [28], and appearing
in [26], in the case of V=G.
Theorem 29. Let
V
be a positive variety of regular languages
and assume
V-
separation is decidable in time
cV(n)
. Let
Tower(n, i)
be defined by
Tower(n, 0) = n
and
Tower(n, i) =
2Tower(n,i−1)
. For any
i≥0
,
chi+1
2(V)-
separation is decidable
in time Tower(O(cV(n)), i).
Corollary 30. Assume Conjecture 26 holds. Let
V
be a positive
variety of regular languages and assume
V-
separation is
decidable in time
cV(n)
. For any
i≥0
,
chi+1
2(V)-
separation
is decidable in time polynomial in
cV(n)
. Further, so is
ch(V)-separation.2
Proof.
The first part is immediate. As for
ch(V)-
separability,
note that Conjecture 26 implies that there is a function that
takes a set of pairs
E
of a monoid
M
, and produces another set
of pairs
E′
, such that for any variety
W
, if
E
were the
W-
pairs
of
M
, then
E′
is the
Pol(W)-
pairs of
M
. In particular, this
means that for any
i≥0
, if the
chi−1
2(V)-
pairs of
M
are
the same as the
chi+1
2(V)-
pairs, then the set of pairs is the
same for any
chj+1
2(V)
,
j≥i
. Since the sets of pairs only
shrink in size from
chi−1
2(V)
to
chi+1
2(V)
, this means that the
ch(V)-
pairs of
M
are the same as its
ch|M|2+1
2(V)-
pairs. We
can thus compute the
ch(V)-
pairs in polynomial time and, by
Corollary 11, decide
ch(V)-
separation in polynomial time.
XI. PERSPECTIVES AND OPEN QUESTIONS
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
Boolean closures.
We leave open whether membership is
decidable for the Boolean closure of
Σi[<]
. Even though the
literature has long been focused on a solution for both
Σi[<]
and its Boolean closure, the historical “dot-depth” question
was asked about the latter. However, we conjecture:
Conjecture 31. For any positive variety
V
of regular languages,
if
V-
separation is decidable, then membership in the Boolean
closure of Vis decidable.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
FO with other predicates.
Common extensions of
FO
revolve
around equipping first-order logic with additional predicates on
the numerical value of positions to define more languages. The
logic
FO[arb]
, for instance, allows any predicate (e.g.,
x=y×
z
), while
FO[<, +1,mod]
only enriches
FO[<]
with predicates
x=y+1
and
x= 0 mod p
, for any constant
p
. A striking fact,
conjectured by McNaughton in the 1960s [11] and showed by
Barrington, et al. [5], is that the regular languages of
FO[arb]
are precisely the languages expressible in
FO[<, +1,mod]
.
Techniques of [20] translate the decidability of membership
to the alternation hierarchy
Σi[<]
to the alternation hierarchy
Σi[<, +1]
, while an ordered-monoid version of [9] would show
the same for
Σi[<, +1,mod]
. This leaves open the decidability
for
Σi[mod]
. If the requirement of “variety” were relaxed to
“quotienting algebra” in our main result, then a result of [24]
would entail:
Conjecture 32. Separation for Σi[mod]is decidable.
2
With
V={∅, A∗}
, the class
ch(V)
is equal to the class
SF
of star-free
languages. It is known that
SF-
pairs are computable in exponential time,
hence, by Corollary 11,
SF -
separation is also decidable in exponential time.
REFERENCES
[1]
Jorge Almeida. “Some algorithmic problems for pseu-
dovarieties”. In: Publ. Math. Debrecen 54.1 (1999),
pp. 531–552.
[2]
Jorge Almeida, Ondˇ
rej Klíma, and Michal Kunc. In:
Forum Mathematicum 30.3 (2018), pp. 663–679. DOI:
10.1515/forum-2016-0028.
[3]
Mustapha Arfi. “Opérations polynomiales et hiérarchies
de concaténation”. In: Theoretical Computer Science
91.1 (1991), pp. 71–84.
[4]
Corentin Barloy, Michaël Cadilhac, Charles Paperman,
and Thomas Zeume. “The Regular Languages of First-
Order Logic with One Alternation”. In: LICS ’22: 37th
Annual ACM/IEEE Symposium on Logic in Computer
Science, Haifa, Israel, August 2 - 5, 2022. Ed. by Christel
Baier and Dana Fisman. ACM, 2022, 58:1–58:11. DOI:
10.1145/3531130.3533371.
[5]
David A. Mix Barrington, Kevin Compton, Howard
Straubing, and Denis Thérien. “Regular Languages in
NC
1
”. In: J. Computer and System Sciences 44.3 (1992),
pp. 478–499.
[6]
Janusz A. Brzozowski and Robert Knast. “The dot-depth
hierarchy of star-free languages is infinite”. In: Journal
of Computer and System Sciences 16.1 (1978), pp. 37–55.
ISSN: 0022-0000. DOI: http://dx.doi.org/10.1016/0022-
0000(78)90049-1.
[7]
Rina S. Cohen and J.A. Brzozowski. “Dot-depth of
star-free events”. In: Journal of Computer and System
Sciences 5.1 (1971), pp. 1–16. IS SN: 0022-0000. DOI:
https://doi.org/10.1016/S0022-0000(71)80003-X.
[8]
Thomas Colcombet. “The factorisation forest theorem”.
In: Handbook of Automata Theory. Ed. by Jean-Éric
Pin. European Mathematical Society Publishing House,
Zürich, Switzerland, 2021, pp. 653–693. DOI: 10.4171/
AUTOMATA-1/18.
[9]
Luc Dartois and Charles Paperman. “Alternation Hierar-
chies of First Order Logic with Regular Predicates”. In:
Fundamentals of Computation Theory - 20th Interna-
tional Symposium, FCT 2015, Gda´nsk, Poland, August
17-19, 2015, Proceedings. Ed. by Adrian Kosowski and
Igor Walukiewicz. Vol. 9210. Lecture Notes in Computer
Science. Springer, 2015, pp. 160–172. DOI: 10.1007/978-
3-319-22177-9\_13.
[10]
Samuel Eilenberg. Automata, Languages, and Machines,
Volume B. Pure and Applied Mathematics. Academic
Press, 1976.
[11]
Robert McNaughton. Symbolic Logic and Automata.
Tech. rep. 60-244. Wright Air Development, 1960.
[12]
Robert McNaughton and Seymour Papert. Counter-Free
Automata. Cambridge, Mass.: The MIT Press, 1971.
[13]
Dominique Perrin and Jean-Éric Pin. “First-order logic
and star-free sets”. In: Journal of Computer and System
Sciences 32.3 (1986), pp. 393–406. IS SN: 0022-0000.
DOI: https://doi.org/10.1016/0022-0000(86)90037-1.
[14] Jean-Éric Pin. “A variety theorem without complemen-
tation”. In: Russian Mathematics 39 (1995), pp. 80–90.
[15]
Jean-Éric Pin. “An Explicit Formula for the Intersection
of Two Polynomials of Regular Languages”. In: Devel-
opments in Language Theory. Ed. by Marie-Pierre Béal
and Olivier Carton. Berlin, Heidelberg: Springer Berlin
Heidelberg, 2013, pp. 31–45. ISBN: 978-3-642-38771-5.
[16]
Jean-Éric Pin. “Bridges for concatenation hierarchies”.
In: Automata, Languages and Programming. Ed. by
Kim G. Larsen, Sven Skyum, and Glynn Winskel. Berlin,
Heidelberg: Springer Berlin Heidelberg, 1998, pp. 431–
442. ISBN: 978-3-540-68681-1.
[17]
Jean-Éric Pin. “Profinite Methods in Automata The-
ory”. In: 26th International Symposium on Theoretical
Aspects of Computer Science. Ed. by Susanne Albers
and Jean-Yves Marion. Vol. 3. Leibniz International
Proceedings in Informatics (LIPIcs). Dagstuhl, Germany:
Schloss Dagstuhl – Leibniz-Zentrum für Informatik,
2009, pp. 31–50. IS BN: 978-3-939897-09-5. DO I: 10.
4230/LIPIcs.STACS.2009.1856.
[18]
Jean-Éric Pin. “The Dot-Depth Hierarchy, 45 Years
Later”. In: The Role of Theory in Computer Science.
2017.
[19]
Jean-Éric Pin and Pascal Weil. “Polynomial closure and
unambiguous product”. In: Theory of Computing Systems
30 (1997), pp. 383–422.
[20]
Jean-Éric Pin and Pascal Weil. “The wreath product
principle for ordered semigroups”. In: Communications
in Algebra 30.12 (2002), pp. 5677–5713.
[21]
Pascal Pin Jean-Éricand Weil. “A Reiterman theorem
for pseudovarieties of finite first-order structures”. In:
Algebra Universalis 35 (1996), pp. 577–595.
[22] Thomas Place. “Separating regular languages with two
quantifier alternations”. In: Log. Methods Comput. Sci.
14.4 (2018). DOI: 10.23638/LMCS-14(4:16)2018.
[23]
Thomas Place and Marc Zeitoun. “Characterizing level
one in group-based concatenation hierarchies”. In: In-
ternational Computer Science Symposium in Russia.
Springer. 2022, pp. 320–337.
[24]
Thomas Place and Marc Zeitoun. “Generic Results for
Concatenation Hierarchies”. In: Theory Comput. Syst.
63.4 (2019), pp. 849–901. DOI: 10.1007/S00224-018-
9867-0.
[25]
Thomas Place and Marc Zeitoun. “Going Higher in First-
Order Quantifier Alternation Hierarchies on Words”. In:
J. ACM 66.2 (2019), 12:1–12:65. DOI: 10.1145/3303991.
[26]
Thomas Place and Marc Zeitoun. “Group Separation
Strikes Back”. In: 38th Annual ACM/IEEE Symposium
on Logic in Computer Science, LICS 2023, Boston, MA,
USA, June 26-29, 2023. IEEE, 2023, pp. 1–13. DOI:
10.1109/LICS56636.2023.10175683.
[27]
Thomas Place and Marc Zeitoun. “Separation and
covering for group based concatenation hierarchies”.
In: 2019 34th Annual ACM/IEEE Symposium on Logic
in Computer Science (LICS). IEEE. 2019, pp. 1–13.
[28]
Lorijn van Rooijen. “A combinatorial approach to the
separation problem for regular languages. (Une approche
combinatoire du problème de séparation pour les lan-
gages réguliers)”. PhD thesis. University of Bordeaux,
France, 2014. UR L: https://tel.archives-ouvertes.fr/tel-
01154542.
[29]
Marcel-Paul Schützenberger. “On Finite Monoids Having
Only Trivial Subgroups”. In: Information and Control
8.2 (1965), pp. 190–194.
[30]
Imre Simon. “Factorization forests of finite height”. In:
Theoretical Computer Science 72.1 (1990), pp. 65–94.
ISSN: 0304-3975. DOI: https://doi.org/10.1016/0304-
3975(90)90047-L.
[31]
Imre Simon. “Piecewise testable events”. In: Automata
Theory and Formal Languages. Ed. by H. Brakhage.
Berlin, Heidelberg: Springer Berlin Heidelberg, 1975,
pp. 214–222. IS BN: 978-3-540-37923-2.
[32]
Howard Straubing. Finite Automata, Formal Logic, and
Circuit Complexity. Boston: Birkhäuser, 1994. DOI: 10.
1007/978-1-4612-0289-9.
[33]
Howard Straubing. “Finite semigroup varieties of the
form V
∗
D”. In: Journal of Pure and Applied Algebra
36 (1985), pp. 53–94. IS SN: 0022-4049. DO I: https :
//doi.org/10.1016/0022-4049(85)90062-3.
[34]
Denis Thérien. “Classification of finite monoids: the
language approach”. In: Theoretical Computer Science
14.2 (1981), pp. 195–208. ISSN: 0304-3975. DO I: https:
//doi.org/10.1016/0304-3975(81)90057-8.
[35]
Wolfgang Thomas. “An application of the Ehrenfeucht-
Fraisse game in formal language theory”. eng. In:
Mémoires de la Société Mathématique de France 16
(1984), pp. 11–21. UR L: http://eudml.org/doc/94847.
CONTENTS
I Introduction 1
II Preliminaries 3
II-A Sets, languages . . . . . . . . . . . . . 3
II-B Algebraic structures . . . . . . . . . . . 3
II-C Profinite words and inequalities . . . . . 4
II-D V-pairs.................. 5
II-E The semantics of V-pairs: separation . 5
III Main theorem and structure of the proof 6
IV
Correct, complete, and stable relational expan-
sions 6
VPol(V)-stable implies Pol(V)-complete 7
VI
A sequence
M0, M1, M2, . . .
of relational expan-
sions 8
VI-A The difficulty . . . . . . . . . . . . . . 8
VI-B Construction . . . . . . . . . . . . . . . 8
VII
The sequence eventually reaches a monoid
M∞
that is a Pol(V)-stable relational expansion 9
VIII M∞is Pol(V)-correct 9
VIII-A Burst and Lift . . . . . . . . . . . . . . 9
VIII-B Correctness lemma . . . . . . . . . . . 10
IX Wrapping up: The algorithm 11
IX-A The algorithm . . . . . . . . . . . . . . 11
IX-B Tightness: Early exit and a conjecture . 11
X Application to Concatenation Hierarchies 12
XI Perspectives and Open Questions 12
Appendix 15
A Proof of Lemma 4 . . . . . . . . . . . . 15
B Proof of Lemma 5 . . . . . . . . . . . . 15
C Proof of Lemma 6 . . . . . . . . . . . . 15
D Proof of Lemma 7 . . . . . . . . . . . . 16
E Proof of Theorem 8 . . . . . . . . . . . 16
F Proof of Lemma 9 . . . . . . . . . . . . 17
G Proof of Corollary 11 . . . . . . . . . . 17
H Proof of Lemma 16 . . . . . . . . . . . 17
I Proof of Theorem 17 . . . . . . . . . . 18
J Proof of Fact 1 . . . . . . . . . . . . . 18
K Proof of Fact 2 . . . . . . . . . . . . . 18
L Proof of the first part of Corollary 20 . 19
M Proof of Fact 3 . . . . . . . . . . . . . 19
N Proof of Fact 4 . . . . . . . . . . . . . 19
O Proof of Fact 5 . . . . . . . . . . . . . 19
P Proof of Fact 6 . . . . . . . . . . . . . 19
Q Proof of Theorem 23 . . . . . . . . . . 19
R Proof of Theorem 29 . . . . . . . . . . 19
APPENDIX
A. Proof of Lemma 4
Let
(ui)i>0
be a convergent sequence of profinite words
over
A.
We will show how to construct a profinite word
w
such that
bµ(w) = limibµ(ui).
We enumerate all the morphisms
ϕi:A∗→Mi, i > 0
into finite monoids. This can be done because there are only
countably many such morphisms. We then form direct products
of these, giving a morphism
ψi=ϕ1× · · · ϕi:A∗→M1× · · · × Mi, i > 0.
Let
mi= limkb
ϕi(uk)∈Mi.
We then have
limkb
ψi(uk) = (m1, . . . , mi).
We claim that there is a convergent sequence
(wi)i>0
of
finite words over Asuch that for each j > 0,
limkψj(wk)=(m1, . . . , mj).
Assuming the claim for the moment, we complete the proof
as follows: If
µ:A∗→M
is a morphism with
M
finite, then
µ=ϕj
and
M=Mj
for some
j.
Since
ϕj
is the composition
of
ψj
with projection onto the
jth
component, our claim shows
that
limkµ(wk) = mj= limkbµ(uk).
Letting
w
denote the profinite word represented by the sequence
(wi)i≥0gives the desired result.
To prove the claim, for each
p > 0,
we choose
kp
large
enough so that
c
ψp(uk) = limkc
ψp(uk)=(m1, . . . , mp)
for all
k≥kp.
We let
vp,1, vp,2, . . .
be a sequence of finite
words representing ukp.Since
limrψp(vp,r) = c
ψp(ukp)=(m1, . . . , mp),
we can choose
rp
large enough so that
ψp(vp,rp) =
(m1, . . . , mp).Set wp=vp,rp.
Let
j > 0.
If
k≥j
then
ψj(wk)
is the projection of
ψk(wk)
onto the first jcomponents; that is
ψj(wk)=(m1, . . . , mj)
for k≥j, which proves the claim.
Observe that by definition, all sequences of finite words
representing the sequence
(ui)i>0
in this sense are equivalent,
so the profinite word wis unique.
B. Proof of Lemma 5
Let µ:A∗→c
B∗as in the statement of the lemma. Let
u= (ui)i≥0
be a convergent sequence of finite words over
A.
The sequence
(µ(ui))i≥0
of profinite words over
B
is also convergent, since if
ν:B∗→M
is a morphism into a finite monoid, the sequence
(bν(µ(ui)))i>0
is ultimately constant. It is immediate that if
u′= (u′
i)i≥0
is another convergent sequence equivalent to u,
then the sequences
(µ(ui))i≥0,(µ(u′
i))i≥0
are also equivalent,
in the sense that they have the same limit under any morphism
into a finite monoid. By Lemma 4,
(µ(ui))i≥0
is represented by
a unique profinite word
w
over
B.
We thus get a well-defined
map bµ:u7→ wfrom c
A∗to c
B∗.
It is clear that if
bµ(u),bµ(u′)∈c
A∗
are represented by
w, w′
in this way, then
ww′
represents
bµ(uu′),
and thus
bµ
is a
morphism. We are also able to conclude, again by composing
bµ
with a morphism
ν:B∗→M
with
M
finite, that if
(ui)i>0
is a convergent sequence of profinite words over A, then
limibµ(ui) = bµ(limiui).
Thus bµis continuous.
C. Proof of Lemma 6
(Many readers will recognize this as an application of the
König infinity lemma; however we will spell out the argument
in detail.) Let
u= (ui)i≥0
be a sequence of profinite words. We
will show how to extract a convergent subsequence. As in the
proof of Lemma 4, we consider an enumeration
ϕi:A∗→Mi
of the morphisms from
A∗
into finite monoids, and the products
ψi=ϕ1× · · · ϕi:A∗→M1× · · · × Mi
of these morphisms. Since
M1
is finite, there is some
m1∈M1
such that
ϕ1(ui) = m1
for infinitely many values of
i.
We
thus extract the subsequence consisting of these
i.
We denote
this subsequence by (u1,i)i≥0as well.
We next look at the elements
ψ2(u1,k)
of
M1×M2
for
k≥2.
The first component of all these elements is
m1
. Again,
since
M1×M2
is finite there is some
m2∈M2
that appears
infinitely many times among these elements. We extract the
subsequence of these elements, and label the resulting terms
u2,2, u2,3,....
We continue in this fashion. After the
jth
step we have a
subsequence
u1,1, u2,2...,uj,j , uj,j+1 , uj,j+2
such that if
i≤k≤j, ψi(uk,k )=(m1, . . . , mi).
The result
is an infinite subsequence
(uk,k )k>0
such that for each
i,
the sequence
(ψi(uk,k ))k>0
is ultimately constant. Thus the
sequence
(ϕi(uk,k ))k>0
is ultimately constant as well. Since
every morphism from
A∗
into a finite monoid is one of the
ϕi,
we conclude that the subsequence (uk,k )k>0is convergent.
D. Proof of Lemma 7
(a) Let
ui≤Vvi,
for
i= 1,2,
and let
M∈V.
Let
µ:A∗→
M
be a morphism and let
mi=bµ(ui), ni=bµ(vi).
We then
have, by multiplicativity of the partial order on M,
bµ(u1u2) = m1m2≤n1n2=bµ(v1v2),
so u1u2≤Vv1v2.
(b) let
µ:A∗→M
be a morphism, with
M∈V.
Let
u=
(ui)i>0, v = (vi)i>0,
be convergent sequences of profinite
words over A. For isufficiently large, we have
bµ(u) = µ(ui)≤µ(vi) = bµ(v).
So u≤Vv.
(c) Let
µ:A∗→B∗
be a morphism with
u, v ∈c
A∗, u ≤Vv.
Let ν:B∗→M, where M∈V.From u≤Vvwe have
bν(bµ(u)) ≤bν(bµ(v)),
and thus bµ(u)≤Vbµ(v).
E. Proof of Theorem 8
We start with some routine lemmas on profinite words.
Lemma 33. Assume that
pxωq=u1· · · un
for some profinite
words
p, x, q, u1, . . . , un
. There are
i≥1
and profinite words
α, β such that:
1) pxω=u1· · · ui−1α,
2) αxωβ=ui,
3) xωq=βui+1 · · · un.
Proof. This is shown by induction on n.
If
n= 1
, then
pxωq=u1
. We let
i:= 1
,
α:= pxω
,
β:= xωq
, and we indeed have that
αxωβ=pxω·xω·xωq=
pxωq=u1.
For
n > 1
, assume
pxωq
can be decomposed as
u1· · · un
.
Writing
u′
n−1=un−1un
and applying the induction hypothesis
on the decomposition
pxωq=u1· · · un−2u′
n−1
, we obtain
i′, α′, β′
with the properties of the lemma. If
i′< n −1
then
we are done, since the same
i′, α′, β′
also satisfy the properties
for the original decomposition. If i′=n−1, then we have:
α′xωβ′=u′
n−1=un−1un,
and so in particular
un−1un= (α′xω)(xωβ′)
. By equidivisi-
bility, we get that either:
•un−1=α′xωβ
for some profinite word
β
, in which case
we set i:= n−1,α:= α′, and we are done;
•un=αxωβ′
for some profinite word
α
, in which case
we set i:= n,β:= β′, and we are done.
Lemma 34. Assume that
pxω+1q=u1· · · un
for some
profinite words
p, x, q, u1, . . . , un
. There are
i≥1
and profinite
words α, β such that:
1) pxω=u1· · · ui−1α,
2) αxω+1β=ui,
3) xωq=βui+1 · · · un.
Proof.
We apply Lemma 33 on
(px)·xω·q
, that is, by
setting
p:= px
, and obtain
i′, α′, β′
such that: 1)
(px)xω=
u1· · · ui′−1α′
, 2)
α′xωβ′=ui′
, 3)
xωq=β′ui′+1 · · · un
. By
multiplying the first equation by
xω−1
on the right, we obtain:
pxω= (px)xωxω−1=u1· · · ui′−1α′xω−1.
Let
i:= i′
,
α:= α′xω−1
, and
β:= β′
. We obtain, as required:
1) pxω=u1· · · ui−1α,
2) αxω+1β=α′xω−1xω+1 β′=α′xωβ′=ui′=ui,
3) xωq=β′ui′+1 · · · un=βui+1 · · · un.
The main statement is now proved as follows:
Theorem 35. Let
V
be a variety of ordered monoids. The
following holds:
Pol(V) = Jxω+1 ≤xωyxω,for any x≤VyK.
Proof. Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Left to right inclusion.
Let
L∈Pol(V)
, we
show that it satisfies the inequalities. By definition,
L=
L0a1L1· · · anLn
for languages
Li
in
V
. By continuity of
concatenation, we also have that:
L=L0a1L1· · · anLn.
Let
x≤Vy
and
p, q ∈c
A∗
. We need to show that
pxω+1q∈L
implies
pxωyxωq∈L
in order to conclude
this proof direction. From pxω+1q∈Lwe get that
pxω+1q=v0a1v1· · · anvn,with, for all i, vi∈Li.
If
x=ε
, we need to show that
pyq ∈L
. In that case,
let
i, α, β
be such that
p=v0a1v1· · · aiα
,
vi=αβ
, and
q=βai+1 vi+1 · · · anvn
. Since
vi∈Li
and
Li∈V
, and since
ε≤Vyby hypothesis, αyβ ∈Li, and thus pyq ∈L.
If
x=ε
, we apply Lemma 34 and since no letter
ai
can be
written as αxω+1β, we have i, α, β such that:
1) pxω=v0a1v1· · · aiα,
2) αxω+1β=vi,
3) xωq=βai+1 vi+1 · · · anvn.
Since
vi∈Li
and
Li∈V
, since
x≤Vy
, and since
αxω+1β=
αxωxxωβ
, we get that
v′
i:= αxωyxωβ∈Li
. Finally, we have,
as required, that:
(pxω)·y·(xωq) = (pxω)·(xωyxω)·(xωq)
= (v0a1v1· · · aiα)(xωyxω)(βai+1vi+1 · · · anvn)
=v0a1v1· · · ai(αxωyxωβ)ai+1 vi+1 · · · anvn
=v0a1v1· · · aiv′
iai+1vi+1 · · · anvn∈L.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Right to left inclusion.
Let
L
be a language recognized
by a morphism
η:A∗→(M, ≤)
such that
M
satisfies the
inequalities of the statement. We show that L∈Pol(V).
We may assume that
η
is surjective by restraining its range
to its image, which would be a quotient of
M
, hence still
satisfy the inequalities of the statement.
Fix, for all
m∈M
, a language
Lm∈V
with
η−1(m)⊆
Lm
and such that
η−1(m′)∩Lm=∅
if, and only if,
(m, m′)
is a V-pair of M. Such languages exist by Lemma 10.
For any word
w∈A∗
, we build a language
Kw
by induction
on the factorization forest for w:
•
If
w
is a single letter or the empty word, then
Kw=
L1M{w}L1M,
•
In the binary case, with
w=w1w2
, we set
Kw=
Kw1Kw2,
•
In the idempotent case, with
w=w1· · · wn
and
η(w) =
η(w1) = · · · =η(wn) = m
an idempotent of
M
, we set
Kw=Kw1LmKwn.
Clearly, for any
w
,
Kw∈Pol(V)
. Moreover, there are only
finitely many different
Kw
since factorization forests are of
bounded depth. To conclude, we show that L=Sw∈LKw.
Fact 7. For any w∈A∗,w∈Kw.
Proof.
This is shown by induction on the forest factorization
of
w∈A∗
. This is true if
w
is a single letter, since
ε∈L1M
.
The binary case is immediate from the induction hypothesis.
In the idempotent case, using the notations of the construction,
it is enough to note that
w1∈Kw1, wn∈Kwn
by induction
hypothesis and w2· · · wn−1∈η−1(m)⊆Lm.
Fact 8. For any w∈A∗and any w′∈Kw,η(w)≤η(w′).
Proof.
This is shown by induction on the forest factorization
of
w∈A∗
. If
w
is a letter, then
w′=w′
1ww′
2
with
η(w′
1) =
η(w′
2)=1M. This implies that η(w) = η(w′).
In the binary case,
w
is factorized as
w1w2
, and thus
w′=
w′
1w′
2
with
w′
1∈Kw1
,
w′
2∈Kw2
. By induction hypothesis,
we have that
η(w1)≤η(w′
1)
and
η(w2)≤η(w′
2)
, and so
η(w) = η(w1)η(w2)≤η(w′
1)η(w′
2) = η(w′).
In the idempotent case,
w=w1w2· · · wn
with
η(wi) = m
an idempotent of
M
for every
i
. By construction,
w′=
w′
1w′
2w′
3
with
w′
1∈Kw1, w′
2∈Lm, w′
3∈Kwn
. From
w′
2∈
Lm
, we deduce that
(m, η(w′
2))
is a
V-
pair of
M
, that is,
there are profinite words
u≤Vv
such that
bη(u) = m, bη(v) =
η(w′
2)
. Since
L
satisfies the inequality
uω+1 ≤uωvuω
, and
since
m
is idempotent, we get that
m≤mη(w′
2)m
. The
induction hypothesis asserts that
m≤η(w′
1), η(w′
3)
, and so
we obtain
η(w) = m≤η(w′
1)η(w′
2)η(w′
3) = η(w′)
and we
are done.
The first fact shows that
L⊆Sw∈LKw
. For the converse
direction, let
E⊆M
such that
L=η−1(↑E)
. Let
w∈L
and w′∈Kw. The second fact shows that η(w)≤η(w′)and
since
η(w)∈ ↑E
, we have that
η(w′)∈ ↑E
, implying that
w′∈L.
F. Proof of Lemma 9
Let
u, v ∈c
A∗
with
u≤Vv,
and
µ:A∗→M
a morphism
into a finite monoid. We first show that
(bµ(u),bµ(v))
is a
V
-
pair of
M.
If we consider
M
as a finite alphabet, then the map
µ|A:A→M
extends to a morphism
θ:A∗→M∗
such that
πMθ=µ. We have b
θ(u)≤Vb
θ(v)by Lemma 7. Thus
(bµ(u),bµ(v)) = ( d
πMθ(u),d
πMθ(v)) = (dπM(b
θ(u)),dπM(b
θ(v)))
is a V-pair of M.
Conversely, suppose that
(m1, m2)
is a
V
-pair of
M,
and
that
µ
is surjective. We will show that
(m1, m2) = (bµ(u),bµ(v))
for some
u, v ∈c
A∗
with
u≤Vv.
There exist
x, y ∈M∗
such
that
x≤Vy
and
u=dπM(x), v =dπM(y).
Viewing
M
as a
finite monoid, we map
θ:M→A∗
by setting
θ(m) = w,
where
µ(w) = m.
(Surjectivity guarantees that such a word
w
exists.) Let
u=b
θ(x), v =b
θ(y).
By Lemma 7,
u≤Vv.
Since
bµ(u) = m1,bµ(v) = m2,(m1m2)is a V-pair.
G. Proof of Corollary 11
The left-to-right direction is an immediate consequence of
Lemma 10.
For the other direction, consider
L, L′⊆A∗
two regular
languages. First, we note that we can assume that
L
and
L′
are recognized by the same surjective morphism. Indeed, let
µ:A∗→M, µ′:A∗→M′
be surjective morphisms such that
there are sets
E, E ′
with
L=µ−1(↑E)
and
L′=µ′−1(↑E′)
.
We can construct the surjective morphism
θ:A∗→M×M′
by letting
θ(a)=(µ(a), µ′(a))
for any
a∈A
, and then
L=θ−1(↑E×M′)
and
L′=θ−1(M× ↑E′)
— note that
these are indeed inverse images of upsets.
Thus, we assume that
L, L′
are recognized by the surjective
morphism
µ:A∗→M
, so that
L=µ−1(↑E)
and
L′=
µ−1(↑E′)
. We show that
L
is
V-
separable from
L′
if, and only
if, for every
m∈ ↑E
and
m′∈ ↑E′
, the language
µ−1(m)
is
V-
separable from
µ−1(m′)
. This, together with Lemma 10,
entails that
V-
separation is decidable if
V-
pairs are computable.
Assume
K∈ V
separates
L
from
L′
, then certainly, for
every
m∈ ↑E
and
m′∈ ↑E′
,
K
separates
µ−1(m)
from
µ−1(m′)
. Conversely, if for any
m∈ ↑E
and
m′∈ ↑E′
there is a language
Km,m′∈ V
that separates
µ−1(m)
from
µ−1(m′)
, then
K=Sm∈↑ETm′∈↑E′Km,m′
separates
L
from
L′
; indeed,
Tm′∈↑E′Km,m′
separates
µ−1(m)
from every
µ−1(m′)
,
m′∈ ↑E′
, which means that it separates
µ−1(m)
from
Sm′∈↑E′µ−1(m′) = L′
. Consequently,
K
separates
L
from
L′
. Additionally,
K
is a finite union of finite intersections
of languages in V, and since Vis a positive variety, K∈ V .
H. Proof of Lemma 16
Let
W
be a variety of ordered monoids such that
Pol(W) =
W.
Let
u, v ∈c
A∗
with
u≤Wv.
Suppose that there are
u′, u′′ ∈c
A∗
and
a∈A
such that
u=u′au′′
. We will show
that there exist
v, v′∈c
A∗
with
u′≤Wv′, u′′ ≤Wv′′
and
v=v′av′′.
As in previous proofs, we can enumerate all the morphisms
from
A∗
into members of
W
and, for each
j,
form the direct
product of the first
j
of these. The result is a sequence of
morphisms
ψi:A∗→Mi,
where each
Mi∈W,
with the following property: If
ϕ:A∗→
M∈W
is any morphism, then
ϕ
factors through
ψi
for all
sufficiently large
i.
(Note that we are using the fact that
W
is
closed under finite direct products.)
Since
u∈c
A∗, u = lim ui
for some convergent sequence of
finite words
ui.
Likewise, there are finite words
vi, u′
i, u′′
i
with
v= lim vi, u′= lim u′
i, u′′ = lim u′′
i.
Let
k > 0.
There is some
j0,
depending on
k,
such that if
j≥j0,then the values
ψk(uj), ψk(u′
j), ψk(u′′
j), ψk(vj)
are all constant independent of j. Now we define languages
L′
k=ψ−1
k(↑ψk(u′
j0))
L′′
k=ψ−1
k(↑ψk(u′′
j0))
Since these languages are pre-images of up-sets in
Mk,
they
belong to
W
. Let
Lk=L′
kaL′′
k.
Since
Pol(W) = W, Lk∈
W.
Let
µk:A∗→Nk∈W
be a morphism that recognizes
Lk.
We have
u′
jau′′
j∈L′
kaL′′
k
for
j≥j0.
There is thus some
jk> j0
such that if
j≥jk
then
µk(vj)
is constant independent
of j, and
µk(u′
jau′′
j) =
\
µk(u)≤
\
µk(v) = µk(vj).
This implies that
vj∈Lk,
and thus
vj=v′
jav′′
j
for some
v′
j∈L′
k, v′′
j∈L′′
k.
We thus have (by the definitions of
L′
k
and
L′′
k)
ψk(u′
j)≤ψk(v′
j), ψk(u′′
j)≤ψk(v′′
j)
whenever j≥jk.In this manner we obtain sequences
(u′
jk)k>0,(u′′
jk)k>0,(v′
jk)k>0,(v′′
jk)k>0
of finite words. By sequential compactness, we can extract
parallel convergent subsequences from
v′
jk
and
v′′
jk
converging
to limits
v′, v′′
respectively. Let’s label the terms of these sub-
sequences
V′
i, V ′′
i.
From Lemma 7 we have
u′≤Wv′, u′′ ≤W
v′′.
We also have, by Lemma 7
u=u′au′′ ≤Wv′av′′,
and from continuity of multiplication,
v′av′′ = (limiV′
i)a(limiV′′
i) = limiV′
iaV ′′
i.
Each
V′
i
and
V′′
i
has the form
v′
j, v′′
j
for some sequence of
indices jthat increases with i. Since v′
jav′′
j=vj,we get
v′av′′ = limjvj=v.
This completes the proof of the lemma in the case where
ℓ
is a
letter of
A.
We now consider the case
ℓ=ϵ,
so that
u=u′u′′.
We have
u′= lim u′
i
for some sequence of finite words
u′
i
. If
u′′ = 1
then we take
v′′ =v,
so we can assume that
u′′
is the limit of a sequence
(u′′
i)
of nonempty finite words.
Since the alphabet
A
is finite, there is some letter
a∈A
such
that infinitely many terms of the sequence have the form
aw′′
i,
where
w′′
i∈A∗.
By sequential compactness, the sequence
(w′′
i)
has a convergent subsequence
(w′′
ij)
converging to some
w∈c
A∗.So
u=u′u′′ = (limju′
ij)a(limjw′′
ij) = u′aw.
By the previous case
v=v′ax
for some profinite words
v′
and
x
with
u′≤Wv′
and
w≤Wx.
Set
v′′ =ax.
By
Lemma 7, we get
u′′ =aw ≤Wax =v′′.
I. Proof of Theorem 17
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
Finite word case.
The existence of a decomposition depth
δ:A∗→ {0,1, . . . , D}
for finite words is known as Si-
mon’s factorization forest theorem [30] (see also Colcombet’s
survey [8, Chap. 18]). In Simon’s theorem, words
u
with
idempotent type can be further written
u=u1· · · un
with
µ(u) = µ(u1) = · · · =µ(un)
an idempotent of
M
. To
recover the types listed in the present theorem, we simply
see the case
n= 2
as a binary type, and when
n > 2
, we set
u′:= u2· · · un−1
, since
µ(u′) = µ(u2)· · · µ(un−1) = µ(u)
by idempotency.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
General case.
We extend the aforementioned function
δ
into a function
d
for profinite words. For any
u∈c
A∗
we set
d(u)
to be the minimal value
k
such that there is a sequence
(ui)i≥0
of finite words that converges to
u
with
δ(ui)≤k
for
any
i
. (The existence of such a sequence is ensured by the
fact that
d
has finite range.) Note that the number of
ui
’s with
δ(ui)< k
has to be finite, otherwise the sequence of these
ui
’s would converge to
u
, contradicting the minimality of
d(u)
.
We can thus assume that all the ui’s are such that δ(ui) = k.
Let
u= (ui)i≥0∈c
A∗
with
δ(ui) = d(u)
for any
i
. We
show the theorem by induction on
d(u)
. If
d(u)=0
then
every
ui
is either a letter or the empty word, hence so is
u
. If
d(u) = k
, we address two cases depending of the type of the
ui
’s (note that these two cases overlap, and when this happens,
either case can be followed):
•
There are infinitely many
ui
’s with binary type. By
extraction of a subsequence, we can assume that all
ui
’s
are of binary type. For every
i
we can thus write
ui=
ui,1ui,2
with
δ(ui,1), δ(ui,2)< k
. By two applications of
Lemma
??
we can extract a further subsequence of the
ui
so that we can assume that both
(ui,1)i≥0
and
(ui,2)i≥0
are convergent. We write
u1
and
u2
for their respective
limits. This concludes the case since
d(u1), d(u2)< k
and u=u1u2by continuity of the product.
•
There are infinitely many
ui
’s with idempotent type. By
extraction of a subsequence, we can assume that all
ui
’s
are of idempotent type and that their images in
M
are
equal to the same idempotent
e∈M
. For every
i
, we
can write
ui=ui,1u′
iui,2
with
δ(ui,1), δ(ui,2)< k
, and
µ(ui,1) = µ(u′
i) = µ(ui,2) = e
. Again, by Lemma 6 we
can assume that
(ui,1)i≥0
,
(u′
i)i≥0
, and
(ui,2)i≥0
are all
convergent, and we write
u1
,
u′
, and
u2
for their respective
limits. This concludes the case since
d(u1), d(u2)< k
and
u=u1u′u2
by continuity of the product. By continuity
of µ, we also have bµ(u1) = bµ(u′) = bµ(u2) = bµ(u) = e.
J. Proof of Fact 1
Since
bπ
and
b
▷i
are continuous morphisms, we need only
show the claim for
u=x∈Mi
, that is, we need to show
x⪯
▷i(x)
. This is clear by construction, since
π
ℓ
(x) = π
ℓ
(▷i(x))
and πr(x)⊆πr(▷i(x)).
K. Proof of Fact 2
Since
bπ
ℓ
and
bπr
are continuous morphisms, we need only
show the claim for u=x∈Mi.
The claim holds for
M0
by construction. Let
i≥0
and
assume the claim holds for
i
. Let
x∈Mi+1
and let
w∈K+
i
be
such that
▷i(w) = x
. The claim at level
i
with
u:= w
asserts
that
π
ℓ
(w)∈πr(w)
. Fact 1 then asserts that
π(w)⪯▷i(w)
,
and so π
ℓ
(x) = π
ℓ
(w)∈πr(w)⊆πr(▷i(w)) = πr(x).
L. Proof of the first part of Corollary 20
We show that
M∞
is a monoid. There is an element
e=
(1M, E)∈K∞
, by construction. Take any other element
x=
(m, F )∈K∞
. Their product
ex
is
(m, E F )
, and since
1M∈
E
by Fact 2,
(m, F )⪯(m, E F )
. However,
ex ∈M∞
since
▷∞
is the identity map, and so there is an element
(m, F ′)
in
K∞
such that
(m, E F )⪯(m, F ′)
. As
K∞
contains only
incomparable elements,
F=EF =F′
, and so
ex =x
.
Similarly,
xe =x
, and thus
e
is a neutral element for
⟨K∞⟩
.
Note that
π
ℓ
is a monoid morphism, since
π
ℓ
(e)=1M
, and that
πr
is a relational morphism of monoids, since
1M∈πr(e) = E
.
M. Proof of Fact 3
Since
bπ
and
bursti
are continuous morphisms, we need only
show this over letters, so let us assume that
u=x∈Ki
.
By definition of
bursti(x)
, if
x
is not an idempotent then
the property holds. Otherwise,
bursti(x)
is the concatenation
of profinite words that all map to
xω+1 =x
via
bπ
, and so
bπ(bursti(x)) = x.
N. Proof of Fact 4
Since
b
▷i◦lifti
and
bπ
are continuous morphisms, we need
only show this over letters, so let us assume that
u=x∈Mi+1
.
This is then immediate since
▷i(lifti(x)) = x
by construction.
O. Proof of Fact 5
We have that
bπ
ℓ
(lifti−1(bursti(u))) = bπ
ℓ
(lifti−1(u))
by
Fact 3. Applying
bπ
ℓ
on both sides of the equation given by
Fact 4, we get that
π
ℓ
(b
▷i−1(lifti−1(u))) = bπ
ℓ
(u)
, and by
construction of
▷i−1
,
π
ℓ
(b
▷i−1(lifti−1(u))) = bπ
ℓ
(lifti−1(u))
.
This entails that
bπ
ℓ
(u) = bπ
ℓ
(lifti−1(bursti(u)))
, and in turn,
iterating the argument, that bπ
ℓ
(u) = bπ
ℓ
(τi(u)).
P. Proof of Fact 6
First, we show that the statement is closed under concatena-
tion over finite words, in the sense that if the claim is satisfied
for two finite words
u:= w1
and
u:= w2
then it will also be
satisfied for
u:= w1w2
. Let
m∈πr(▷i(w1w2))
. Since
πr◦▷i
is a morphism,
πr(▷i(w1w2)) = πr(▷i(w1))πr(▷i(w2))
, and
so
m=m1m2
with, for
b= 1,2
,
mb∈πr(▷i(wb))
. By
hypothesis, there are two words
um1, um2∈d
M+
such that,
for
b= 1,2
,
bπ
ℓ
(umb) = mb
and
τi(wb)≤Pol(V)umb
. This
implies that bπ
ℓ
(um1um2) = m1m2=mand:
τi(u) = τi(w1w2) = τi(w1)τi(w2)≤Pol(V)um1um2,
so that
um:= um1um2
, which is not
ε
by hypothesis, satisfies
the conclusion of the claim.
Next, we argue that showing the claim over finite words
implies that it holds over profinite words. Let
u= (wj)j≥0∈
d
K+
0
and
m∈πr(b
▷i(u))
. Since
πr◦▷i
is a morphism into a
finite semigroup, it is eventually constant over
(wj)j≥0
; we
can thus assume that
m∈πr(▷i(wj))
for any
j≥0
(taking a
subsequence of
(wj)j≥0
as needed). For any
j≥0
, applying
the lemma with
u:= wj
and
m
provides a profinite word
vj
such that bπ
ℓ
(vj) = mand τi(wj)≤Pol(V)vj.
The sequence
(vj)j≥0
has a convergent subsequence, by
compactness; let
um
be its limit. By continuity of
bπ
ℓ
, we
have that
bπ
ℓ
(um) = m
. Since
τi
is continuous,
(τi(wj))j≥0
converges to τi(u).
We are thus considering two sequences,
(τi(wj))j≥0
and
(vj)j≥0
, with
τi(wj)≤Pol(V)vj
for any
j≥0
, such that the
former sequence converges to
τi(u)
and the latter to
um
. By
Lemma 7, τi(u)≤Pol(V)um.
Q. Proof of Theorem 23
Consider one pass through the main loop of Algorithm 1.
Note that
|Exp| ∈ O(2n)
. The internal loop at line
??
takes
time
n2
, while the loop at line
??
is executed
n2×22n
times,
each time requiring to call the algorithm for
V-
pairs. All in
all, lines
??
to
??
take time
O(cV(2O(n)))
, since
cV
is at
least linear. As was shown in the proof of Lemma 19, the
set
↓MaxElts
always grows, hence the main loop will be
executed at most O(2n), showing the claim.
R. Proof of Theorem 29
We first note that Theorem 1 is an example of a more
general property, easily shown using De Morgan’s laws [24,
Lemma 33]: for any positive variety
V
,
Pol(Bool(V))
is equal
to the polynomial closure of the dual of
V
, that is, the set
co-Vof complements of languages in V. Hence:
chi+1
2(V) = Pol(co-chi−1
2(V)).
Consequently, if we inductively assume that we can decide
separation for
chi−1
2(V)
, we simply need to show that we
can decide separation for
co-chi−1
2(V)
with the same time
complexity in order to be able to apply Theorem 23 and
complete the proof. The time complexity bound given in the
statement holds inductively.
On the algebraic side, the ordered version of Eilenberg shows
that the variety of ordered monoids that corresponds to
co-V
,
for any positive variety of regular languages
V
, is the set
co-V
of ordered monoids of
V
in which the partial order is reversed.
This implies that for any profinite words
u, v
, we have that
u≤Vv
if, and only if,
v≤co-Vu
. Hence for any monoid
M
,
(m, m′)
is a
co-V-
pair of
M
if, and only if,
(m′, m)
is
a
V-
pair of
M
, so that computing
co-V-
pairs is as easy as
computing
V-
pairs. By Corollary 11, this means we can decide
co-V-separation, and we are done.