Automata Logics, and Infinite Games. A Guide To Current Research PDF
Automata Logics, and Infinite Games. A Guide To Current Research PDF
Automata Logics,
and Infinite Games
A Guide to Current Research
13
Volume Editors
Erich Grädel
RWTH Aachen, Mathematische Grundlagen der Informatik
52056 Aachen, Germany
E-mail: [email protected]
Wolfgang Thomas
RWTH Aachen, Lehrstuhl Informatik VII
52056 Aachen, Germany
E-mail: [email protected]
Thomas Wilke
Universität Kiel
Institut für Informatik und Praktische Mathematik
Christian-Albrechts-Platz 4, 24118 Kiel, Germany
E-mail: [email protected]
ISSN 0302-9743
ISBN 3-540-00388-6 Springer-Verlag Berlin Heidelberg New York
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, re-use of illustrations, recitation, broadcasting,
reproduction on microfilms or in any other way, and storage in data banks. Duplication of this publication
or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965,
in its current version, and permission for use must always be obtained from Springer-Verlag. Violations are
liable for prosecution under the German Copyright Law.
Springer-Verlag Berlin Heidelberg New York
a member of BertelsmannSpringer Science+Business Media GmbH
http://www.springer.de
over certain graphs of this kind, and in the final part the logical framework is
extended to guarded logics. Each part ends with notes with further references;
however, these pointers to the literature are not meant to be exhaustive.
The volume is the outcome of a research seminar which took place in Dagstuhl
in February 2001. There were 19 young researchers participating in the seminar;
each of them prepared a presentation based on one or several recent articles,
reshaping the material in a form with special emphasis on motivation, examples,
justification of constructions, and also exercises.
Thanks are due to the International Conference and Research Center of
Dagstuhl and the “Gesellschaft für Informatik (GI)” for the support it provided.
Achim Blumensath and Christof Löding provided substantial help in technical
and editorial matters; we express our sincere thanks to them.
The editors hope that this book will help many readers to enter this fasci-
nating, mathematically attractive, and promising area of theoretical computer
science. As an incentive, many open problems are mentioned in the text. The
best success which the book could have would be to guide readers to the solution
of some of these problems.
Aachen, Kiel, October 2002 Erich Grädel
Wolfgang Thomas
Thomas Wilke
Contents
Part I. Introduction
1 ω-Automata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Berndt Farwer
2 Infinite Games . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
René Mazala
3 Determinization of Büchi-Automata . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Markus Roggenbach
Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
1 ω-Automata
Berndt Farwer
Fachbereich Informatik
Universität Hamburg
1.1.1 Notation
The symbol ω is used to denote the set of non-negative integers, i.e. ω :=
{0, 1, 2, 3, . . . }.
By Σ we denote a finite alphabet. Symbols from a given alphabet are denoted
by a, b, c . . . . Σ ∗ is the set of finite words over Σ, while Σ ω denotes the set of
infinite words (or ω-words) over Σ (i.e. each word α ∈ Σ ω has length |α| = ω).
Letters u, v, w, . . . denote finite words, infinite words are denoted by small greek
letters α, β, γ . . . . We write α = α(0)α(1) . . . with α(i) ∈ Σ. Often we indicate
infinite runs of automata by , σ, . . . . A set of ω-words over a given alphabet is
called an ω-language.
For words α and w, the number of occurrences of the letter a in α and w is
denoted by |α|a and |w|a , respectively. Given an ω-word α ∈ Σ ω , let
Occ(α) = {a ∈ Σ | ∃i α(i) = a}
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 3-21, 2002.
Springer-Verlag Berlin Heidelberg 2002
4 Berndt Farwer
1.2 ω-Automata
(1) (0) = qI ,
(2) (i) ∈ δ((i − 1), ai ) for i ≥ 1 if A is nondeterministic,
(i) = δ((i − 1), ai ) for i ≥ 1 if A is deterministic.
1 ω-Automata 5
With the different acceptance conditions defined in the following sections the
question arises how they are related in expressive power, i.e. whether there exist
transformations from one acceptance condition to another. If such transforma-
tions can be established another question naturally arises: what is the complexity
for the respective translations?
The size of an automaton A, denoted by |A|, is measured by the number
of its states, i.e. for A = (Q, Σ, δ, qI , Acc) the size is |A| = |Q|. In addition to
the number of states of an automaton the size of the acceptance condition is
also of some importance for the efficiency of the transformation. This is usually
measured by the number of designated sets or pairs of such. Details are given in
the respective sections.
The Büchi acceptance condition has originally been introduced for nondetermin-
istic ω-automata. In this case, the acceptance component is a set of states.
Inf() ∩ F = ∅
i.e. at least one of the states in F has to be visited infinitely often during the
run. L(A) := {α ∈ Σ ω | A accepts α} is the ω-language recognized by A.
Example 1.4. Consider the ω-language L over the alphabet {a, b} defined by
a,b
q1 a
a
q0
b
a
q2 q3
a
Fig. 1.1. A Büchi automaton accepting the words from (a + b)∗ aω + (a + b)∗ (ab)ω
6 Berndt Farwer
i.e. the set of infinitely recurring states of is exactly one of the sets in F .
Example 1.7. Consider again the ω-language L over {a, b} consisting of the ω-
words which end with aω or with (ab)ω . The deterministic Muller automaton
of Figure 1.2 recognizes L, where the acceptance component consists of the two
sets {qa } and {qa , qb }.
We now verify that nondeterministic Büchi automata and nondeterministic
Muller automata are equivalent in expressive power.
One direction is straightforward: for a Büchi automaton A = (Q, Σ, δ, qI , F )
define the family F of sets of states by collecting all subsets of Q which contain
a state from F .
1 ω-Automata 7
b a
a
qb qa
b
Fig. 1.2. A state transition diagram where the state qa is reached after reading a and
qb after reading b.
and the set F of final states of A consists of the states (qG , ∅) for G ∈ F. We
do not give a formal definition of the transitions, which should be clear from the
description above.
If Q has n states and F contains m sets then |Q | has at most n + mn2n = 2O(n)
states. Summarizing, we obtain the following result.
b a
a
qb qa
b
c ba c
qc
c
Fig. 1.3. A state diagram where qx is reached after reading x.
Example 1.12. The Rabin automaton with state transition diagram from Fig-
ure 1.2 and Rabin condition Ω = {({qb }, {qa })} accepts all words that consist
of infinitely many a’s but only finitely many b’s.
To specify the language consisting of all words that contain infinitely many
b’s only if they also contain infinitely many a’s with a Rabin automaton based
on the state graph from Figure 1.2 we have to use Ω = {(∅, {qa }), ({qa , qb }, ∅)}.
This condition can be paraphrased by saying that each word in the accepted
language has either infinitely many a’s or it has neither infinitely many a’s nor
infinitely many b’s. It is clear that in the latter case no ω-word can be accepted
1 ω-Automata 9
and the condition could be simplified to Ω = {(∅, {qa })}. But in the presence of
a third symbol and a third state as depicted in Figure 1.3 two pairs are needed,
as the ω-word cω must be recognized: Ω = {(∅, {qa }), ({qa , qb }, {qc })}.
Example 1.14. Let Σ = {a, b}. The language consisting of all words that contain
infinitely many a’s if they contain infinitely many b’s can be recognized by a
Streett automaton with the state graph from Figure 1.2. The condition can be
paraphrased as |α|b = ω ∨ |α|a = ω, i.e. |α|b = ω ⇒ |α|a = ω. In the automaton
of Figure 1.2 the two states qa and qb indicate that respectively symbol a or b has
been read in the previous step. The appropriate Streett automaton is obtained
by taking as acceptance component the set Ω = {({qa }, {qb })}.
Rabin automata and Streett automata are transformed into Muller automata
by simply gathering all state sets that satisfy the Rabin condition, respectively
Streett condition, into a Muller acceptance set.
The parity condition amounts to the Rabin condition for the special case where
the accepting pairs (E1 , F1 ), . . . , (Em , Fm ) form a chain with respect to set inclu-
sion. We consider the case of an increasing chain E1 ⊂ F1 ⊂ E2 ⊂ . . . Em ⊂ Fm .
10 Berndt Farwer
Let us associate indices (called colours) with states as follows: states of E1 re-
ceive colour 1, states of F1 \ E1 receive colour 2, and so on with the rule that
states of Ei \ Fi−1 have colour 2i − 1 and states of Fi \ Ei have colour 2i. An ω-
word α is then accepted by the Rabin automaton iff the least colour occurring
infinitely often in a run on α is even (hence the term “parity condition”).
Sometimes it is more convenient to work with the condition that the maximal
colour occurring infinitely often in the run under consideration is even. This
applies to some constructions in later chapters of this book.
Example 1.17. Consider the parity automaton from Figure 1.4 with colouring
function c defined by c(qi ) = i.
a b
c
q0 a q1 b q2 q3
c
Fig. 1.4. Another ω-automaton
It accepts the ω-words with start with ab, continue by a finite sequence
of segments in a∗ cb∗ c, and end with aω ; so L(A) = ab(a∗ cb∗ c)∗ aω . For the
parity automaton A with the same transition graph but colouring c defined by
c (qi ) = i + 1 we obtain L(A ) = ab(a∗ cb∗ c)∗ bω ∨ ab(a∗ cb∗ c)ω .
It is obvious how a parity condition is cast into the form of a Rabin condition.
1.3.5 Discussion
(2) The ω-languages recognized by these ω-automata form the class ω-KC(REG),
i.e. the ω-Kleene closure of the class of regular languages.
Both questions can be answered affirmatively; and both involve tedious work.
The complementation problem can be attacked via several approaches (see
Chapter 4 below). One possibility is to work with deterministic ω-automata and
thus use a reduction to the determinization problem.
Let us see that Büchi automata are too weak to recognize even very simple
ω-languages from ω-REG. The Büchi automaton depicted in Figure 1.5 with
F = {q1 } accepts those ω-words over the alphabet {a, b} that have only finitely
many b’s.
a,b
a
q0 q1 a
Let us now show that deterministic Muller automata, Rabin automata, Streett
automata, and parity automata all have the same expressive power. We show first
the crucial step, namely that deterministic Muller automata can be transformed
into deterministic Rabin automata.
We use a technique called latest appearance record (LAR). The idea is
to use permutations of the states of the given Muller automaton as new states,
extended by a hit position. So the memory of the new automaton stores lists
of states from the original automaton; this is in contrast to the construction
of Theorem 1.10 which produced a nondeterministic Büchi automaton from a
Muller automaton; in that case we stored sets of states of the original automaton
in the memory of the constructed one.
In a list of (distinct) states, we use the last entry for the current state in the
run on the given Muller automaton. The hit position (the position of the marker
() indicates where the last change occurred in the record. For every transition
from one state p to q in the original automaton, the state q is moved to the last
position of the record while the symbols which were to the right of q are shifted
one position to the left (so the previous place of q is filled again). The marker is
inserted in front of the position where q was taken from. So the positions before
the marker are untouched by the transition under consideration.
The Muller automaton A accepts by the run if the set J considered in the
Lemma belongs to F . This means that the run will infinitely often visit a state
in the defined set Fk−j but only finitely often visit states u(v with |u| < k − j,
i.e. states from Ek−j . So the Rabin condition of A is satisfied and A accepts in
this case. The converse implication (“if A accepts an input word, then A does”)
is shown analogously.
From the definition of the sets Ej , Fj we see that they are arranged in a chain:
E1 ⊆ F1 ⊆ E2 . . . ⊆ Ek ⊆ Fk . We can shorten the chain by admitting only pairs
where Ej = Fj , without altering the set of accepting runs. Then we are left with
a strictly increasing chain of sets, and thus have defined an ω-automaton which
is presentable as a parity automaton.
14 Berndt Farwer
n
…
2
q0 1 q1 q2 … qn
The first lower bound will useful in Chapter 3, where a transformation from
Büchi automata to deterministic Rabin automata is presented, using the con-
struction of Safra [158]. The lower bound will show that Safra’s construction is
optimal.
The second lower bound is of interest in connection with the conversion of
Streett automata into Rabin automata (or conversely) presented above. The
lower bound result will be taken up again in Chapter 5, where Streett automata
are studied in more depth.
The proof idea of the present section is due to Michel [128]. We follow the
presentation as given by Löding in [114].
In order to keep the representation of nondeterministic automata small, a
set of initial states is used in the examples that follow. It is obvious that the
automata can be presented in the usual format by adding just one state and
adding arcs from this new state for each arc leaving an initial state of the given
automaton.
Example 1.28. Consider the family of Büchi automata from Figure 1.6. This
family of automata (An )n≥2 is defined over the alphabets {1, . . . , n, #} respec-
tively. (The constraint n ≥ 2 is introduced for the proof of Lemma 1.29 where
two different permutations of symbols from {1, . . . , n} are assumed to exist.)
The languages Ln accepted by these automata can be characterised by the
condition: A word α is accepted by An iff there exists k and i1 , . . . , ik ∈ {1, . . . , n}
such that each pair ij ij+1 for j < k and ik i1 appears infinitely often in α.
We encode the symbols 1, . . . , n by words over {0, 1}∗ such that
0i 1 if i < n,
i is encoded by
0i 0∗ 1 if i = n
furthermore we keep # unaltered. Now we can specify the same family of lan-
guages w.r.t. the encoding by the family of automata (An )n≥2 over the fixed
alphabet {0, 1, #}. The size of An (in either of the two versions) is O(n).
The family of automata from the previous example can be used to prove the
following lemma.
1 ω-Automata 17
v R
u
q0 q p
w
S
Lemma 1.29. There exists a family of languages (Ln )n≥2 over the alphabet
{0, 1, #} recognizable by nondeterministic Büchi automata of size O(n) such
that any nondeterministic Streett automaton accepting the complement language
of Ln has at least n! states.
n n n'
n n
…
1 1
1 1 1'
Theorem 1.30. There exists a family of languages (Ln )n≥2 over the alphabet
{0, 1, #} recognizable by nondeterministic Büchi automata of size O(n) such that
any equivalent deterministic Rabin automaton must be of size n! or larger.
The technique of latest appearance records is used for the transformation of var-
ious automata into parity automata. Two variants are studied in the literature:
state appearance records and index appearance records.
State appearance records have been introduced in Section 1.4.2 for the trans-
formation of Muller automata into Rabin automata. Löding [114] shows that any
transformation of a deterministic Streett automaton of size n with r pairs of des-
ignated sets into a deterministic Rabin automaton will result in an automaton
where the number of states is a factorial in min(n, r), and by the fact that par-
ity automata are special cases of Rabin automata, a transformation to a parity
condition will result in an automaton with at least min(n, r)! states and O(r)
colours. Since the automata used in the proof consist of n states and n pairs,
this also proves the optimality of the best known transformation from Muller
automata to automata with parity condition.
Due to the duality of Rabin and Streett conditions the result is transferrable
to the case with Rabin condition and Streett condition interchanged.
Index appearance records (IAR) are used for example by Safra [159] for the
transformation of nondeterministic Streett automata into deterministic Rabin
automata. The transformation, to be presented in full detail in Chapter 5 below,
takes a deterministic Streett automaton of size n with an acceptance condition
consisting of r pairs of designated sets to an equivalent deterministic Rabin
automaton of size nO(r)! that uses O(r) accepting pairs.
By Theorem 1.32 we obtain the optimality of the IAR construction. The
following example gives the family of automata on which the proof is based.
Example 1.31. Consider the family of deterministic Streett automata (An )n≥2
from Figure 1.8 with pairs of designated state sets Ωn = {(E1 , F1 ), . . . , (En , Fn )}
and Ei = {i}, Fi = {i }.
The language accepted by the automaton An can be characterised by the
symbols occurring in odd and even positions of the accepted words. Each word
1 ω-Automata 19
α in L(An ) satisfies the condition that each symbol occurring infinitely often in
an odd position must also occur infinitely often in an even position of α.
This family of automata (An )n≥2 is defined over the alphabets {1, . . . , n},
respectively. By encoding the symbols 1, . . . , n by words over {0, 1}∗ such that
0i 1 if i < n,
i is encoded by i ∗
0 0 1 if i = n
we can specify the same family of languages w.r.t. the encoding by the family of
automata (An )n≥2 over the fixed alphabet {0, 1}. The construction is similar to
that in Section 1.5.1.
Theorem 1.32 ([114]). There exists a family of languages (Ln )n≥2 over the al-
phabet {0, 1} recognizable by deterministic Streett automata with O(n) states and
O(n) pairs of designated state sets such that any deterministic Rabin automaton
accepting Ln requires at least n! states.
Proof. The idea for proving Theorem 1.32 is motivated by the observation that
for any finite word u ∈ {1, . . . , n}∗ of even length, the word uα is accepted
by An iff α is accepted by An . It can be shown by induction over n that any
deterministic Rabin automaton accepting L(An ) must have at least n! states.
The base case for the induction is obvious: Any (Rabin) automaton accepting
a proper subset of the infinite words over a 2-letter alphabet with some word
having occurrences of both letters needs at least two states.
The induction step relies on the fact that any given deterministic Rabin
automaton A accepting L(An ) can be modified to a deterministic automaton
over {1, . . . , n} \ {i} for any i ∈ {1, . . . , n} by simply removing all arcs labelled
by i. Setting the initial state of the modified automaton to any q that is reachable
in An by an even number of state transitions we obtain a deterministic Rabin
automaton Aqi .
Because of the characterisation of L(An ) given above, it is clear that Aqi
accepts a language isomorphic up to the renaming of symbols to L(An−1 ). The
induction hypothesis requires the automaton Aqi to have at least (n − 1)! states.
For a complete proof the reader is referred to [114].
There are different possibilities to use the set Occ() for acceptance. The
analogue to the Muller condition, introduced by Staiger and Wagner [166], uses
a family F of state sets and declares the run accepting if
Occ() ∈ F .
Occ() ∩ F = ∅,
Occ() ⊆ F,
by an occurrence condition. For example, the set L of ω-words over {a, b} with
infinitely many b is not recognizable by an ω-automaton with Staiger-Wagner
acceptance. Assuming such an automaton which recognizes L, say with n states,
one would consider an accepting run on the input word (an+1 b)ω . After some
finite prefix, say after (an+1 b)k , the run would have visited the states which
are visited at all. In the succeeding block an+1 the automaton assumes a loop,
which can be repeated if the input is changed to (an+1 b)k aω . So over this input
the same states would be visited as in the considered run over (an+1 b)ω . Hence
(an+1 b)k aω would be accepted, a contradiction.
1.7 Conclusion
We have shown the expressive equivalence of
n+kn·2n NB
DM DR DS
n·2O(k log k) ,O(k)
2O(n log n) ,n
DP
René Mazala
2.1 Introduction
2.2 Games
2.2.1 Arenas
An arena is a triple
A = (V0 , V1 , E) (2.1)
2.2.2 Plays
A play of a game with an arena as above may be imagined in the following way:
a token is placed on some initial vertex v ∈ V . If v is a 0-vertex then Player 0
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 23-38, 2002.
Springer-Verlag Berlin Heidelberg 2002
24 René Mazala
is then called a game, where A is the arena of the game and Win its winning
set. The plays of that game are the plays in the arena A. Player 0 is declared
the winner of a play π in the game G iff
We will only be interested in winning sets that can be described using the ac-
ceptance conditions that were discussed in the previous chapter. But recall that
these acceptance conditions made only sense when used with automata with a
finite state space—a run of an infinite-state automaton might have no recurring
state. We will therefore colour the vertices of an arena and apply the acceptance
conditions from the previous chapter on colour sequences.
Let A be as above and assume χ : V → C is some function mapping the
vertices of the arena to a finite set C of so-called colours; such a function will
be called a colouring function. The colouring function is extended to plays in
a straightforward way. When π = v0 v1 · · · is a play, then its colouring, χ(π), is
given by χ(π) = χ(v0 )χ(v1 )χ(v2 ) · · · . So, when C is viewed as the state set of a
2 Infinite Games 25
finite ω-automaton and Acc is an acceptance condition for this automaton (in
the sense of the previous chapter), then we will write Wχ (Acc) for the winning
set consisting of all infinite plays π where χ(π) is accepted according to Acc.
Depending on the actual acceptance condition we are interested in, this means
the following, where π stands for any element of V ω .
P
• Muller condition (Acc = F ⊆ 0 (C)): π ∈ Wχ (Acc) iff Inf(χ(π)) ∈ F .
• Rabin condition (Acc = {(E0 , F0 ), (E1 , F1 ), . . . , (Em−1 , Fm−1 )}):
π ∈ Wχ (Acc) iff ∃k ∈ [m] such that Inf(χ(π)) ∩ Ek = ∅ and Inf(χ(π)) ∩ Fk =
∅,
• Streett condition (Acc = {(E0 , F0 ), (E1 , F1 ), . . . , (Em−1 , Fm−1 )}):
π ∈ Wχ (Acc) iff ∀k ∈ [m].(Inf(χ(π)) ∩ Ek = ∅ ∨ Inf(χ(π)) ∩ Fk = ∅),
• Rabin chain condition (Acc = {(E0 , F0 ), (E1 , F1 ), . . . , (Em−1 , Fm−1 )} where
E0 ⊂ F0 ⊂ E1 ⊂ F1 ⊂ . . . ⊂ Em−1 ⊂ Fm−1 ): like the Rabin condition.
• Parity conditions (the colour set C is a finite subset of the integers):
– max-parity condition: π ∈ Wχ (Acc) iff max(Inf(χ(π))) is even.
– min-parity condition: π ∈ Wχ (Acc) iff min(Inf(χ(π))) is even.
• Büchi condition (Acc = F ⊆ C): π ∈ Wχ (Acc) iff Inf(χ(π)) ∩ F = ∅.
• 1-winning (Acc = F ⊆ C): π ∈ Wχ (Acc) iff Occ(χ(π)) ∩ F = ∅.
For simplicity, we will just write (A, χ, Acc) instead of (A, Wχ (Acc)). To indicate
that we are working with a certain acceptance/winning condition, we will speak
of Muller, Büchi, . . . games. We will say a game is a regular game if its
winning set is equal to Wχ (Acc) for some χ and some acceptance condition Acc
from above, except for 1-acceptance.
Example 2.1. Let A = (V0 , V1 , E, χ) be the (coloured) arena presented in Fig-
ure 2.1. We have the 0-vertices V0 = {z1 , z2 , z5 , z6 } (circles) and the 1-vertices
V1 = {z0 , z3 , z4 } (squares). The colours are C = {1, 2, 3, 4}. The edge relation E
and the colour mapping χ may be derived from the picture, i.e. χ(z4 ) = 2
or χ(z0 ) = 1. Note that we don’t have a dead end in our example. As a
winning condition we choose the Muller acceptance condition given by F =
{{1, 2}, {1, 2, 3, 4}}.
A possible infinite play in this game is π = z6 z3 z2 z4 z2 z4 z6 z5 (z2 z4 )ω . This
play is winning for Player 0 because χ(π) = 23121224(12)ω and Inf(χ(π)) =
{1, 2} ∈ F . The play π = (z2 z4 z6 z3 )ω yields χ(π ) = (1223)ω and Inf(χ(π )) =
/ F . Hence π is winning for Player 1.
{1, 2, 3} ∈
When we want to fix a vertex where all plays we consider should start, we
add this vertex to the game: an initialized game is a tuple (G, vI ) where vI is
a vertex of the arena of G. A play of such a game is a play of the uninitialized
game which starts in vI .
1 2
z0 z1
1 z2
3 z3 z4 2 z5 4
z6
First, it would be interesting to know if one of the players can play in such a
game that regardless of how the other moves, the emerging plays will be wins for
him. This is the question whether the game is “determined”. We will formalize
this by introducing the notions of strategy and winning strategy, and we will
state the fundamental result that every regular game is determined; the result
itself will be proved in Chapter 6.
Second, when we consider games on finite graphs these can be input for an
algorithm and an obvious question to ask is if one can effectively (and maybe
efficiently) determine which of the two players wins the game. This question will
be answered in Chapter 7; the complexity of determining the winner heavily
depends on the type of the game (the winning condition) one is interested in.
Third, it is not only interesting to know who wins a game, but also how
a winning strategy looks like. Clearly, a winning strategy will tell the player
what to do next depending on the moves that have been taken thus far. We
will be interested in situations where the winning strategies are simple in the
sense that the next move of the player does only depend on the current vertex
or on a bounded amount of information on the moves that led to the current
position—we will be interested in “memoryless” or “forgetful” strategies. We
will describe this formally and state the main result that for every regular game
there is a forgetful winning strategy and that parity games even allow memoryless
strategies.
In order to be able to define formally what it means for a player to win a game,
we need to introduce the notion of strategy.
2 Infinite Games 27
2.4.1 Strategies
Example 2.3. In the game from Examples 2.1 and 2.2, Player 1 wins on {z0 , z1 }
whereas Player 0 wins on {z2 , z3 , z4 , z5 , z6 }.
28 René Mazala
Remark 2.4. For any game G, if Player 0 wins G on U0 and Player 1 wins G on
U1 , then U0 ∩ U1 = ∅.
Given a game G, we define the winning region for Player σ, denoted Wσ (G)
or Wσ for short, to be the set of all vertices v such that Player 0 wins (G, v).
Clearly:
Exercise 2.2. Proof the above remark by showing that if U is a family of sets of
vertices
such that Player σ wins on each element U ∈ U, then Player σ wins on
U.
U ∈U
Remark 2.6. For every regular game (A, χ, Acc) there exists a Muller winning
condition Acc such that (A, χ, Acc) and (A, χ, Acc ) have the same winning
regions.
The main result says that it is enough to consider parity games. Therefore,
parity games are of our interest in the whole volume.
Theorem 2.7. For every Muller game (A, χ, F ) there exists a parity game
(A , χ , Acc ) and a function r : V → V such that for every v ∈ V , Player σ
wins ((A, χ, F ), v) if and only if Player σ wins ((A , χ , Acc ), r(v)).
is the set of all words w over the alphabet C ∪ {} where and at least one
C
colour are infixes of w and each colour appears at most once.
Now we can define our game (A , χ , Acc ). As vertices we choose
and V1 := V1 × C
V := V0 ∪ V1 with V0 := V0 × C . (2.5)
The set of edges is given by
E := ((v, q), (v , ϕ(v , q))) v ∈ V, v ∈ vE, q ∈ C (2.6)
where ϕ : V × C→C is the memory update function that deletes the marker,
replaces the colour c := χ(v ) of the given vertex v by the marker and finally
appends c. Formally, ϕ is defined as
xyzc if q = xcyz
ϕ(v , q) := xyzc if q = xycz (2.7)
qc else (c is not an infix of q)
for each v ∈ V and each q ∈ C with c := χ(v ). The function that transforms
the initial vertex can be set to
r(v) := (v, χ(v)) . (2.8)
The new colouring function χ : V → ω is defined by
2 ∗ |y| − 1 if { c ∈ C | c infix of y } ∈
/F
χ (v, xy) := . (2.9)
2 ∗ |y| otherwise
a a b
z0 z1 z2
Example 2.8. Let A be the arena in Figure 2.2, and F = {{b}} a Muller accep-
tance condition. The example play π = z1 z2 z0 z1 z2ω is winning for Player 0. The
winning regions are W0 = {z2 } and W1 = {z0 , z1 }. The constructed max-parity
game A is presented in Figure 2.3. We get
with the colouring χ (π ) = 133132ω which is winning for Player 0 too. The
winning region W0 for Player 0 is the set of all vertices with z2 in the first
component. W1 is the complement of W0 .
1 1 3 2
z0 , a z1 , a z2 , ab z2 , ab
1 1 3 2
2.4.3 Determinacy
In all of our example games, the winning regions for Player 0 and Player 1
partition the set of vertices of the game. When a game has this property, we will
say it is determined.
Martin (see, e. g., [119], [95]) showed that every game with a Borel type
winning set is determined. In Chapter 6, we will show the following special case
of Martin’s theorem.
In Example 2.2, we already stated that Player 0 must not move from z6 to
the same successor every time he visits z6 . So, Player 0 can’t have a memoryless
winning strategy.
We say that Player σ wins a game G forgetful when he has a forgetful
strategy for each point of his winning region. Accordingly, it is defined what it
means to win with finite memory, memoryless, and positional.
Exercise 2.3. Give an example for a game G such that Player 0 wins forgetful
on each {v} for v ∈ W0 , but he has no forgetful winning strategy on W0 . Can
you give an example where G is regular?
In exercise 2.2, the reader was asked to show that if U is some set of vertices
such that Player σ wins a given game G on every element of U , then he wins
G on U . This is easy to see. In Exercise 2.3, the reader is asked to provide an
example that shows that the corresponding statement is not true for forgetful
strategies. However, a corresponding statement is true for memoryless strategies
under a certain condition:
Lemma 2.12. Let G = (A, Win) be any game with countable vertex set V ,
V ∗ Win ⊆ Win and Win/V ∗ ⊆ Win, (2.13)
∗ ∗
where Win/V := { η ∈ V | ∃w ∈ V with wη ∈ Win } is the set of all suffixes
ω
of Win. Let U be a set of vertices such that Player σ has a memoryless winning
strategy for each element from U . Then Player σ has a memoryless winning
strategy on U .
Before we turn to the proof observe that the two conditions on the winning set
are satisfied in every regular game: A prefix of a winning play can be substituted
by any other finite word; the set of infinitely often visited colours stays the same.
Proof. The proof uses the axiom of choice. For every u ∈ U , let fσu : Vσ → V
be a partial function which is a memoryless winning strategy for Player σ on u.
Without loss of generality, we assume that for every u ∈ U the domain of fσu ,
denoted Du , is minimal with respect to set inclusion.
Let< be a well-ordering on U (therefore we choose V to be countable) and
D := Du . We have to define a memoryless winning strategy fσ : D → V .
u∈U
For each v ∈ D, let u(v) be the minimal vertex in U (with respect to the
u(v)
well-ordering) with v ∈ Du(v) , and set fσ (v) := fσ (v). Clearly, fσ is well
defined and memoryless. We have to show that fσ is a winning strategy on U .
Assume π = v0 v1 · · · is a play starting in U and conform with fσ . In each
u(v )
σ-vertex vj of the play π, Player σ has to choose the strategy fσ j . Let i
be such that u(vi ) is minimal (with respect to the well-ordering) in the set
{ u(vj ) | j ∈ ω and vj ∈ D }. Then, from this moment i on, the strategy fσ
u(v )
follows the strategy fσ i . The domain Du(vi ) was minimal with respect to
set inclusion, thus, the play vi vi+1 · · · is a suffix of a play that starts in u(vi ),
visits vi , and is conform to fσ i . Hence, π ∈ V ∗ (Win/V ∗ ) ⊆ Win by our two
u(v )
2 1 1 2
z0 z1 z2 z3
Example 2.13. Let A be the max-parity game in Figure 2.4. Clearly, Player 0
wins on each v ∈ U = {z1 , z2 }, i. e. with the memoryless strategies
• f0z1 (z1 ) = z2 and f0z1 (z2 ) = z3 ,
• f0z2 (z2 ) = z1 and f0z2 (z1 ) = z0 .
To find a memoryless strategy on U , Player 0 can not set f0 (z1 ) = f0z1 (z1 ) and
f0 (z2 ) = f0z2 (z2 ) because this yields an infinite loop in z1 and z2 which is a loss
for him. If z1 < z2 in the well-ordering of U , then we get f0 ≡ f0z1 . This is a
memoryless winning strategy on U .
Theorem 2.14. In every parity game, both players win memoryless. This is
called memoryless determinacy of parity games.
From this, together with the construction in the proof of Theorem 2.7, we
can conclude:
Corollary 2.15. In every regular game, both players win forgetful. Analogously,
this is called forgetful or finite memory determinacy of regular games.
Note that the initial memory state in the previous construction could be
chosen arbitrarily.
34 René Mazala
Exercise 2.4. Using the results from the previous chapter, determine how much
memory is sufficient and necessary to win Rabin and Muller games.
Theorem 2.14 states that parity games enjoy memoryless determinacy, that
is, winning strategies for both players can be chosen memoryless. It is easy to
show that in certain Muller games both players need memory to win. In between,
we have Rabin and Streett conditions. For those, one can actually prove that
one of the two players always has a memoryless winning strategy, but we will
not carry out the proof in this volume.
Theorem 2.16. In every Rabin game, Player 0 has a memoryless winning strat-
egy on his winning region. Symmetrically, in every Streett game, Player 1 has a
memoryless strategy on his winning region.
This theorem can also be applied to certain Muller automata on the grounds
of the following observation. A Muller condition (F0 , F1 ) can be rephrased as
Rabin condition if and only if F1 is closed under union.
Example 2.17. We got a memoryless strategy for Player 1 in our Example 2.11.
His winning condition F1 is expressible as Rabin condition:
{({3}, {4}), ({4}, {3}), ({1}, {2})}. He wins a play if it loops, for instance, finitely
often through one of the colours 3 or 4 and infinitely often through the other
colour. Note that the winning condition cannot be rephrased as a parity condi-
tion, that is, Rabin chain condition (on the same graph).
Proof. The proof is constructive in the sense that on finite graphs it can be
immediately turned into an algorithm which computes the winning regions and
the memoryless winning strategies.
2 Infinite Games 35
P P
R(A, X) and a memoryless winning strategy for Player 0 are defined inductively.
In the inductive step, we use the function pre : (V ) → (V ) defined by
pre(Y ) = { v ∈ V0 | vE ∩ Y = ∅ } ∪ { v ∈ V1 | vE ⊆ Y } (2.15)
Inductively, we set X 0 = X,
for each limit ordinal ξ. Let ξ be the smallest ordinal such that X ξ = X ξ+1 . We
claim that W := X ξ is Player 0’s winning region. Clearly, for every v ∈ W \ X
there exists a unique ordinal ξv < ξ such that v ∈ X ξv +1 \ X ξv . By the above
definition, we furthermore know that for every v ∈ W ∩V0 \X there exists v ∈ vE
such that v ∈ X ξv . We set f0 (v) = v and claim that f0 is a memoryless strategy
for Player 0 on W . This can be easily proved by transfinite induction: One shows
that f0 is winning for Player 0 on X ν for every ν ≤ ξ. Hence, W ⊆ W0 .
On the other hand, let W = V \ W and assume v ∈ W . Then v ∈ / X. If
v is a dead end, it must be a dead end of Player 0 because all dead ends of
Player 1 belong to X 1 . But, on a dead end belonging to Player 0, Player 1 wins
immediately. If v is no dead end and belongs to V0 , we have v ∈ / W for every
v ∈ vE because otherwise v would belong to W . Similarly, if v is no dead end
and belongs to V1 , there exists v ∈ vE such that v ∈ / W because otherwise
v would belong to W . If we set f1 (v) = v in this case, then f1 is clearly a
memoryless strategy for Player 1. Every play conform with this strategy and
starting in W has the property that all its vertices belong to W . Since W
does not contain vertices from X or dead ends of Player 1 this play must be
winning for Player 1. Hence, f1 is a winning strategy for Player 1 on W and
V \ W = W ⊆ W1 , that is, W0 = W and W1 = V \ W .
aX (U ) := X ∪ pre(U ) . (2.18)
Show that aX is monotone with respect to set inclusion and that Attr0 (A, X) is
the least fixed point of aX .
36 René Mazala
Exercise 2.6. Show that in a finite arena with n vertices and m edges the at-
tractor of any set can be computed in time O(m + n).
The above remark tells us that, without loss of generality, we can assume
that arenas have no dead ends. Let (A, Acc) be an arbitrary game with A =
(V0 , V1 , E). For σ ∈ {0, 1}, we set Uσ = Attrσ (A, ∅). Then Player σ wins (A, Acc)
on Uσ memoryless. Now, let V0 = V0 \ (U0 ∪ U1 ) and V1 = V1 \ (U0 ∪ U1 ) and
consider the arena A = (V0 , V1 , E ∩ ((V0 ∪ V1 ) × (V0 ∪ V1 ))). Clearly, A does
not have any dead end. Further, for every v ∈ V0 ∪ V1 , Player 0 wins (A , Acc, v)
iff he wins (A, Acc, v) and, symmetrically, Player 1 wins (A , Acc, v) iff he wins
(A, Acc, v). More specifically, winning strategies for (A , Acc) can be used in
(A, Acc).
z0
z1 z2
Example 2.20. In the game depicted in Figure 2.5, Player 1 may prevent an
infinite play by moving the token to z2 . This is a dead end for Player 0 and
Player 1 wins.
2.5.2 1-acceptance
Using what we have proved about reachability games, we can now easily solve
1-games.
Exercise 2.9. Show that for finite arenas, the winning regions of 1-games can be
computed in time O(m + n). (See also Exercise 2.6.)
Proof. Like in the other solutions, we first describe how to construct the winning
region for Player 0 in a Büchi game (A, χ, F ).
We set Y = χ−1 (F ), and define inductively:
Z0 = V , (2.19)
X ξ = Attr0 (A, Z ξ ) , (2.20)
ξ ξ
Y = pre(X ) , (2.21)
Z =Y ∩Y ,
ξ+1 ξ
(2.22)
ξ
Z = Zν , (2.23)
ν<ξ
where the last equation only applies to limit ordinals ξ. Let ξ be the least ordinal
≥ 1 such that Z ξ = Z ξ+1 . We claim W := Attr0 (A, Z ξ ) is the winning region of
Player 0.
To prove W ⊆ W0 , we describe a memoryless winning strategy f0 for Player 0
on W . For every v ∈ V0 ∩ Z ξ , there exists v ∈ vE ∩ Attr0 (A, Z ξ ) and we set
f0 (v) = v . For every other v ∈ V0 ∩ W , we know v ∈ Attr0 (A, Z ξ ), and thus
we set f0 (v) to the value of a respective attractor strategy. Now, the following is
38 René Mazala
easy to see. First, if a finite play starting in W is conform with f0 , then it ends
in a dead of Player 1, which means Player 0 wins. Second, if an infinite play
starting in W is conform with f0 it eventually reaches Z ξ and from this point
onwards it will reach Z ξ over and over again. But since Z ξ ⊆ Y (this is because
ξ ≥ 1), the play will be winning for Player 0.
To prove that W0 = W , we argue that Player 1 has a memoryless winning
strategy on W := V \ W . The winning strategy is defined as follows. For every
v ∈ W there exists a least ν such that v ∈ X ν \ X ν+1 . (Note that X 0 = V and
X ν ⊆ X ν for all ordinals ν and ν with ν < ν .) Since X ν+1 is a 0-attractor,
V \X ν+1 is a 0-trap. We set f1 (v) to the value of a trapping strategy for Player 1
if v ∈ Y . Otherwise, it follows that v ∈ pre(X ν ), and thus, there exists some
v ∈ vE ∩ V \ X ν . We set f1 (v) = v . By induction on ν, it is now easy to show
that f1 is a winning strategy for Player 1 on V \ X ν . It follows that f1 is a
winning strategy on W .
Exercise 2.10. Show that for a finite arena, the winning regions of a Büchi game
can be computed in time O(n(m + n)).
3 Determinization of Büchi-Automata
Markus Roggenbach
For Bene
Introduction
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 43-60, 2002.
Springer-Verlag Berlin Heidelberg 2002
44 Markus Roggenbach
The above picture also shows how determinization can be used for comple-
mentation: given a nondeterministic Büchi automaton accepting a language L,
use Safra’s construction to obtain an equivalent deterministic Muller automa-
ton with state set Q and system F of final state sets. With 2Q \F as system
of final state sets, this automaton accepts the complement of L. Applying the
construction of section 1.3.2 results in a Büchi automaton for the complement
of L.
Another application of Safra’s construction can be found in Klarlund, Mu-
kund and Sohoni [99]: they generalize the construction to asynchronous finite
automata accepting infinite Mazurkiewicz traces.
This chapter is organized as follows: section 3.1 shows that the inclusion
of the deterministic variant in the nondeterministic variant is strict for Büchi
automata. Then Safra’s construction is discussed and proven to be correct in
section 3.2. Finally, section 3.3 deals with the optimality of Safra’s construction.
Example 3.1. Consider the Büchi automaton A = ({qI , f }, {a, b}, ∆, qI, {f }),
where
∆ = {(qI , a, qI ), (qI , b, qI ), (qI , a, f ), (f, a, f )}
(c.f. Figure 3.11 ). This automaton A accepts the language
F = { {qI , f } }, which would also accept the word (ab)ω ∈ / L. The problem with
the corresponding run = {qI }{qI , f }{qI}{qI , f } . . . is that – although there
are infinitely many macrostates containing f – we cannot extract a run of A
from exhibiting infinitely many f : at any time when we could choose f from
a macrostate in , there is no transition with label b available in A.
Note that with
a,b a
a
qI f
b a
a
{qI } {qI , f }
b
Proof. Let Σ := {a, b}. Consider again the language L := {α ∈ Σ ω |
b (α) < ∞}.
As shown in the above example, L is accepted by a nondeterministic Büchi
automaton.
Assume that there is a deterministic Büchi automaton A = (Q, Σ, δ, qI , F )
accepting L. This automaton accepts all words of the form σaω , where σ ∈ Σ ∗ .
Consider a reachable state q ∈ Q. Any finite word σq leading in A from the
initial state qI to q can be extended to an infinite word σq aω ∈ L, i.e. some
state f ∈ F occurs infinitely often in its run on A. Thus there must be a finite
sequence of a-transitions from q to a recurring state.
46 Markus Roggenbach
1
qI q1 1,
1
q
1,
Before presenting the construction in detail, we discuss Safra’s tricks for extend-
ing the powerset construction in an informal way:
Exercise 3.2. Apply Trick 1 on the Büchi automaton of example 3.1. How
does it prevent the run of (ab)ω to be accepting?
M with M ∩ F = ∅ a {q ∈ Q | (m, a, q) ∈ ∆, m ∈ M },
{q ∈ Q | (m, a, q) ∈ ∆, m ∈ F ∩ M }
{. . . , q1 , . . . } {. . . , qn+1 , . . . },
{qn+1 }, {qn+1 }
a1 an
{. . . , q2 , q2 , . . . } {. . . , qn , qn , . . . },
{qn }, {qn }
a2 an−1
.. ..
. .
ai−2 aj
{. . . , qj , qj , . . . },
{. . . , f, qi−1 , . . . }
{qj }, {qj }
ai−1 aj−1
··· M ,
M
K1 , . . . , Kk
(1) The initial state qI is the Safra tree consisting of the single node 1 labelled
with macrostate {qI }.
(2) The value of the transition function δ(T, a) for a given input a ∈ Σ and a
Safra tree T with a set N of nodes is computed as follows:
Step 1: Remove all marks ‘!’ in the Safra tree T.
Step 2: For every node v with macrostate M such that M ∩ F = ∅, create
a new node v ∈ (V \N ), such that v becomes the youngest son of v and
carries the macrostate M ∩ F.
Step 3: Apply the powerset construction on every node v, i.e. replace its
macrostate M by {q ∈ Q | ∃(m, a, q) ∈ ∆ : m ∈ M }.
Step 4 (horizontal merge): For every node v with macrostate M and
state q ∈ M, such that q also belongs to an older brother of v, remove q
from M.
Step 5: Remove all nodes with empty macrostates.
Step 6 (vertical merge): For every node whose label is equal to the union
of the labels of its sons, remove all the descendants of v and mark v with
‘!’.
(3) The set of states Q consists of all reachable Safra trees.
We should check first, that – given a Safra tree T and an input symbol a
– the transition function δ computes indeed a Safra tree. This ensures that Q
consists of Safra trees, as the initial state qI is obviously a Safra tree.
Removing the marks ‘!’ from all nodes of a Safra tree T does not violate
Condition 1 or Condition 2, and as all macrostates are nonempty in T, they are
also nonempty after Step 1. Thus Step 1 preserves the Safra tree properties.
Applying Step 2 on a Safra tree T with a node v carrying a macrostate
M ⊆ F, yields a tree violating Condition 1, as v and its youngest son carry
afterwards the same label M. Computing new macrostates for all nodes of a tree
in Step 3 may lead to even more trouble:
Step 4, Step 5, and Step 6 deal with these problems, resp.: Step 4 ensures Con-
dition 2 by horizontal merge of brother macrostates. Step 5 removes nodes with
empty macrostates. By vertical merge Step 6 fixes situations where Condition 1
is violated. Thus, we finally obtain after all six steps a Safra tree.
a,b c
a
a
qI f g a
a
1 − {qI , f } 1 − {qI , f } 1 − {f } 1 − {f } 1 − {f } 1 − {f } - !
2 − {f } 2 − {f } 2 − {f } 2 − {f }
As the above example indicates, the constructed Rabin and Muller conditions
are not ‘minimal’. For Muller conditions the following restriction might lead to
a smaller a system of final state sets (the proof is left for Exercise 3.5):
Computing δ( 1 − {qI , f, g} , a) :
↓
2 − {g, f }−!
4 − {g} 4 − {f, g}
2 − {f, g} 2 − {f, g} - !
2 − {f, g} 3−∅
4 − {f, g}
4 − {f, g}
Exercise 3.5. Prove Remark 3.5. Is it possible to generalize this result to arbi-
trary Muller conditions?
a,b,c
c
1 − {qI }
b b
b a b a
a
c a
1 − {qI , f } 1 − {f }−! 1 − {g}−! 1 − {f, g}−!
b c
c
c
a
b
1 − {qI , f, g}
↓ c
2 − {g}
1 − {qI , f, g}
↓
2 − {g, f }−!
a
Proof. Let α ∈ L(B). As M and R are deterministic and have the same initial
state and transition relation there exists one run on α of both automata. We
claim that there is at least one node v in the Safra trees of such that
Concerning the Muller condition this proves that Inf( ) equals a set in the
system F of final states: condition Muller 1 is true because a Safra tree of not
including v is not in Inf( ). As v is marked ‘!’ infinitely often in and Q is a
finite set, there exists some Safra tree in Inf( ) with v marked ‘!’. Therefore also
3 Determinization of Büchi-Automata 55
The following lemma makes use of a result which is known as König’s Infinity
Lemma (for a proof and further discussion see e.g. [62]).
Proof. Let α ∈ L(M). Then there exists an accepting run of the deterministic
Muller automaton M on α, i.e. Inf( ) ∈ F . Thus there exists some node v such
that
• v appears in all Safra trees of Inf( ), and
• v is marked at least once in Inf( ).
This has as consequences that
• v – from a certain point on – is a node of all Safra trees in , and
• in Safra trees Ti occur infinitely often with node v marked ‘!’, i.e. has
the form
qI . . . T1 . . . T2 . . . T3 . . .
The same situation is achieved if we consider a word α ∈ L(R) : Then there
exists an accepting run of the deterministic Rabin automaton R on α, i.e. there
exist a node v and an accepting pair (Ev , Fv ) such that Inf( ) ∩ Ev = ∅ and
Inf( ) ∩ Fv = ∅. By construction Ev consists of all Safra trees without node v
(Rabin 1), i.e. v – from a certain point on – is a node of all Safra trees in . As
Fv consists of all Safra trees with node v marked ‘!’ (Rabin 2), infinitely many
Safra trees Ti with node v marked ‘!’ occur in .
Thus we can proceed with the proof independently of the automaton under
consideration, taking a run on a word α, which is accepted either by the Muller
automaton M or by the Rabin automaton R.
In order to mark the node v with ‘!’ in Step 6 of Safra’s construction, it is
necessary that – at least during the computation of the transition function δ – v
has to be a parent. To become a parent, Step 2 is the only possibility in Safra’s
construction. Thus in run the node v carries before any occurrence of a Safra
tree Ti a macrostate containing a recurring state f ∈ F of the nondeterministic
Büchi automaton B.
We consider a subrun of after the point, where v occurs in all Safra trees,
in more detail: Let T and U be Safra trees of with node v marked ‘!’, such
that in no Safra tree between T and U node v is marked ‘!’. Let B be a Safra
tree between T and U such that v carries a macrostate with Q ∩ F = ∅, i.e.
T . . . B . . . U,
say (i) = T, (j) = B and (k) = U, for some 0 ≤ i ≤ j < k. Note that T
and B might be identical. Let P, H, R be the macrostate of v in T, B, U, resp.
For the sake of simplicity assume for the moment that B is the only Safra
tree between T and U, where v carries a macrostate including a recurring state.
Later we will also deal with the general situation.
As is a run on α, there exist subwords
α[i, j) := α(i)α(i + 1) . . . α(j − 1) and α[j, k) := α(j)α(j + 1) . . . α(k − 1)
of α corresponding to the finite subruns T . . . B and B . . . U of .
Consider the computation of the successor state of B and the computation
of state U from some predecessor state, say X (which might be identical with
B), at certain points in Safra’s construction of the transition function δ :
3 Determinization of Büchi-Automata 57
Point 1: During the computation of δ(B, α(j −1)) we obtain in Step 2 of Safra’s
construction a node w with macrostate H ∩ F as son of v. This node w
remains in all Safra trees before U, as no vertical merge takes place before
the computation of U.
Point 2: During the computation of U = δ(X, α(k − 1)), the condition of Step
6 of Safra’s construction becomes true, i.e., before Step 6 the nodes v and
it’s son w carry the same macrostate R.
The following picture shows the macrostates of v and w at these points, adds the
macrostate of v in T, and shows also the subwords corresponding to the subruns:
node in T at Point 1 at Point 2
α[i,j) α[j,k)
v P =⇒ H =⇒ R
⊆
=
α[j,k)
w H ∩F =⇒ R
As new macrostates on a node are computed in Step 3 by the classical powerset
construction, the lower row can be read: for all r ∈ R, there exists a h ∈ H∩F and
a finite run h . . . r of the nondeterministic Büchi automaton B on the subword
α[j, k). This run can be completed by the upper row: for all h ∈ H ∩ F, there
exists a p ∈ P and a finite run p . . . h of the nondeterministic Büchi automaton
B on the subword α[i, j). I.e., for all r ∈ R, there exists a p ∈ P and a run of B
on α[i, k) which leads from p to r while visiting a recurring state. Note that
• there might exist several such run segments, and
• that for any r ∈ R, there exists some predecessor p ∈ P – but not vice versa.
In general, there might occur several Safra trees between T and U, in which v
carries a macrostate including a recurring state. This changes our picture in the
way that we have to deal with several ‘Point 1’-situations, which might lead to
several sons of v. At Point 2 we take the union of all son macrostates. Looking
now for the run of B on α[i, k) ending in some r ∈ R, we take the first suitable
‘Point 1’-situation to switch from the lower to the upper row. This situation
arises, when the predecessor of some state r ∈ Q is a recurring state. As all
states in the macrostates of the sons of v stem from recurring states, such a
situation will always arise.
It remains to combine these finite run segments to a complete infinite run of
B : Let 0 < i1 < i2 < . . . be the positions of at which v is marked ‘!’. Let
S0 := {qI } and Sj be the macrostate of v at position ij . Now we construct a
directed tree with
• pairs (q, j) as nodes, where q ∈ Sj , j ≥ 0, and
• as parent of a node (r, j + 1) we pick one of the pairs (p, j), such that p ∈ Sj
and there exists a subrun from p to r as described above.
Obviously, this is a well formed tree with (qI , 0) as root. It has infinitely many
nodes and is finitely branching. Thus, by König’s Lemma, c.f. Theorem 3.8, there
exists an infinite path (qI , 0)(q1 , 1) . . . in the tree. Collecting all subruns along
58 Markus Roggenbach
(qI , 0)
(f, 1)
(g, 2)
(f, 3) (g, 3)
(f, 4)
(qI , 0)
(f, 1) (g, 1)
(f, 2) (g, 2)
(f, 3) (g, 3)
• S0 = {qI } and
• Si = {f, g}, i ≥ 1.
In Figure 3.12 one can see that the tree constructed is not uniquely determined:
as a parent for (g, j + 1) we have the choice between (f, j) and (g, j). For (g, 2)
we choose (f, 1), while (g, 3) has (g, 2) as parent.
Exercise 3.6. Apply Safra’s construction to the nondeterministic Büchi automa-
ton of Figure 3.1. Compare the result with the automaton of Figure 3.2 – which
states of the automaton obtained by Safra’s construction are equivalent?
Note that this result holds for Rabin automata, and that it is open whether
Safra’s construction can be improved for Muller automata.
4 Complementation of Büchi Automata Using
Alternation
Felix Klaedtke
4.1 Introduction
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 61-77, 2002.
Springer-Verlag Berlin Heidelberg 2002
62 Felix Klaedtke
4.2 Preliminaries
In this section, we provide the needed background material for this chapter. We
denote the domain of a (total) function f with dom(f ). If f is a partial function
then dom(f ) denotes the set of elements for which f is defined. It will be always
clear from the context whether f is a total function or a partial function.
Words
Let Σ be a nonempty alphabet. A finite word α of length m ∈ ω over Σ is a
mapping from {0, . . . , m−1} to Σ. We often represent α as α(0)α(1) . . . α(m−1).
An infinite word α over Σ is a mapping from ω to Σ. We often represent α as
α(0)α(1)α(2) . . . . With Σ ω we denote the set of infinite words over Σ. Occ(α)
denotes the set of letters occurring in α ∈ Σ ω , i.e.
Occ(α) := { a ∈ Σ | there exists an i ∈ ω such that α(i) = a } ,
and Inf(α) denotes the set of letters occurring infinitely often in α ∈ Σ ω , i.e.
Inf(α) := { a ∈ Σ | for all i ∈ ω there is a j ∈ ω with j > i and α(j) = a } .
Büchi Automata
A Büchi automaton, BA for short, B is a tuple (Q, Σ, δ, qI , F ), where Q is
a finite set of states, Σ is a nonempty finite alphabet, δ : Q × Σ → (Q) is P
the transition function, qI ∈ Q is the initial state, and F ⊆ Q is the set of final
states.
A run of B on α ∈ Σ ω is an infinite word over Q with (0) = qI and
(i + 1) ∈ δ((i), α(i)), for all i ∈ ω. is accepting if a final state occurs
infinitely often in , i.e. F ∩ Inf() = ∅. B accepts α if there is an accepting
run of B on α; α is rejected otherwise. L(B) denotes the set of words that B
accepts.
4 Complementation of Büchi Automata Using Alternation 63
Graphs
A graph G is a pair containing a set of vertices V and a set of edges E ⊆ V × V .
A path π in G is a nonempty finite word or an infinite word over V such that
(π(i), π(i + 1)) ∈ E, for all i + 1 ∈ dom(π). A path π is maximal if π is infinite,
or π is a finite path of length n and (π(n − 1), u) ∈ E, for all u ∈ V . A vertex
u ∈ V is reachable from v ∈ V if there is a path v0 v1 . . . vn with v = v0 and
u = vn . If n ≥ 1 then we say that u is nontrivially reachable from v. R(v)
denotes the set of vertices that are reachable from v.
We will often use graphs (V, E) together with a set C that contains some
information about the vertices in V . We will call the triple (V, E, C) also a
graph.
for i ≥ 0,
(ii) the set of edges E := { ((p, i), (q, i + 1)) ∈ V × V | q ∈ δ(p, α(i)) }, and
(iii) the set of marked vertices C := { (q, i) ∈ V | q ∈ F }.
Note that each infinite path in the run graph G corresponds to a run of B on α.
From this it follows that B accepts α iff there is an infinite path in G on which
occur infinitely many vertices in C.
64 Felix Klaedtke
The finite boundary B(G) is the set of vertices that have only finitely many
reachable vertices, i.e.
Proof. If R(v) \ B(G) is infinite then, by König’s Lemma, there exists an infinite
path in G \ B(G) starting with v, since R(v) \ B(G) is infinite and G \ B(G) is
finitely branching.
It remains to show that R(v)\B(G) is infinite. So, for a contradiction assume
that R(v) \ B(G) is finite. Let
The set B is finite since R(v) \ B(G) is finite and G is finitely branching. Since
B ⊆ B(G), we have that R(u) is finite, for all u ∈ B. We have the following
equality:
R(v) = R(v) \ B(G) ∪ R(u) .
u∈B
In particular, R(v) is a finite union of finite sets. This is not possible since R(v)
is infinite, for all v ∈ V \ B(G).
for i ≥ 0.
Lemma 4.4. Let G = (V, E, C) be a sliced graph that is finitely marked with
||G2i+1 || > 0, for some i ≥ 0. Then ||G2i+2 || < ||G2i+1 ||.
Proof. Since ||G2i+1 || > 0 the set of vertices of G2i+1 is not empty. From
Lemma 4.2 it follows that there is a vertex v0 ∈ U (G2i+1 ). From the definition of
G2i+1 = G2i \ V2i it follows that v0 ∈ V \ B(G) if i = 0, and v0 ∈ V \ B(G2i−1 )
if i > 0, where V is the set of vertices of G2i . From Lemma 4.3 we can conclude
that there exists an infinite path v0 v1 v2 . . . in G2i+1 . Obviously, vj ∈ U (G2i+1 ),
for all j ≥ 0. Let vj = (qj , kj ). It holds ||G2i+2 || < ||G2i+1 || since each slice of
G2i+2 with index kj does not contain qj .
Corollary 4.5. Let G = (V, E, C) be a sliced graph that is finitely marked and
let n = ||G||. Then G2n+1 is the empty graph.
Proof. Note that n ≤ |Q| assuming V ⊆ Q × ω for some finite set Q. Assume
that G2n+1 is not the empty graph. It holds ||G2n+1 || > 0, since G2n+1 =
G2n \ B(G2n−1 ). From the lemma above it follows that n > ||G1 || > ||G3 || >
· · · > ||G2n+1 ||. This contradicts ||G2n+1 || > 0.
Proof. (⇒) Note that the run graph G is finitely marked by Lemma 4.1. Let
µ : V → {1, . . . , 2|Q| + 1} be the function defined by µ(v) := i + 1, where i is
the uniquely determined index with v ∈ Vi and v ∈ Vi+1 . From Corollary 4.5 it
follows that 1 ≤ i ≤ 2|Q| and thus µ is well-defined. It remains to show that µ
is a progress measure.
First, we show that there is no infinite path v0 v1 . . . with µ(v0 ) = µ(v1 ) = . . .
where µ(v0 ) is odd. Assume that µ(v0 ) = 2i + 1 for v0 ∈ V . Then v0 ∈ V2i . By
definition of V2i , the vertices in V2i have only finitely many reachable states in
G if i = 0 and G2i−1 if i > 0. Thus, every path v0 v1 . . . with 2i + 1 = µ(v0 ) =
µ(v1 ) = . . . must be finite.
Second, for (u, v) ∈ E, it holds µ(u) ≥ µ(v). This follows from the fact that
(i) u ∈ U (G ) implies v ∈ U (G ), and (ii) u ∈ B(G ) implies v ∈ B(G ), for every
sliced graph G = (V , E , C ) with (u, v) ∈ V .
Third, we show by contraposition that if µ(u) = µ(v) then µ(u) is odd or
v ∈ C, for (u, v) ∈ E. Assume that µ(u) is even and v ∈ C. Since µ(u) is
even, we have that u ∈ U (G2i+1 ), for some 0 ≤ i ≤ |Q|. Since v ∈ C, it holds
u ∈ U (G2i+1 ). Hence µ(u) = µ(v).
(⇐) Let µ : V → {1, . . . , 2|Q| + 1} be a progress measure for G. Let π be an
infinite path in G. Since µ is monotonicly decreasing, there exists a k ≥ 0 with
µ(π(k)) = µ(π(k + 1)) = . . . . By the definition of a progress measure, µ(π(k))
must be even and µ(π(k + i)) ∈ C. Thus, the corresponding run of π is not
accepting. Since π was chosen arbitrarily there is no accepting run of B on α by
Lemma 4.1.
4.3.2 Complementation
The next lemma shows that BAs can check whether there exists a progress
measure or not.
Lemma 4.7. Let B = (Q, Σ, δ, qI, F ) be a BA. For every m ∈ ω, we can con-
struct a BA B with 2O(|Q|+m log |Q|) states such that B accepts α ∈ Σ ω iff there
exists a progress measure of size m for the run graph G of B for α.
Proof. Let Ψ be the set of partial functions from Q to {1, . . . , 2m + 1}. Note
that the cardinality of Ψ is |Q|O(m) = 2O(m log |Q|) . Moreover, let fI ∈ Ψ be the
partial function, where fI (qI ) := 2m + 1 and fI (q) is undefined for q = qI . Let B
P
be the BA (Ψ × (Q), Σ, δ , (fI , ∅), Ψ × {∅}) with (f , P ) ∈ δ ((f, P ), a) iff the
following conditions are satisfied:
for Pk = ∅, and
Pk+1 := { q ∈ Q | fk (p) = fk+1 (q) is odd and ((p, k), (q, k + 1)) ∈ E }
otherwise.
By induction over k it is straightforward to show that is a run of B on α.
It remains to show that is accepting, i.e., there are infinitely many k ∈ ω such
that Pk = ∅. For the sake of absurdity, assume that there is an n ∈ ω such that
Pn = ∅ and Pn+1 , Pn+2 , . . . = ∅. Note that if q ∈ Pk with k > n then there exists
a p ∈ Pn+1 such that the vertex (q, k) is reachable from a vertex (p, n + 1) in G.
Thus, there is an infinite path v0 v1 . . . with vi = (qi , ki ) for i ≥ 0, and there is
an infinite sequence of indices i0 < i1 < . . . such that qij ∈ Pkij for all j ≥ 0.
Since µ is a progress measure, it is µ(vij ) ≤ µ(vij ) for j ≥ j. Thus, there exists
a k > n such that µ(vk ) is odd and µ(vk ) = µ(vk+1 ) = . . . . This contradicts the
assumption that µ is a progress measure.
exists a progress measure of size |Q| for the run graph G of B for α, i.e., by the
construction of B that B accepts α.
Theorem 4.8. For any BA B = (Q, Σ, δ, qI, F ) we can construct a BA B with
L(B) = Σ ω \ L(B). Moreover, B has 2O(|Q| log |Q|) states.
In the remainder of the text, we reformulate the above described construc-
tion by using weak alternating parity automata. Much of the complexity of the
construction is handled by the existential and universal choice of a transition
of an alternating automaton. Complementing an alternating automaton can be
done by dualizing its transition function and its acceptance condition. We will
use games to show the correctness of this construction. This is an appetizer
for Chapter 9 where games are used to show that dualization of alternating
automata over Kripke structures corresponds to complementation.
(q0 , 0) (q0 , 2)
(q1 , 1) (q1 , 2) (q1 , 3)
(q2 , 1) (q2 , 2) (q2 , 3) (q2 , 4) ...
In such a graph the vertices are of the from (q, i), where q is a state and i ∈ ω.
The intuitive meaning of a vertex (q, i) is that A is in state q before reading the
ith input letter. Formally, a run dag G of A on α ∈ Σ ω is a graph (V, E), where
V ⊆ Q × ω with (i) (qI , 0) ∈ V , (ii) every vertex v is reachable from (qI , 0), (iii)
((p, i), (q, j)) ∈ E implies j = i + 1, and (iv) for all (p, i) ∈ V it holds that
Example 4.9. Let A be the WAPA ({q0 , q1 , q2 }, {a}, δ, q0, c) with c(q0 ) = c(q1 ) :=
2, c(q2 ) := 1, and
A run dag of A on the input word α = aaa . . . is depicted in Figure 4.1. Every
run dag on aaa . . . is non-accepting, since the slice S1 always contains q2 and
c(q2 ) = min{c(q0 ), c(q1 ), c(q2 )} is odd.
Before we turn to the definition of the games, we define the dual of a WAPA. The
dual of a formula θ ∈ B + (X) is the formula θ ∈ B + (X), obtained by exchanging
0 and 1, and ∨ and ∧ in θ. Later, we will need the following lemma.
P
(i) The set of vertices V can be partitioned into the two sets V0 and V1 with
V0 = Q × ω and V1 = Q × (Q) × ω.
(ii) The edge relation E ⊆ (V0 × V1 ) ∪ (V1 × V0 ) is defined by
and
for p, q ∈ Q, M ⊆ Q and i ∈ ω.
(iii) C ⊆ ω is a finite set of parities with c(qI ) ∈ C.
Example 4.11. The game G∗A,α , where A and α are taken from Example 4.9, is
depicted in Figure 4.2. Game positions that cannot be reached from the game
position (q0 , 0) are omitted. A solid outgoing line from a game position represents
a possible move of player 0 from that game position. The dashed lines are the
edges that correspond to the moves that player 1 can make.
1
Since we do not need to distinguish between arenas and games as, e.g., in Chapter 2
and Chapter 6, we will only use the notion of a game.
4 Complementation of Büchi Automata Using Alternation 71
A play π is a maximal path in GA,α with π(0) = (qI , 0). In the following, let
σ denote the opponent of player σ ∈ {0, 1}, i.e. σ := 1 − σ. Player σ wins the
play π if
Example 4.12. We define strategies for player 0 and player 1 for the game in
Example 4.11. The function f1 : V1 → V0 is defined by
(q1 , i) if S = {q0 , q1 } or S = {q1 , q2 },
f1 (q, S, i) := (q2 , i) if S = {q2 },
(q , i) otherwise (for some fixed q ∈ S),
The next lemma shows the tight relation between the acceptance condition
of WAPAs and winning strategies for player 0.
Lemma 4.13. Let A = (Q, Σ, δ, qI , c) be a WAPA and let α ∈ Σ ω . Player 0
has a winning strategy for the game G∗A,α iff A accepts α.
Proof. (⇐) Assume that there is an accepting run dag G of A on α with the
slices S0 , S1 , . . . . We define a strategy f0 of player 0 as follows: Given the game
position (q, i) with q ∈ Si , player 0 picks M ⊆ Si+1 with M ∈ Mod(δ(q, α(i)),
i.e. f0 (q, i) := (q, M, i + 1). For all other game position (q , i ) ∈ V0 with q ∈ Si ,
f0 (q , i ) is arbitrary. In this way, starting from the game position (qI , 0), player 0
ensures that the play proceeds along a path through the run dag. Since the run
dag is accepting, player 0 has a winning strategy.
(⇒) A winning strategy f0 defines an accepting run dag: For i ≥ 0, the
slices Si are built-up inductively, beginning with the singleton S0 := {qI }. For
i > 0 and a game position (q, i − 1) that is picked by player 1, player 0’s strategy
prescribes a set of states as the next move. The union of these states defines the
next slice Si , i.e. Si := q∈Si−1 { S ⊆ Q | f0 (q, i − 1) = (q, S, i) }. The edges from
the vertices in Si × {i} to the vertices in Si+1 × {i + 1} are inserted according
to the transition function. The run dag is accepting.
The next lemma shows the determinacy of the games. For its proof we need
the definition of an attractor set of player σ of a set of game positions for a game
GA,α = (V, E, C).
The attractor of player σ of X ⊆ V , Attrσ (X) for short, contains all game
positions from which player σ can force player σ a visit (after finitely many
moves) to a game position in X or to a game position where player σ cannot
make a move.
Attr0σ (X) := X ,
σ (X) := Attrσ (X) ∪
Attri+1 i
and
Attrσ (X) := Attriσ (X) .
i∈ω
V0 ∪ V1
Attr20 (X)
Attr0 (X) Attr10 (X)
Attr00 (X) = X
(q, S, i) •
... •
• •
(q , i ) •
..
.
•
Fig. 4.4. Construction of the attractor of player 0.
possible to stay outside of Attrσ (X), and if it is the turn of player σ, then it is
not possible for him to choose a game position in Attrσ (X).
Lemma 4.14. Let A = (Q, Σ, δ, qI, c) be a WAPA and let α ∈ Σ ω . Either
player 0 or player 1 has a winning strategy for the game GA,α = (V, E, C).
Proof. Let k := min C, and let T denote the set of game positions where player 1
can force player 0 to visit a game position from which player 0 cannot make a
move, i.e. T := Attr1 ({ u ∈ V0 | for all v ∈ V is (u, v) ∈
/ E }).
We show by induction over |C|, i.e., the number of parities in GA,α , that
either player 0 or player 1 has a winning strategy. For |C| = 1, player 0 has a
winning strategy if k is even and (qI , 0) ∈
/ T , and player 1 has a winning strategy
if k is odd or (qI , 0) ∈ T .
Let |C| > 1. If (qI , 0) ∈ T then obviously player 1 has a winning strategy.
Let (qI , 0) ∈ T and assume that k is even (the proof for the case that k is odd is
analogous). Moreover, let U = { v ∈ V | c(v) = k }.
If (qI , 0) ∈ Attr0 (U ), then player 0 has a winning strategy. If (qI , 0) ∈
Attr0 (U ) then we define the game GA,α by removing the parity k and by deleting
the edges (u, v) ∈ E ∩ V1 × V0 where u is a game position in U . By the induction
hypothesis we know that either player 0 or player 1 has a winning strategy for
GA,α . We will show that if player σ has a winning strategy in GA,α then the
player σ has also a winning strategy in GA,α .
Assume that player 1 has a winning strategy for GA,α . Playing according
to this strategy ensures the players to stay outside of Attr0 (U ) in any play in
GA,α , as otherwise there would be a game position belonging to Attr0 (U ) in
GA,α . Therefore, player 1 also has a winning strategy for GA,α .
Suppose that player 0 has a winning strategy for GA,α . If player 0 plays
according to this strategy in GA,α , then the only possibility for player 1 to give
another progression to the play as in GA,α is to move into Attr0 (U ) if possible.
But if player 1 picks a game position in Attr0 (U ), player 0 can force to visit a
game position in U . Then player 0 wins the play, due to the choice of U (note
that k is even). Therefore, player 0 has a winning strategy for GA,α .
74 Felix Klaedtke
The next lemma connects winning strategies for the games G∗A,α and G∗A,α .
Proof. (⇒) Let f0 be a winning strategy for player 0 for G∗A,α . Further, let
(q, S, k) be a game position of G∗A,α . If there exists a play in G∗A,α such that
(q, S, k) appears in this play, then S ∈ Mod(δ(q, α(k − 1))). We do not need
to define a strategy for vertices that do not appear in any play in G∗A,α . From
Lemma 4.10 it follows that there exists a p ∈ S ∩ W , for f0 (q, k − 1) = (q, W, k),
since W ∈ Mod(δ(q, α(k − 1))). We define a strategy of player 1 as f 1 (q, S, k) :=
(p, k).
We show that f 1 is a winning strategy for G∗A,α . For a play π in G∗A,α played
according to f 1 , there exists a play π in G∗A,α played according to f0 such that
c(π(i)) − 1 = c(π(i)), for all i ≥ 0. Since player 0 wins π in G∗A,α , player 1 wins
π in G∗A,α .
(⇐) Let f 1 be a winning strategy of player 1 for G∗A,α and let (q, k) be a
game position of G∗A,α . The set T := { p ∈ Q | (p, k + 1) = f 1 (q, R, k) with R ∈
Mod(δ(q, α(k))) } is a model of δ(q, α(k)) by Lemma 4.10. Let S ⊆ T be a
minimal model of δ(q, α(k)). As above, it can be shown that the strategy f0
with f0 (q, k) := (q, S, k + 1) is a winning strategy for G∗A,α for player 0.
Now, we have all ingredients for proving that the dual automaton accepts
the complement.
Exercise 4.3. The natural number m is an upper bound for the progress
measures for the BA B = (Q, Σ, δ, qI , F ) if for all α ∈ Σ ω , B rejects α iff there
exists a progress measure of size m for the run graph. Show that it is Pspace-
complete to decide if m ∈ ω is the smallest upper bound for the progress measure
for a BA. (Hint: Use the fact that the emptiness problem for WAPAs is Pspace-
complete.)
Theorem 4.18. For any WAPA A we can construct a Büchi automaton B with
2
L(A) = L(B). Moreover, if A has n states then B has 2O(n ) states; and if A is
stratified then B has 2 O(n)
states.
and
δ ((S, ∅), a) := { (S , S \ E) | S ∈ Mod δ(q, a) } ,
q∈S
Exercise 4.4. Show that there is no linear translation from BAs to stratified WA-
PAs. (Hint: Use the family of languages (Ln )n≥2 from Theorem 1.30 in Chap-
ter 1.)
4.5.2 Complementation
The construction can be improved by tuning the translation from the WAPA
A to the BA B = (P, Σ, δ, qI , F ). The improvement is due to Kupferman and
P
Vardi [102]. Let Q be the set of states of B. Note that P = (Q × {0, . . . , 2n}) ×
P (Q × {0, . . . , 2n}) by the construction of B. A set S ⊆ Q × {0, . . . , 2n} is
consistent if (q, i), (q, j) ∈ S implies i = j. Let B be the BA (P , Σ, δ, qI , F ),
where P be the set of pairs of consistent sets (S, O) with O ⊆ S, and the
transition function δ is restricted to P , i.e. δ ((S, O), a) := δ((S, O), a), for
(S, O) ∈ P .
Note that (S, O) ∈ P can be represented by the set { q ∈ Q | (q, i) ∈ O }
and the partial function f : Q → {0, . . . , 2n}, where f (q) := i, for (q, i) ∈ S,
and otherwise f is undefined. It is easy to see that B and the BA obtained by
the construction described in Section 4.3 are isomorphic. This was observed by
Kupferman and Vardi in [107]. Thus, L(B ) = Σ ω \ L(B), and B has 2O(n log n)
states.
5 Determinization and Complementation of
Streett Automata
Stefan Schwoon
5.1 Introduction
Several classes of ω-automata have been proposed in the literature, most im-
portantly Büchi automata, Muller automata, and Rabin automata. It has been
shown that the expressiveness of all these models is equivalent.
Streett automata were first suggested in [171]. They differ from the other for-
malisms in their acceptance condition which models strong fairness constraints.
Again, their expressiveness is equal to that of Büchi automata; however, for
certain properties of ω-words Streett automata can have an exponentially more
succinct representation.
Here, we survey results about upper and lower bounds on the problems of
determinization and complementation of Streett automata. A relatively simple
argument yields a doubly exponential upper bound for determinization; how-
ever, Safra [159] found an exponential procedure, involving a transformation
from (non-deterministic) Streett automata to deterministic Rabin automata (see
Section 5.3), and another transformation from deterministic Streett automata
to deterministic Rabin automata (Section 5.4). We present a slight modification
of Safra’s method and try to provide ample intuition for the method and its
proof of correctness. Moreover, the results of [114] lead to lower bounds on these
problems; these, together with their proofs, are portrayed in Section 5.5.
5.2 Definitions
A (non-deterministic) Streett automaton A is a 5-tuple (Q, Σ, δ, q0 , Ω) where Q
is a finite set of states, Σ is a finite alphabet, δ : Q × Σ → 2Q is a transi-
tion function, q0 is an initial state, and Ω = {(L1 , U1 ), . . . , (Lh , Uh )} is an
acceptance condition where Li and Ui , 1 ≤ i ≤ h, are subsets of Q. We
call the tuples (Li , Ui ) the acceptance pairs of A. For deterministic automata,
|δ(q, a)| = 1 holds for all q ∈ Q, a ∈ Σ.
For an infinite word α ∈ Σ ω of the form α = α0 α1 . . . , we say that ξ ∈ Qω is
a run of A over α if ξ = ξ0 ξ1 . . . where ξ0 = q0 and ξi+1 ∈ δ(ξi , αi ) for all i ≥ 0.
The infinity set Inf(ξ) is the set of states visited infinitely often by ξ.
A word α ∈ Σ ω is accepted by the Streett automaton A if there is an infinite
run over α such that the following condition holds for all i ∈ {1, . . . , h}: If some
state in Li is visited infinitely often, then some state in Ui is visited infinitely
often, too. The set of infinite words accepted by A is denoted L(A).
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 79-91, 2002.
Springer-Verlag Berlin Heidelberg 2002
80 Stefan Schwoon
The acceptance condition can be seen as a model for strong fairness. If, for
each i, the states in Li represent a request for some action and Ui corresponds
to the execution of said action, then the condition postulates that every request
which is repeated an unbounded number of times is eventually granted.
A Rabin automaton has the same structure as a Streett automaton, but
with a different interpretation. The Rabin acceptance condition declares a run as
accepting if there exists an index i, 1 ≤ i ≤ h, such that Li is visited infinitely
often, but Ui is visited only finitely often.
Since their structures are the same, each Streett automaton can be inter-
preted as a Rabin automaton, and vice versa. Observe that the Rabin acceptance
condition is exactly the opposite of the Streett condition. Hence, in the deter-
ministic case where there is exactly one possible run for every input, the same
automaton under the Streett interpretation represents the complement of its lan-
guage under the Rabin interpretation. We will use this fact in our constructions
later on.
Definition 5.1. A set J ⊆ {1, . . . , h} is a witness set for α ∈ L(A) if there
is an accepting run over α such that for all indices j ∈ J the set Uj is visited
infinitely often, and for all indices j ∈
/ J the set Lj is visited only finitely often.
It is easy to see that α ∈ L(A) exactly if α has at least one witness set. Hence,
the problem of checking acceptance of a Streett automaton can be reduced to
finding a witness set for the input. The constructions in later sections are based
on this idea.
Note also that an accepting run can yield multiple witness sets if there are
indices j where Lj is visited finitely often, and Uj is visited infinitely often.
node in such a tree represents a process that is “hoping” for a certain witness
set to be realised. The witness set of a parent node is a superset of that of its
child nodes.
A process for a witness set J observes a subset of the possible runs and waits
for all the Uj , j ∈ J, to be visited in order. If that happens without visiting
any Lj , j ∈ / J, the process “succeeds” and starts over. If some Lj , j ∈
/ J,
is encountered in a run, the process discards that run. If some process succeeds
infinitely often, D accepts.
Fix a Streett automaton A = (Q, Σ, δ, q0 , Ω) for the rest of the section,
where Ω = {(L1 , U1 ), . . . , (Lh , Uh )}. Let H = {1, . . . , h}. A (Q, H)-tree is a
finitely branching rooted tree whose leaves are labelled with non-empty subsets
of Q and whose edges are labelled by elements of H ∪ {0}. The labels of the
leaves are pairwise disjoint and, other than 0, no edge label occurs twice on
any path from the root to one of the leaves. Each node bears a name from the
set {1, . . . , 2 · |Q| · (h + 1)}. The child nodes of a parent are ordered (from left
to right, say). For each parent node, at least one child is connected by an edge
with a non-zero label.
The root node represents the “hope” that H is a witness set for the input.
If some node n represents witness set J, and an edge labelled j goes from n to
a child node, then the child node represents witness set J \ {j}. Informally, the
child node has given up on hoping for Uj to occur.
With this intuition in mind, we now present the formal construction of D.
We will explain this construction a bit more before we prove its correctness and
the size bound.
q1
a b
b a
q2 q3 L1 = {q2 }
b b U2 = {q3 , q5 }
b b
b
q4 q5
b
Fig. 5.1. Example Streett automaton.
completed cycle). If two runs are advanced to the same level, we keep the older
one, i.e. the one in the leftmost child (otherwise a state might change its position
in the tree too often).
In step 5 we clean up “dead” branches of the tree. In step 6, when all the
runs observed by a process have succeeded, we reset the process. In this way we
limit the growth of the tree.
5.3.2 An Example
Consider the non-deterministic Streett automaton A shown in Figure 5.1 which
has two accepting pairs. For the purposes of this example we will consider the
sets L1 = {q2 } and U2 = {q3 , q5 } only. Assume that the automaton has to
process input starting with abb.
Figure 5.2 shows the behaviour of the corresponding deterministic automa-
ton. Part (a) shows the steps made by the construction upon receiving the in-
put a. For a better demonstration we assume that the six steps are executed
on all leafs simultaneously; the correctness is not affected by this. Diagram (a1)
shows the initial state of the deterministic automaton, consisting of the root
labelled {q1 } and named 1.
Diagram (a2) shows the tree after expansion and simulation of A (i.e. steps 1
and 2 of the transition function). Diagram (a3) shows the result of step 3; q2
is in L1 and is thus “reset”; q3 is in U2 and hence advanced to the next index
by the root node. Now q3 appears twice below the root node; this is repaired in
step 4, shown in (a4). In the last step, the now empty sub-tree with the node
named 3 is removed; the result is in diagram (a5).
Part (b) of the figure shows the processing of the second input character.
Diagram (b1) shows the situation after steps 1 and 2. In (b2) the final result is
shown: q5 was advanced to a 0-labelled branch, and the duplicate q1 was removed
from the less advanced sub-tree of the root.
Part (c) demonstrates the third step; again we read a b. Diagram (c1) shows
the result of steps 1 and 2. Afterwards, q5 is removed from the left branch (since
it is less advanced), and the occurrence of q5 in the right branch is advanced to
a 0-labelled branch. Since that leaves the label 0 on all edges leaving node 4, the
84 Stefan Schwoon
(a)
(a1) (a2) (a3) (a4) (a5)
1 1 1 1 1
{q1 } 2 2 1 2 1 2 1
2 2 4 2 4 2 4
1
{q3 } {q3 } {q3 }
1 1 1 1 1
3 3 5 3 5 5
{q2 , q3 } {q3 } {q2 } ∅ {q2 } {q2 }
(b) (c)
(b1) (b2) (c1) (c2)
1 1 1 1
2 1 2 1 2 1 2 1
2 4 2 4 2 4 2 4
1 2 1 2 0 1 2 0 1 {q4 , q5 }
5 6 5 6 7 5 6 7 5
{q1 , q4 } {q1 , q5 } {q4 } {q1 } {q5 } {q1 , q5 } {q5 } 2
{q1 }
8
{q4 }
Fig. 5.2. (a) After input a. (b) After input ab. (c) After input abb.
subtrees of that node are collapsed according to step 6 of the transition. The
result is shown in Diagram (c2).
i ≥ 0, ξAi i
occurs in ξD . We first wait until ξA makes no more visits to any set Lj ,
j ∈/ J (which happens after consumption of a finite number of input symbols).
From now on, either ξD i
becomes a leaf infinitely many times, then α ∈ L(D).
i i
Otherwise we wait until ξD becomes a leaf for the last time. Thereafter, ξA
i i
always occurs in some sub-tree of ξD . We will see that ξA can ‘migrate’ from
one sub-tree to another as time goes on, but the number of migrations is finitely
bounded: Since all Uj , j ∈ J, occur infinitely often, ξA i
is moved to other sub-
trees until it eventually ends up in one whose edge is labelled with an element
of (H ∪ {0}) \ J. From there, ξA i
cannot migrate further via steps 3 or 6 of the
construction anymore, only via rule 4. But this can happen only finitely often
i
since the set of edge labels is finite, and since there is a leftmost sub-tree of ξD .
i
Eventually, ξA must end up in some sub-tree which it never leaves anymore. We
can repeat the argument recursively for this sub-tree until we arrive at one that
becomes a leaf infinitely many times. This recursion is of finite depth because
eventually we get to processes for the witness set J. If we do not get a leaf at
i
this level, then ξA must be in a 0-indexed sub-tree, and some other state must
be trapped in a non-0-indexed sub-tree. Thus, if we continue the recursion, the
i
0-indexed sub-trees containing ξA have less and less states. In the worst case, we
i
continue until a sub-tree contains only the single state ξA . Since ξA is accepting
with witness set J, and since all the sets Lj , j ∈ J occur infinitely often in the
accepting run of A, this node must become a leaf infinitely many times.
Let us count how many nodes a (Q, H)-tree can have. A tree has at most |Q|
leaves. Suppose a tree has no 0-labelled edges, then the length of a path from the
root to a leaf can be at most h. Such a tree can have at most |Q| · h nodes apart
from the root. Now consider trees with 0-edges. From the root, there must be at
least one path to a leaf without passing a 0-edge, and likewise from each node
which is the target of such an edge. So there can be at most |Q| − 1 such edges
(and nodes). Therefore, in total, a (Q, H)-tree can have at most m := |Q|·(h+1)
nodes.
With this knowledge we can determine the size of D, i.e. the number of
(Q, H)-trees. Let us consider unlabelled trees without names on the nodes first.
According to Cayley’s Theorem there can be at most mm−2 different trees with
m nodes (even ignoring the fact that we are only interested in trees with at most
|Q| leaves). We then have no more than (m − 1)h+1 choices to label the edges,
(2m)!/m! choices to name the nodes, and (|Q| + 1)|Q| choices to distribute states
among the leaves. Taking these results together, we get at most
2m!
mm−2 · (h + 1)m−1 · · (|Q| + 1)|Q| = 2O(m log m) = 2O(|Q|h log(|Q|h))
m!
different states.
In a D-state of the form (q, π, l, u), the state q simply records the state that
A would be in after the same input. The values for l and u contain information
about visits to the accepting pairs. If a set Lj is visited, then its position in
the current permutation π is recorded in l. Similarly, if a set Uj is visited, then
its position is written to u and additionally the index is shifted to the right in
the permutation. If multiple sets are visited in one step, then we choose the one
that’s leftmost in π.
The intuition behind this construction is that the permutation of the indices
maintained by D will take on the form outlined in Figure 5.3. Those indices j
for which Uj is visited only finitely often will eventually gather on the “left”
side of the permutation whereas the others will be moved to the back infinitely
often. If a run satisfies the acceptance condition, then the third component of
the D-states must always indicate the right half of the permutation from some
point on.
(L(D) ⊆ L(A)) Let α ∈ L(D). In the unique run of D over α, there must be
an index j such that (a) Lj is visited infinitely often and (b) Uj is visited only
finitely often. Because of (b), if some Lk is visited infinitely often in A, then k
must occur infinitely often at position j or to the right of it in the permutation.
Because of (a) we know that no index with this property is allowed to keep its
position forever. Therefore it follows that k is also moved back to the end of the
permutation infinitely often, and every such move means a visit to Uk .
(L(A) ⊆ L(D)) Suppose α ∈ L(A), and consider the maximal witness set J
associated with its run. Let i = h − |J| + 1. At some point, for all k ∈/ J all the
finitely many visits to Lk and Uk have been made, so k is never moved back in
the permutation again. After that, we wait until all indices in J are eventually
moved to the back. From now on, all indices on positions left of i are outside of
J, so the D-run will never visit Ui again. Whatever the index k on position i is,
it is from now on always an element of J. When A next visits Uk , D visits Li .
5.4.2 Summary
Let A be a (non-deterministic) Streett automaton with n states and h accept-
ing pairs. According to Section 5.3 we can transform A to a deterministic au-
tomaton D accepting the complement of L(A) with 2O(nh log(nh)) ) states and
2n(h + 1) accepting pairs. The result of this section lets us transform D into an-
other deterministic Streett automaton accepting L(A) with the same asymptotic
bound on the number of states and 2n(h + 1) + 1 accepting pairs.
i1 ij−1 ij ih
1 j−1 j h
q1 q2 qn
0, ..., n 0, ..., n 0, ..., n
...
2 2
1
1 n
n
q0
0 0 0
1 1 1
0
q1 q2 qn
1 1 1
q0
0 1 0 1 0 1
0 0 0
1 0 0 0
0
0 0
Analogously, we switch from απ back to απ when we have passed the sequence
j1 . . . jn and visited the states in Inf(ξπ ).
We now derive a contradiction: Clearly, Inf(ξ) = Inf(ξπ ) ∪ Inf(ξπ ), i.e. ξ
satisfies An ’s acceptance condition: for every acceptance pair (Lk , Uk ) of An
such that Inf(ξ) has a state in Lk , said state must come from either Inf(ξπ ) or
Inf(ξπ ); and since both απ and απ are in L(An ), Inf(ξ) also has a state from Uk .
So α ∈ L(An ).
On the other hand, one can show that α ∈ L(Bn ) which is a contradiction.
Take the smallest k such that ik = jk . Since ik must occur in π later on, there
is l > k such that jl = ik . Similarly, there is m > k such that im = jk . Since α
infinitely often cycles through π and π , the sequences ik ik+1 . . . im = jk and
jk jk+1 . . . jl = ik happen infinitely often, hence α matches the characterization
given above.
Note that the Streett automaton derived from Bn has only one acceptance
pair (albeit one of size O(n)). The alphabet is also of size O(n). However, the
idea can be transferred to a family of automata with a binary alphabet. Such
an automaton (still of size O(n)) is shown in Figure 5.5, (which corrects a slight
mistake in [114]).
In Figure 5.5 the input characters are replaced with the sequence ai b for
0 ≤ i < n and an a∗ b for n.
• Qn = {−n, . . . , n} \ {0};
• Σn = {1, . . . , n};
• q0n = −1;
90 Stefan Schwoon
1, ..., n
q0 q1
1, ..., n
q10 q20 0
qn
q11 q21 1
qn
q1f q2f f
qn
Ralf Küsters
6.1 Introduction
In Chapter 2, parity games were introduced and it was claimed that these games
are determined and that both players have memoryless winning strategies. The
main purpose of this chapter is to provide proofs for this claim.
The determinacy and the memorylessness of parity games is exploited in
various areas inside and outside of game theory. For the purpose of this book,
automata theory, modal logics, and monadic second-order logics are the most
interesting among them.
More specifically, the word and emptiness problem for alternating tree au-
tomata as well as model checking and satisfiability in modal µ-calculus [100]
can be reduced to deciding the winner of a parity game. In fact, model checking
µ-calculus is equivalent via linear time reduction to this problem [56, 55, 202].
In Chapter 9 and 10, these reductions are presented.
In addition, parity games provide an elegant means to simplify the most diffi-
cult part of Rabin’s proof of the decidability of the monadic second-order theory
of the binary infinite tree [148], the complementation lemma for automata on
infinite trees. Although, from Rabin’s proof the determinacy of parity games fol-
lows implicitly, Rabin did not explicitly use games to show his result. The idea
to use games is due to Büchi [21] and it was applied successfully by Gurevich
and Harrington [77]. In turn, their paper has been followed by numerous other
attempts to clarify and simplify the proof of the complementation lemma; see,
for instance, a paper by Emerson and Jutla [55]. For the proof of the complemen-
tation lemma see Chapter 8. We refer to Part VI and VII for more on monadic
second-order logics.
The determinacy of parity games follows from a result due to Martin [119],
who has shown that Borel games, a class of games much larger than the class
of parity games we consider here, are determined. For our purpose, however,
this result does not suffice since the strategies employed there require to store
the complete history of a play, and thus, they require infinite memory. Gurevich
and Harrington [77] showed that finite-memory strategies suffice to win Muller
games, a class more general than parity games, but smaller than Borel games
(see Chapter 2).1 Later, it turned out that for parity games the winner only
needs a memoryless strategy. This was proved for the first time independently
by Emerson and Jutla [55] and Mostowski [132]. While these proofs were quite
1
Apparently, Büchi was the first to prove the existence of finite-memory strategies in
a manuscript sent to Gurevich and Harrington.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 95-106, 2002.
Springer-Verlag Berlin Heidelberg 2002
96 Ralf Küsters
involved and non-constructive in the sense that the proofs did not exhibit memo-
ryless winning strategies, McNaughton [126] proposed a simpler and constructive
proof for Muller games played on finite graphs, from which he could derive an
exponential-time algorithm for computing finite-memory strategies. His results
also establish the existence of memoryless winning strategies for parity games
on finite graphs.
In the present chapter, we follow a proof proposed by Zielonka [203] to show
that parity games (on possibly infinite graphs) are determined and that the
winner of a game has a memoryless winning strategy. We present both a con-
structive and a non-constructive proof. In addition, we sketch algorithmic and
complexity-theoretic issues. We show that the problem of deciding the winner
of a parity game belongs to the complexity classes NP and co-NP. Based on
the constructive proof of determinacy, a simple deterministic exponential-time
algorithm is derived to compute the winning positions of players along with their
memoryless strategies. Jurdziński [92, 93] proved tighter complexity results and
developed more efficient algorithms. An in-depth treatment of his results and
other approaches for computing winning regions is provided in Chapter 7.
The present chapter is structured as follows. In Section 6.2, some basic no-
tions are introduced. They prepare for the proof of the main theorem of this
chapter, which is shown in Section 6.3. Finally, in Section 6.4 the mentioned
complexity-theoretic and algorithmic issues are discussed.
We assume that the reader is familiar with the notions introduced in Chap-
ter 2, such as parity games, (memoryless) strategies, determinacy, etc.
Throughout this chapter let G = (A, χ) denote a parity game with arena
A = (V0 , V1 , E) and colouring function χ. The set of vertices of G will be denoted
by V := V0 ∪ V1 .
6.2.1 Subgames
Let U ⊆ V be any subset of V . The subgraph of G induced by U is denoted
Example 6.1. Figure 6.1 depicts a simple parity game, subsequently called Gex ,
with the vertices v0 , . . . , v7 and colours 0, 1, 2. As in Chapter 2, circles denote
0-vertices and boxes 1-vertices. In this game, G[{v5 , v6 }] is a subgame of G.
6 Memoryless Determinacy of Parity Games 97
v0 v1 v2
1 1 0
v7 0 1 v3
1 2 2
v6 v5 v4
Fig. 6.1. A parity game
6.2.2 σ-Traps
The notion of a σ-trap was introduced in Chapter 2. Recall that if a token is
in a σ-trap U , then Player σ can play a strategy consisting in choosing always
successors inside of U . On the other hand, since all successors of σ-vertices in U
belong to U , Player σ has no possibility to force the token outside of U . In our
example, the set {v0 , v7 } is a 1-trap, while the set {v0 , v1 , v2 , v3 , v7 } is a 0-trap.
We summarize some simple properties of σ-traps.
Lemma 6.3. (1) For every σ-trap U in G, G[U ] is asubgame.
(2) For every family {Ui }i∈I of σ-traps Ui , the union i∈I Ui is a σ-trap as well.
(3) If X is a σ-trap in G and Y is a subset of X, then Y is a σ-trap in G iff Y
is a σ-trap in G[X].
Proof. Exercise.
Note that the converse of (1) is not true. In our example, the set {v3 , v4 , v5 , v6 }
induces a subgame in G, but it is neither a 0-trap nor a 1-trap. Also observe that
the equivalence in (3) does not hold for nested traps of different types: If X is a
σ-trap in G and Y is a σ-trap in G[X], then, in general, Y is not a trap of any
kind (neither σ nor σ) in G.
98 Ralf Küsters
Attractors and attractor sets were introduced in Chapter 2. Recall that the
attractor set Attrσ (G, X) ⊆ V for Player σ and set X is the set of vertices
from which Player σ has a strategy — and according to Proposition 2.18 a
memoryless strategy — to attract the token to X or a dead end in Vσ in a finite
(possibly 0) number of steps. In our example, Attr1 (Gex , {v2 }) = {v1 , v2 } and
Attr0 (Gex , {v2 }) contains all vertices of Gex .
We summarize relevant relationships between attractors and traps in the
following lemma.
Lemma 6.4. (1) The set V \ Attrσ (G, X) is a σ-trap in G.
(2) If X is a σ-trap in G, then so is Attrσ (G, X).
(3) X is a σ-trap in G iff Attrσ (G, V \ X) = V \ X.
(4) Attrσ (G, X) = V \ U where U is the greatest (w.r.t. set inclusion) σ-trap
contained in V \ X; U exists since ∅ is a σ-trap, and by Lemma 6.3, the
union of σ-traps is a σ-trap.
6.2.4 σ-Paradise
• U is a σ-trap, and
• there exists a memoryless winning strategy fσ for σ on U , i.e.,
– fσ is a total mapping from U ∩ Vσ into U such that, for all v ∈ U ∩ Vσ ,
fσ (v) ∈ vE; and
6 Memoryless Determinacy of Parity Games 99
6.3 Determinacy
Following Zielonka [203] in this section we show that parity games are deter-
mined and that the winner of a parity game has a memoryless winning strategy.
Formally, the main theorem of this chapter reads as follows.
Theorem 6.6. The set of vertices of a parity game is partitioned into a 0-
paradise and a 1-paradise.
Note that the 0- and 1-paradises are the winning regions of the players. We
provide two proofs of this theorem. The first proof is non-constructive, whereas
the second one is constructive. For parity games on finite graphs, the latter proof
can even be turned into a recursive algorithm for computing the winning regions
of the players, along with their memoryless winning strategies (see Section 6.4)
100 Ralf Küsters
Attrσ (G[Xσ ], N )
Z
Zσ Zσ
Xσ Xσ
Fig. 6.2. Construction of Xσ and Xσ
Proof. Since the maximum priority of G is 0, Player 1 can only win G on dead
ends in V0 or vertices from which he can force the token to such a dead end.
That is, the 1-paradise is the set Attr1 (G, ∅) with attr1 (G, ∅) as a memoryless
winning strategy. Since V \ Attr1 (G, ∅) is a 1-trap and the maximum priority of
G is 0, it es easy to see that V \ Attr1 (G, ∅) is a 0-paradise.
σ ≡ n mod 2 (6.1)
be the player that wins if the token visits infinitely often the maximum priority
n. Let Xσ be a σ-paradise such that Xσ := V \ Xσ is a σ-trap. Finally, let
G: Once in Xσ , σ cannot move the token outside this set; although from Zσ , σ
can move the token inside Xσ , σ cannot move it outside Zσ in G[Xσ ]. Moreover,
when playing according to xσ in Xσ and according to zσ in Zσ two cases can
occur:
(1) At some moment in a play the token hits the set Xσ . Then, from this moment
on, σ plays according to xσ and wins the play.
(2) The token stays forever in Zσ . Since in this set, σ plays according to zσ , σ
wins as well.
This shows:
Lemma 6.8. The union Xσ ∪ Zσ is a σ-paradise.
This lemma will later allow us to extend σ-paradises. Conversely, if Xσ cannot
be extended in this way, one can show that it is not possible to extend Xσ at all
and that Xσ is a σ-paradise:
With these lemmas at hand, the non-constructive and the constructive proofs of
Theorem 6.6 are rather straightforward.
A non-constructive proof of Theorem 6.6. Let n be the maximum priority
occurring in G. If n=0, then Theorem 6.6 follows from Lemma 6.7.
Suppose that n ≥ 1 and let σ be defined as in (6.1). Let Wσ = {Wσq }q∈Q
be the family of all σ-paradises. Because of Lemma 6.5 we know that Wσ =
q
q∈Q Wσ is the greatest among these σ-paradises, say with memoryless winning
102 Ralf Küsters
Alternative proofs. We conclude this section with some remarks on yet an-
other proof of determinacy. The proof presented by Emerson and Jutla [55] is
a non-inductive proof. The idea is that given a game the set W of winning po-
sitions of a player is expressed by a µ-calculus formula ϕ. Now it is possible to
deduce that the complement of W is indeed the set of winning positions for the
opponent from the fact that the negation of ϕ has the same form as ϕ after
exchanging the roles of both players. This shows that from every vertex one of
the players has a winning strategy, and thus, the game is determined. Note that
the µ-calculus formula and its negation, describing the winning positions of a
player and its adversary, respectively, allow to calculate the winning sets of both
players independently. In the non-constructive and constructive proofs presented
above, we first constructed Wσ , and depending on this set defined Wσ .
Finally, using a ranking argument, Emerson and Jutla proved (in a non-
constructive manner) the existence of memoryless winning strategies.
w(v ) are eliminated, i.e., for a 0-vertex we only keep the outgoing edge referred
to by w.
Given Gw , we need to check whether there exists a vertex v reachable from
v in Gw such that a) χ(v ) is odd and b) v lies on a cycle in Gw containing only
vertices of priority less or equal χ(v ). If, and only if, such a vertex v does not
exist, w is a winning strategy for Player 0. Checking this can be carried out in
polynomial time. (We leave the proof as an exercise.) Thus, Wins ∈ NP.
We now show Wins ∈ co-NP. By Theorem 6.6, deciding (G, v) ∈ Wins means
deciding whether v is a winning position for Player 1. This can be achieved by
the above algorithm if we require χ(v ) to be even. (Alternatively, one can apply
the above NP-algorithm to the dual game, i.e., the one where 0-vertices and
1-vertices are switched and the priorities are increased by 1). Consequently,
Wins ∈ co-NP.
Exercise 6.1. Complete the proof of Corollary 6.10.
The result just proved also follows from the work by Emerson, Jutla, and Sistla
[56], who showed that the modal µ-calculus model checking problem is in NP ∩
co-NP. This problem is equivalent via linear time reduction to Wins. Jurdziński
[92] has proved the even stronger result that Wins ∈ UP ∩ co-UP, where UP
is the class of languages recognizable by unambiguous polynomial-time non-
deterministic Turing machines, i.e., those with at most one accepting computa-
tion of length polynomially bounded in the size of the input; as usual, co-UP
denotes the problems whose complement is in UP.
winning-regions(G)
n := max{ χ(v) | v ∈ V }
If n = 0 then return ((V \ Attr1 (G, ∅), w0 ), (Attr1 (G, ∅), attr1 (G, ∅)))
// w0 is some memoryless strategy for Player 0
// otherwise
σ := n mod 2
// compute Wσ , wσ
(Wσ , wσ ) :=win-opponent(G, σ, n)
// compute Wσ , wσ
Wσ := V \ Wσ
N := { v ∈ Wσ | n ∈ χ(v) } // see (6.2)
Z := Wσ \ Attrσ (G[Wσ ], N ) // see (6.2)
((Z0 , z0 ), (Z1 , z1 )) :=winning-regions(G[Z])
∀v ∈ Wσ ∩ Vσ : // see (6.3)
8
< zσ (v) if v ∈ Z,
if v ∈ Attrσ (G[Wσ ], N ) \ N,
: v
wσ (v) = attrσ (G[Wσ ], N )(v)
if v ∈ N and v ∈ vE ∩ Wσ
Fig. 6.3. A deterministic algorithm computing the winning regions of a parity game
win-opponent(G, σ, n)
Repeat
(W , w ) := (W, w)
X := Attrσ (G, W )
∀v ∈ X ∩ Vσ :
w(v) if v ∈ W,
x(v) =
attrσ (G, W )(v) if v ∈ X \ W.
Y := V \ X;
N := { v ∈ Y | n = χ(v) } // see (6.2)
Z := Y \ Attrσ (G[Y ], N ) // see (6.2)
((Z0 , z0 ), (Z1 , z1 )) =winning-regions(G[Z])
W := X ∪ Zσ
∀v ∈ W :
x(v) if v ∈ X,
w(v) =
zσ (v) if v ∈ Zσ .
Until W = W
return (W, w)
Corollary 6.11. Computing the winning regions of finite parity games and the
corresponding memoryless winning strategies can be carried out in time O(m·ln ).
The best known deterministic algorithm for computing winning regions is due
to Jurdzińzski [93] and is discussed in Chapter 7 (see Theorem 7.25). Unlike the
algorithm presented here, Jurdzińzski’s algorithm only needs polynomial space.
The following chapter also includes other up-to-date approaches to the problem
of deciding the winner of a parity game.
7 Algorithms for Parity Games
Hartmut Klauck
School of Mathematics
Institute for Advanced Study, Princeton
7.1 Introduction
It is the aim of this chapter to review some of the algorithmic approaches to the
problem of computing winning strategies (resp. of deciding if a player has a win-
ning strategy from a given vertex) in parity games with finite arenas and other
two-player games. Parity games are equivalent via linear time reductions to the
problem of modal µ-calculus model checking (see Chapters 10 and 9), and this
model checking problem plays a major role in computer-aided verification. Fur-
thermore we will see that the problem is not too hard in a complexity-theoretic
sense, while no efficient algorithm for it is known so far. Also parity games are
the simplest of a whole chain of two-player games for which no efficient solu-
tions are known, further underlining the importance of looking for an efficient
algorithm solving this particular problem.
We will explain why the problem of solving parity games lies in UP ∩ co-UP,
explore its relations to some other games, and describe the theoretically most
efficient algorithm for the problem known so far. Furthermore we describe work
on more practically oriented algorithms following the paradigm of strategy im-
provement, for which a theoretical analysis stays elusive so far.
Recall that in a parity game we are given a (finite) graph with vertices labeled
by natural numbers. The vertex set is partitioned into vertices in which Player 0
moves and vertices in which Player 1 moves. In an initialized game we are also
given a starting vertex. In a play of the game a token is placed on the starting
vertex and is then moved over the graph by Player 0 and Player 1, each making
their move if the token is on one of their vertices. For simplicity we assume
that the graph is bipartite, so that each move from a Player 1 vertex leads to a
Player 0 vertex and vice versa. Each player follows some strategy. If the highest
priority of a vertex occurring infinitely often in the play is odd, then Player 1
wins, otherwise Player 0 wins. See Chapter 2 for more details.
Exercise 7.1. Show that one can convert any parity game on a nonbipartite game
arena into an equivalent parity game on a bipartite arena in linear time.
It is an important (and deep) result that the players may restrict themselves
to memoryless strategies (i.e., define their strategy by picking once and for all a
neighbor for each of their vertices thus not considering the path on which they
arrive there), see Theorem 6.6 in the previous chapter. This also implies that
for each vertex one of the players has a winning strategy, so there are no draws!
Supported by NSF Grant CCR 9987854.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 107-129, 2002.
Springer-Verlag Berlin Heidelberg 2002
108 Hartmut Klauck
If the players use memoryless strategies, a play of the game leads after a finite
number of steps to a cycle in the underlying graph.
Rabin [148] first showed a complementation lemma for parity automata work-
ing on infinite trees (while providing a decidability result for a certain logic)
implicitly also proving the determinacy result for parity games. The applica-
tion of games to the complementation problem is due to Büchi [21]. Gurevich
and Harrington [77] gave an abbreviated proof of Rabin’s result, and Emerson
and Jutla [55] another simplified proof by showing equivalence to the modal
µ-calculus model checking problem (in which complementation is trivial). Their
result also implies that in fact a player who has a strategy so that he wins in an
initialized game also has a memoryless winning strategy. See also [126, 203] for
further work on these problems.
The first question arising is, of course, whether one can decide the winner
in a parity game efficiently, i.e., whether one can find the player who wins if
the play starts at a given vertex, given that this player plays optimally. We
are also interested in finding winning strategies. The aforementioned result that
the players can restrict themselves to memoryless strategies immediately implies
that the following trivial approach is successful (using exponential time): for a
given vertex go through all strategies of Player 1. For each such strategy go
through all strategies of Player 0 and check who wins. If there is a strategy of
Player 1 that wins against all strategies of Player 0 declare Player 1 the winner,
otherwise Player 0 wins. It is the main purpose of this chapter to review some
more efficient algorithms solving this problem.
Why are we interested in this problem? There are at least two reasons. One
is that the problem is deeply related to several important topics. First of all
the problem is equivalent to the problem of modal µ-calculus model checking
[55, 56], which in turn is of large importance for computer-aided verification.
So better algorithms for the problem lead to better model checkers, making
more expressive types of logical statements about finite systems checkable by
efficient algorithms. The modal µ-calculus was defined first by Kozen in [100],
see Chapter 10. Parity games are also at the heart of an interconnection between
languages defined by automata operating on infinite trees and monadic second
order logic [183], see Chapter 12.
Another important reason to study the problem is its current complexity-
theoretic classification. It is known [92] to lie in UP ∩ co-UP (and thus “not too
far above P”, see [142]), but not known to be in P so far, and it is one of the
few natural problems so. Trying to find a polynomial algorithm for the problem
is a natural pursuit.
In this chapter we describe the best algorithm for the problem known so
far (from [93]), show that the problem is in UP ∩ co-UP (following [92]), and
discuss other promising approaches to get better algorithms for the problem,
mainly the strategy improvement approach first defined in [85], and employed
in a completely combinatorial algorithm given in [191].
Further we discuss the connection of the game to other games of infinite du-
ration played on graphs, and see that it is the least difficult to solve of a series of
such games all lying in NP∩co-NP. So it is the most natural candidate to attack!
7 Algorithms for Parity Games 109
1
t
lim inf w(vi−1 , vi ) ≥ ν.
t→∞ t i=1
We also refer to the above limit as the value that Player 0 wins from Player 1
after the play.
Exercise 7.2. Extend the above definition so that dead ends are allowed. Do this
in a way so that both games are equivalent.
110 Hartmut Klauck
[147] and [92] describe polynomial time reductions from parity games to mean
payoff games. We will show how to perform such a reduction in the next section.
So parity games are not harder to solve than mean payoff games. Again it is
known that memoryless strategies suffice for the players of mean payoff games.
Surprisingly the proof is much easier than in the case of parity games.
Theorem 7.2 ([49]). Let (A, ν, d, w) be a mean payoff game. Then Player 0 has
a winning strategy from a set of vertices iff Player 0 has a memoryless winning
strategy from that set.
More precisely, for each vertex v0 there is a number ν(v0 ), called the value
of v0 , such that
(a) Player 0 has a memoryless strategy so that for every play v0 v1 · · · in which
he follows this strategy
1
t
lim inf w(vi−1 , vi ) ≥ ν(v0 ).
t→∞ t i=1
(b) Player 1 has a memoryless strategy so that for every v0 v1 · · · in which she
follows this strategy
1
t
lim sup w(vi−1 , vi ) ≤ ν(v0 ).
t→∞ t i=1
We also refer to the above left hand side as the value that Player 0 wins from
Player 1 after the play.
The correction term (1 − λ) arises to make sure that the value of a game
using only edges of weight a is also a.
Zwick and Paterson prove that for each vertex in a discounted payoff game
one of the players has a memoryless winning strategy. We will see the reason
for this in Section 7.4. Furthermore we will see in that section that mean payoff
games can be reduced in polynomial time to discounted payoff games. Note
however that the proofs for the facts that memoryless winning strategies exist
become simpler with each game defined so far (and that such a result for a more
difficult game does not immediately imply the corresponding result for the easier
game).
The most general games we mention are the simple stochastic games
defined by Condon [40]. In these finite games the vertex set is partitioned into
three sets of vertices: vertices in which Player 0 moves, in which Player 1 moves,
and random vertices, in which a random successor is chosen, plus two vertices
in which 1 is paid by Player 0 to Player 1 resp. 0 is paid by Player 0 to Player 1
(and the game ends). The expected amount paid to Player 1 is the result of
the game. Zwick and Paterson [204] show that discounted payoff games can
be reduced to simple stochastic games. So these are the most difficult to solve
of the games considered here. Moreover they are the only games of chance we
consider! Still it is possible to decide in NP∩co-NP whether the payoff of Player 1
exceeds a certain threshold. The reduction from parity games to simple stochastic
games that results increases the game arena only by a constant factor. Using an
algorithm by Ludwig [117], which solves simple stochastic games with fan-out 2,
and the reductions we get the following corollary.
This is the best algorithm we√know if the number of different priorities as-
signed to vertices is larger than m. The algorithm is notably subexponential,
if the graph is sparse. The time bound is understood as the expected value of
the running time (over coin tosses of the algorithm) in the worst case (over all
inputs).
In this section we want to describe a relatively simple algorithm for solving parity
games, or rather mean payoff games. The approach can also be adapted to solve
discounted payoff games.
Let us consider a parity game (A, Ω) where Ω assigns d different priorities to
the vertices. Our whole objective is to decrease the dependence of the runtime on
d, see Section 6.4 for the first algorithm in this direction presented here. Actually,
for very large d our algorithms will not be better than the trivial exponential
112 Hartmut Klauck
Proof. Suppose our parity game is (A, Ω). W.l.o.g. the priorities are {0, . . . , d −
1}. The mean payoff game uses the same arena. An edge originating at a vertex v
with priority i = Ω(v) receives the weight w(v, u) = (−1)i ni . Let ν = 0. Clearly
all weights lie in the range {−nd−1, . . . , nd−1 }. This defines our mean payoff
game (A, 0, nd , w).
We claim that the value of the mean payoff game exceeds 0 for a pair of
memoryless strategies iff the same strategies lead to a play of the game in which
the highest priority vertex occurring infinitely often has an even priority.
W.l.o.g. we may assume that the players use memoryless strategies when
playing the mean payoff game, see Theorem 7.2. Then a play corresponds to a
path ending in a cycle. In the limit defining the value of the play the weights
on the initial segment before the cycle contribute zero. So the value of the game
is positive iff the sum of the weights on the cycle is positive. The weights are
from the set {−nd−1, nd−2 , −nd−3 , . . . , −n, 1}, assuming for concreteness that
d is even. Assume the weight wmax with the largest absolute value appearing
on the cycle is positive. Then the sum of the weights on the cycle is at least
wmax − (n − 1)wmax /n > 0, since there are at most n − 1 edges with weights not
equal to wmax in the cycle. The maximal weight is on an edge originating from
the vertex of highest priority, which must be even. Symmetrically if the weight
of largest absolute value is negative, the highest priority vertex must be odd.
So the mean payoff game and the parity game behave in the same way for
each pair of memoryless strategies, thus they are equivalent, and have the same
winning regions, and the same strategies lead to a win.
Now we show how to solve mean payoff games efficiently if the weights are
small.
7 Algorithms for Parity Games 113
Theorem 7.6. Given a mean payoff game (A, ν, d, w) where the arena has n
vertices and m edges, the winning region for Player 0 can be computed in time
O(n3 md) and space O(n · (log d + log n)).
Proof. It is our goal to find the values of the vertices efficiently. This immediately
gives us the winning region. Let νk (v) denote the following value: the players play
the game for k steps starting from vertex v (so they construct a path of length
k), then νk (v) denotes the sum of the edge weights traversed if both players play
optimally.
We want to compute the values ν(v) as the limit over k of the νk (v). First
let us characterize the latter value.
For every v ∈ V :
max(v,u)∈E {w(v, u) + νk−1 (u)} if v ∈ V0 ,
νk (v) = (7.1)
min(v,u)∈E {w(v, u) + νk−1 (u)} if v ∈ V1 .
Clearly ν0 (v) = 0 for all v ∈ V . Using this recursion we can easily compute
νk (v) for all v ∈ V in time O(km). Recall that we allow arithmetic operations
in unit time. Now we investigate how quickly νk (v)/k approaches ν(v).
First let us conclude the theorem from the above lemma. We compute all the
values νk (v) for k = 4n3 d. This takes time O(n3 md). All we have to store are
the νi (v) for the current i and i − 1. These are numbers of O(log(kd)) bits each,
so we need space O(n(log d + log n)).
Now we estimate ν(v) by ν (v) = νk (v)/k. Clearly
1 2nd 2nd 1
ν (v) − < ν (v) − ≤ ν(v) ≤ ν (v) + < ν (v) + .
2n(n − 1) k k 2n(n − 1)
Now ν(v) can be expressed as the sum of weights on a cycle divided by the
length of the cycle due to Theorem 7.2, and is thus a rational with denominator
1
at most n. The minimal distance between two such rationals is at least n(n−1) ,
so there is exactly one rational number of this type in our interval. It is also easy
to find this number. We can go through all denominators l from 1 to n, estimate
ν(v) as ν (v) · l/l and ν (v) · l/l, if one of these numbers is in the interval,
we have found the solution. This takes about O(n) steps.
Knowing the vector of values of the game it is easy to compute winning
strategies by fixing memoryless strategies that satisfy equation 7.1.
Proof of Lemma 7.7. It is proved in [49] that the values of vertices in a mean
payoff game and in its following finite variant are equal: the game is played as
the infinite mean payoff game, but when the play forms a cycle the play ends
and the mean value of the edges on that cycle is paid to Player 0. Also the
optimal such value can be obtained using the same memoryless strategies as in
the infinite case.
114 Hartmut Klauck
Note that the above proof uses the memoryless determinacy theorem for
mean payoff games [49].
Exercise 7.3. (1) Prove that mean payoff games and their finite variants are
equal in the above sense. Hint: Use the above idea with the stack.
(2) Use 1. to show that mean payoff games enjoy memoryless determinacy.
Corollary 7.8. Given a parity game (A, Ω) where d different priorities are as-
signed to vertices, the winning region and strategy of Player 0 can be computed
in time O(nd+2 m) and space O(d · n log n).
Exercise 7.4. Show that the following problem can be decided in polynomial
time: input is a game arena in which Player 0’s strategy is fixed (all vertices of
Player 0 have outdegree 1) plus a vertex in the arena. Is there a path from the
vertex to a cycle in which the highest priority is odd?
7 Algorithms for Parity Games 115
Furthermore since each vertex is either in Player 0’s or in Player 1’s winning
region, the same argument gives an NP algorithm for deciding Player 1’s winning
region, which is a co-NP algorithm for deciding Player 0’s winning region. Thus
parity games are solvable in NP∩co-NP. This strongly indicates that the problem
is not NP-complete, since otherwise NP would be closed under complement and
the polynomial hierarchy would collapse (see e.g. [142]).
Now we review a result by Jurdziński [92] saying that the complexity of
the problem is potentially even lower. First we define (for completeness) the
complexity class UP (see [142]).
Definition 7.9. A problem is in the class UP, if there is a polynomial time non-
deterministic Turing machine, such that for each input that is accepted exactly
one computation accepts.
Lemma 7.10. Let (A, ν, d, w) be a mean payoff game with an arena on n ver-
tices, played beginning in vertex v. Then rounding the value of the discounted
payoff game (A, ν, d, w, λ) with λ ≥ 1 − 1/(4n3 /d) to the nearest rational with
denominator smaller than n yields the value of the mean payoff game.
Lemma 7.11. The value vector ν̄ = (ν(v1 ), . . . , ν(vn )) containing the values of
vertices in a discounted payoff game equals the unique solution of the following
system of n equations
max(v,u)∈E {(1 − λ) · w(v, u) + λxu } if v ∈ V0 ,
xv = (7.2)
min(v,u)∈E {(1 − λ) · w(v, u) + λxu } if v ∈ V1 .
Proof. Let F be a function that maps a vector x̄ to the vector ȳ such that
max(v,u)∈E {(1 − λ) · w(v, u) + λxu } if v ∈ V0 ,
yv =
min(v,u)∈E {(1 − λ) · w(v, u) + λxu } if v ∈ V1 .
116 Hartmut Klauck
Then we are interested in vectors x̄ with x̄ = F (x̄), the fixed points of F . Let
||ȳ||∞ denote the maximum norm, then
∀ȳ, z̄ : ||F (ȳ) − F (z̄)||∞ ≤ λ||ȳ − z̄||∞ .
Since 0 < λ < 1 we have that F is a contracting function (with respect to the
maximum norm). Thus the limit x̄ = limn→∞ F n (0) exists and is the unique
solution to x̄ = F (x̄).
Now Player 0 can use the following strategy, provided he knows the vector
x̄ = F (x̄): at vertex v choose the neighboring vertex u that maximizes (1 −
λ)w(v, u) + λxu . Then Player 0 wins at least xv in a play starting at v. On the
other hand Player 1 may fix a strategy analogously so that her loss is also at
most xv . Thus the solution of F (x̄) = x̄ is the vector of values of the game.
Obviously this lemma leads to a UP algorithm for the solution of discounted
payoff games, if the vector of values can be described by short numbers. Then we
can just guess these numbers and verify that the equations are satisfied. What is
a short number? The number must be representable using a polynomial number
of bits in the size of the game. The size of the game is the length of a description
of the game, including edge weights and λ.
But first let us note that the strategies obtained from the system of equations
are indeed memoryless. The proof of Lemma 7.11 does not presuppose such a
memoryless determinacy result.
Corollary 7.12. Let (A, ν, d, w, λ) be a discounted payoff game. Then Player 0
[Player 1] has a winning strategy from a set of vertices iff Player 0 [Player 1]
has a memoryless winning strategy from that set.
Lemma 7.13. The solution of the equations 7.2 can be written with polynomi-
ally many bits.
Proof. Let N be the size of the binary representation of the discounted payoff
game. Let ν̄ be the unique solution of the equations. Then this vector can be
written
ν̄ = (1 − λ) · w̄ + λ · Q · ν̄,
where w̄ is a suitable vector containing weights w(v, u), and Q is a 0,1-matrix
containing only a single one per row. Note that in order to write down this
system of equations one has to know the winning strategy.
Assume that λ = a/b is a rational included in the game representation, with
integers a, b satisfying log a, log b < N . Let A = b ·I − a·Q for the identity matrix
I, then A is an integer matrix with at most two nonzero integer entries per row.
The above equation can then be rewritten
A · ν̄ = (b − a) · w̄.
Due to Cramer’s rule the solution of this system can be written as the vector
containing det Av / det A on position v where Av is obtained from A by replacing
column v with (a − b) · w̄.
7 Algorithms for Parity Games 117
Exercise 7.6. Devise an algorithm for discounted payoff games similar to the
algorithm described in Theorem 7.6.
m · d/2
) for min-parity games with n vertices, m edges, and d ≥ 2 different
priorities. An algorithm with comparable time complexity has been given by
Seidl in [161]. But as opposed to previous algorithms Jurdziński’s algorithm uses
only space polynomially depending on d, namely O(dn log n), when achieving this
time bound (note that we use the logarithmic measure for space complexity).
The algorithm is fairly simple to describe and analyze after several technical
concepts have been explained.
First note that we will apply comparisons in the following to tuples of natural
numbers, referring to their lexicographical ordering. Furthermore we will use
symbols like <i , referring to the lexicographical ordering when restricted to the
first i components of a tuple (ignoring the other components). So e.g. (2, 4, 3) <
(2, 4, 5), but (2, 4, 3) =2 (2, 4, 5). Denote [i] = {0, . . . , i − 1}.
For a technical reason in this section Player 0 wins, if the lowest priority
occurring infinitely often is even, i.e., we are considering min-parity games. The
max-parity game can obviously be reduced to this variant and vice versa. Also
we exclude dead ends from the game graph, see Exercise 2.8.
Proof. From each vertex either Player 1 or Player 0 has a winning strategy. If
Player 1 has a winning strategy, then this can be assumed to be memoryless. So
assume Player 0 plays according to f0 and consider the resulting solitaire game.
Then Player 1 can win from a vertex v iff she can fix an edge for each vertex so
that the resulting path from v ends in a simple cycle which is odd. If no such
cycle exists, Player 1 cannot win (and Player 0 wins). If such a cycle exists, then
Player 1 wins iff she can find a path to that cycle. This happens at least for all
vertices on that cycle, so there are vertices where f0 is not winning.
The key notion in the algorithm will be a parity progress measure. These are
labelings of the vertices of graphs with tuples of natural numbers having certain
properties. First we consider such labelings for solitaire games.
Definition 7.16. Let (A, Ω) be a solitaire game with vertex priorities Ω(v) ≤ d.
A function ρ : V0 ∪ V1 → INd+1 is a parity progress measure for the solitaire
game, if for all edges (v, w):
The intuition behind the above definition is best explained through the fol-
lowing lemma.
Proof. Let ρ be a parity progress measure for a solitaire game G. Suppose there
is an odd cycle v1 , . . . , vl in G, let i = Ω(v1 ) be the lowest priority on the cycle,
which is odd. Then according to the definition of a parity progress measure
ρ(v1 ) >i ρ(v2 ) ≥i · · · ≥i ρ(vl ) ≥i ρ(v1 ), which is a contradiction.
Lemma 7.18. If all simple cycles in a solitaire game G = (A, Ω) are even, then
there is a parity progress measure ρ : V → MG .
Proof. We define the parity progress measure explicitly from the solitaire game
G (as opposed to the inductive proof given in [93]). Let ai (v) be the maximal
number of vertices with priority i occurring on any path in G starting in v
and containing no vertex with priority smaller than i. This value is infinite,
if infinitely many vertices with priority i occur on some path with no smaller
priority occurring on that path. If v has priority smaller than i or there is no
path featuring a vertex with priority i but no smaller priority, then ai (v) = 0.
We then set ρ(v) = (0, a1 (v), 0, a3 (v), 0, . . . , 0, ad (v)) and claim that this is
a parity progress measure with the desired property.
First assume that some ai (v) is not finite for some odd i. Then there is an
infinite path starting at v such that the path contains no vertex with lower
priority than i, but infinitely many vertices with priority i. Thus the path must
contain some vertex with priority i twice, and we can construct a cycle with
least priority i, a contradiction to the assumption of the lemma.
Now we show that we have actually defined a mapping ρ : V → MG . As-
sume that ai (v) is larger than the number of vertices with priority i. Due to
the definition of ai (v) there is a path originating in v such that ai (v) vertices
with priority i show up before a vertex with priority smaller than i. If ai (v) is
larger than the number of vertices with priority i, such a vertex occurs twice.
Consequently there is a cycle containing as least priority i, again a contradiction.
It remains to show that we defined a parity progress measure. Let (v, w) be
any edge and i any odd number. If i = Ω(v), then ai (v) = ai (w) + 1. For all
smaller odd i we get ai (v) ≥ ai (w), because the edge (v, w) extended by a path
starting in w that contains k vertices with priority i but no smaller priority,
yields a path starting in v that contains k vertices with priority i but no smaller
priority. Thus for all v with odd priority ρ(v) >Ω(v) ρ(w) and for all v with even
priority ρ(v) ≥Ω(v) ρ(w).
Exercise 7.8. Find a parity game and a parity progress measure for which the
above intuition is not true.
Definition 7.19. Let prog(ρ, v, w) denote the least m ∈ MG such that m ≥Ω(v)
ρ(w), and, if Ω(v) is odd, then m >Ω(v) ρ(w) or m = ρ(w) = .
A function ρ : V → MG is a game progress measure, if for all v ∈ V the
following two conditions hold:
Let us explain the intuition behind the above definition. A parity progress
measure captures the existence of a winning strategy for Player 0 from all ver-
tices in a solitaire game. The key feature of a parity progress measure is that
it decreases on edges originating from vertices with odd parity and does not
increase on other edges (with respect to some order depending on the priorities
of vertices).
In a game arena (as opposed to a solitaire game) the strategy of Player 0 is
not fixed, i.e., usually vertices belonging to both players have outdegree larger
7 Algorithms for Parity Games 121
than one. Also there are usually nonempty winning regions for Player 0 and for
Player 1.
A game progress measure is defined with respect to Player 0. For each vertex
the above “decreasing” property must hold for some edge, if the vertex belongs
to Player 0, and for all edges, if the vertex belongs to Player 1. So we demand
the existence of an edge with the “decreasing” property for the multiple edges
originating in vertices belonging to Player 0. Furthermore we have introduced
the element to deal with vertices in the possibly nonempty winning region of
Player 1. Note that in case we have assigned the top element to a vertex we can-
not demand that an edge leading to that vertex decreases the progress measure.
That is the reason for introducing the complications in the prog-notation.
If we restrict a game graph with a game progress measure ρ to the vertices
in ||ρ||, we get a solitaire game with a parity progress measure. Assume that this
parity progress measure equals the one constructed in the proof of Lemma 7.18.
In this case we get the following interpretation of the game progress measure:
the component ρi (v) for some odd i and some v ∈ ||ρ|| contains the number of
times Player 1 may force Player 0 to see priority i before some smaller priority
occurs, if Player 0 tries to minimize that value and Player 1 tries to maximize it.
Unfortunately this intuition does not hold true for all possible parity progress
measures as noted before, see Exercise 7.8
It is easy to find a game progress measure by assigning to all vertices. This
measure does not tell us much. But it will turn out that we can try to maximize
the size of ||ρ|| and find the winning region of Player 0.
First we define a strategy from the measure ρ. Let f0ρ : V0 → V be a strategy
for Player 0 defined by taking for each vertex v a successor w which minimizes
ρ(w).
Proof. Restrict the game arena to the vertices in ||ρ||. If we now fix the strategy
f0ρ we get that ρ is a parity progress measure on the resulting solitaire game. This
implies that all simple cycles in the solitaire game are even (using Lemma 7.17)
and the strategy wins from all vertices in ||ρ||, if f0ρ is closed on ||ρ|| due to
Lemma 7.15. But this is true, since the strategy would violate the conditions of
its game progress measure if it would use an edge leading from ||ρ|| to a vertex
labeled in the solitaire game.
Lemma 7.21. For each parity game there is a game progress measure ρ such
that ||ρ|| is the winning region of Player 0.
Proof. Since each vertex is either in the winning region of Player 0 or of Player 1
we can assume that a winning strategy for Player 0 never leaves his winning set,
otherwise Player 1 could win after such a step. Fixing a memoryless winning
strategy with this winning region and restricting the vertices to the winning
122 Hartmut Klauck
region yields a solitaire game G containing no simple even cycle. Thus due to
Lemma 7.18 there is a parity progress measure ρ with values in MG . If we now
set ρ(v) = for all vertices outside of G we get a game progress measure as
demanded.
We are now almost done. Given a game, we have to find a game progress
measure that has a maximal number of vertices which do not have value . But
it is actually not really clear how to compute game progress measures at all,
except trivial ones.
So we take the following approach. We consider the set of all functions V →
MG . Our goal is to find one such function which is a game progress measure,
and in particular one with a maximal winning region. First we define an ordering
on these functions. Let ρ, σ be two such functions, then ρ σ, if for all v ∈ V
we have ρ(v) ≤ σ(v). If also ρ = σ, then we write ρ < σ. With this ordering we
have a complete lattice structure on our set of functions. We will define certain
monotone operators in this lattice. The game progress measure we are looking
for is the least common fixed point of these operators.
We start from a function mapping all vertices to the all zero vector and apply
the set of operators that “push the function” towards a game progress measure.
Eventually this process will actually stop at a fixed point of the operators.
The applied operators work on one vertex label only, and in the worst case
during a run of the algorithm the label of such a vertex may take on all its
possible values. But then the number of such steps is no more than n times the
number of all labels, which is n · |MG |.
Let us define the operators now.
Definition 7.22. The operator Lift(ρ, v) is defined for v ∈ V and ρ : V → MG
as follows:
ρ(u) if u = v,
Lift(ρ, v)(u) := max{ρ(v), min(v,w)∈E prog(ρ, v, w)} if u = v ∈ V0 ,
max{ρ(v), max(v,w)∈E prog(ρ, v, w)} if u = v ∈ V1 .
The following lemmas are obvious.
Lemma 7.23. For all v ∈ V the operator Lift(·, v) is monotone with respect to
the ordering .
Lemma 7.24. A function ρ : V → MG is a game progress measure iff it is a
simultaneous fixed point of all Lift(·, v) operators, i.e., iff Lift(ρ, v) ρ for all
v ∈V.
Exercise 7.9. Prove the lemmas.
Now we have a correspondence between fixed points and game progress mea-
sures. We are interested in a game progress measure inducing the winning region.
To find such a measure we will be computing the least simultaneous fixed point
of all the operators. Due to a theorem of Tarski [175] and Knaster such a least
fixed point exists and can be computed in the following way (see also Chapter 20
in the appendix):
7 Algorithms for Parity Games 123
n
) and space O(dn log n).
Proof. First let us argue that the algorithm actually finds the winning region
of Player 0. The computed game progress measure µ is the least simultaneous
fixed point of all the lift operators. The strategy f0µ induced by µ is a winning
strategy on the set of vertices ||µ|| due to Lemma 7.20. Therefore ||µ|| is a subset
of Player 0’s winning region. Furthermore ||µ|| is the largest set of vertices not
assigned over all game progress measures. Thus it must be Player 0’s winning
region due to Lemma 7.21.
Now let us calculate the complexity of the algorithm. The space is very easy
to calculate. For each vertex we have to store an element of MG , which consists
of d numbers from the set [n]. Thus space used is O(d · n log n).
The time can be bounded as follows. The Lift(ρ, v) operator can be im-
plemented in time O(d · outdegree(v)). Every vertex may be lifted at most
|MG | times, so the time is upper bounded by O(|MG | · d · v outdegree(v)) =
O(md|MG |), if we ensure that we can always find a liftable vertex in constant
time. This is possible by maintaining a queue of liftable vertices. In the begin-
ning we insert all liftable vertices. Later we get a liftable vertex out of the queue,
lift it, and test all predecessors of the vertex for liftability. Liftable vertices are
marked liftable in an array, and if they change from non-liftable to liftable they
are inserted into the queue. These operations are possible within the given time
bound.
It remains now to estimate the size of MG . First assume that priority 0 is
used, and also assume that there are vertices with priorities i for all 0 ≤ i ≤ d−1.
If some priority is missing, we can diminish the higher priorities by 2 without
changing the game. Then
d/2
|MG | = (|V2i−1 | + 1).
i=1
We have
d/2
d−1
(|V2i−1 | + 1) ≤ |Vi | ≤ n,
i=1 i=0
because there is at least one vertex with every even priority, and there are at
most n vertices. Such a product is maximized when all the factors are equal, and
can thus be bounded by
124 Hartmut Klauck
d/2
n
.
d/2
Now assume that priority 0 is not used. Then w.l.o.g. the priorities used are
{1, 2, . . . , d}. Inspection of the argument shows that it works in this case as well,
by switching the roles of the players in the proof and in the algorithm.
Now let us mention that one has indeed to specify in which order the Lift
operators are applied, leading to a possible improvement by using a suitable such
order. But Jurdziński has shown [93] that there is an example where for each
such order policy the time bound is basically tight.
Exercise 7.10. Consider the following graph H4,3 where quadratic vertices be-
long to Player 1 and all other vertices to Player 0. The numbers in the vertices
are the priorities.
1 1 1
2 2 2 2 2 2 2
8 7 8 7 8 7 8
4 4 4 4 4 4 4
3 3 3
6 6 6 6 6 6 6
5 5 5
Show that the vertices with priority 7 are lifted 44 times, no matter what
lifting policy is used. For this observe that for all vertices except those in the
second, indented layer Player 0 has a winning strategy, for all vertices in the
second layer Player 1 has a winning strategy, and hence is the label assigned
to these vertices in the least progress measure. Furthermore show that the Lift
operator increases the labels of vertices with priority 7 only to their successors.
Generalize the graph to a family of graphs H,b with (+ − 1)·(3b + 1)+ (2b + 1)
vertices and priorities from the set {1, . . . , 2+}. Show that some vertices are lifted
(b + 1) times no matter what lifting policy is used. Conclude that the running
time bound of Theorem 7.25 is basically tight, in particular that the running
time is exponential in d.
7 Algorithms for Parity Games 125
Definition 7.26. Let (A, Ω) be some parity game. Let wπ denote the vertex
with highest priority occurring infinitely often in a play π. Let Pπ denote the
set of vertices encountered in play π before wπ appears first and having higher
priority than wπ . Let l(x) denote the size of the set of vertices encountered before
wπ appears first.
The triple (wπ , Pπ , lπ ) is called the play profile of π.
A valuation is a mapping which assigns a play profile to every vertex.
A valuation is induced by a pair of strategies if it assigns to every vertex v
the play profile of the play consistent with the strategies and starting at v.
Exercise 7.11. Construct a parity game and a valuation so that no pair of strate-
gies corresponds to the valuation.
Construct a parity game and a valuation so that more than one pair of
strategies corresponds to the valuation.
Obviously not all valuations are consistent with strategy pairs. We are looking
for some nice conditions under which this is the case.
Consider the play profiles of vertices u, v with u = f0 (v) in a valuation in-
duced by strategies f0 , f1 . Call the plays originating at those vertices π(u), π(v).
Now obviously the most relevant vertex occurring infinitely often in the plays
starting at u and at v is the same. We can distinguish three cases.
(1) wπ(v) has larger or equal priority than v, but is not equal to v. In this case
Pπ(u) = Pπ(v) and lπ(u) = lπ(v) − 1.
(2) wπ(v) has smaller priority than v. In this case Pπ(v) = Pπ(u) ∪ {v} and
lπ(u) = lπ(v) − 1.
(3) wπ(v) = v. In this case Pπ(v) = ∅ and lπ(v) = 0. Furthermore Pπ(u) = ∅, since
there are no vertices on the cycle, which are more relevant than v.
∀u ∈ V ∃v ∈ V : v ∈ uE ∧ u /φ v.
Exercise 7.12. Prove the lemma. In particular, first show how to extract a strat-
egy pair from a locally progressive valuation so that the strategy pair induces
the valuation. Then show how to compute a locally progressive valuation when
given a pair of strategies. Analyze the time needed to do so.
We now have a close connection between strategy pairs and locally progressive
valuations. Our original goals were to find a way to get an “optimal” response
strategy, and a way to “improve” strategies by some simple operations. We now
define these with respect to valuations.
The first thing we need is a total ordering on the valuations. Since we assume
that no priority occurs twice, we simply take the order on the priorities.
Next we define another ordering on vertices. Let u ≺ v, if the priority of u
is smaller than the priority of v and v has even priority, and if v has smaller
priority than u and v has odd priority. So this ordering tells us how valuable
vertices are from the point of view of Player 0.
This can be extended to sets of vertices P, Q, saying that P ≺ Q if P = Q
and the highest vertex in the symmetric difference between P and Q is in Q, if
even, and in P , if odd.
Now extend the order to play profiles. Let (u, P, l) and (v, Q, r) be two play
profiles. Then (u, P, l) ≺ (v, Q, r) if u ≺ v, or if u = v and P ≺ Q or if u = v
and P = Q and [l < r iff v has odd priority].
This captures how advantageous a play profile may be for Player 0 compared
to another play profile. If the most relevant vertex is advantageous, then so is
the profile. If the most relevant vertex is the same, then the sets of vertices more
relevant but occurring only finitely often is decisive. If these are the same, then
the delay until the most relevant vertex appears decides. This is as much as we
can see from the play profile, and the profile has been designed to let us see that
much from it.
We are now able to state what optimal and improved valuations are.
Definition 7.29. A valuation φ is optimal for Player 0, if two vertices u and
v ∈ uE satisfy the progress relation u /φ v only if v is the ≺-maximal successor
of u or if φ(u) = (u, ∅, 0) and v = (u, ∅, k).
A symmetric definition optimal for Player 1. A valuation is optimal if it
is optimal for both players.
In other words, regarding the above defined value ordering the progress re-
lation increases only on optimal edges. Strategies inducing the valuation send
vertices to optimal neighbors.
Definition 7.30. A locally progressive valuation φ is improved for Player 0
in the following way: first a strategy for Player 0 is extracted from φ so that for
each vertex a successor is chosen which is maximal with respect to the ≺-order
on profiles with respect to φ, then a valuation is constructed which is compatible
with this strategy.
Note that if a locally progressive valuation is optimal for Player 0, then a
strategy for Player 0 can be extracted from the valuation by mapping each vertex
128 Hartmut Klauck
to its successor in the progress ordering. This strategy leads from each vertex
to a neighbor which is maximal in the value ordering. We can also extract a
strategy for Player 1 from the valuation. If Player 1 wins in a play from some
vertex v played as determined by those strategies, then Player 1 wins from v
also if Player 0 choses a different strategy, since this other strategy cannot lead
to vertices with a a more advantageous play profile for Player 0. Hence we can
collect the following consequences.
Exercise 7.13. Show that a locally progressive valuation that is optimal for
Player 1, and which does not change when it is improved for Player 0, is al-
ready optimal for both players.
Theorem 7.33. The above algorithm computes winning strategies for Player 0
and Player 1. It can be implemented so that each iteration runs in time O(nm).
Proof. The first statement follows from the previous lemmas. For the implemen-
tation we have to discuss the computation of an optimal valuation for Player 1
given a strategy of Player 0.
For this Player 1 fixes Player 0’s strategy and then goes in ascending order
over all the vertices in the resulting solitaire game using the “reward ordering”
≺. For such a vertex v Player 1 tests, if there is a cycle containing v and otherwise
7 Algorithms for Parity Games 129
only vertices of smaller priority. If so, then she computes the set of vertices from
which v can be reached (and thus also the cycle). Then a valuation is computed
on this component alone, and the component is removed, whereupon Player 1
continues with the next v.
To find an optimal valuation for the component from which the mentioned
cycle is reachable, notice that it is optimal for Player 1 to go to the cycle, since
v is the most profitable vertex which may occur infinitely often. It is her goal to
find a path from each vertex that reaches the cycle giving the lowest reward for
Player 0. All these computations are possible in time O(nm).
For more details see [191].
7.7 Conclusions
We have considered the problem of finding the winning regions of the two players
in a given parity game and in several other graph-based games. We have seen
that the problem can be solved in polynomial time, if the number of different
priorities assigned to vertices is only a constant.
Our interest in the problem comes from its equivalence to model checking in
the modal µ-calculus. Furthermore the problem is important as one of the few
natural problems in UP ∩ co-UP. We have shown how to prove this complexity
theoretic result. It is promising to investigate the complexity of the problem
further. One message is at least that the problem is very unlikely to be NP-
complete.
Furthermore we have discussed a simple, yet rather efficient algorithm, an
algorithm with a quadratically improved time complexity compared to the first
algorithm, and an attempt to solve the problem following the paradigm of strat-
egy improvement.
8 Nondeterministic Tree Automata
Frank Nießner
8.1 Introduction
The automaton models introduced so far mainly differ in their acceptance con-
ditions. However, they all consume infinite sequences of alphabet symbols, i.e.,
they consume ω-words. We therefore call these automata word automata. In
this chapter we define finite-state automata which process infinite trees instead
of infinite words and consequently we call them tree automata.
Automata on infinite objects, in general, play an important role in those ar-
eas of computer science where nonterminating systems are investigated. System
specifications can be translated to automata and thus questions about systems
are reduced to decision problems in automata theory. Tree automata are more
suitable than words when nondeterminism needs to be modelled.
Furthermore, there are close connections between tree automata and logical
theories, which allow to reduce decision problems in logic to decision problems
for automata. Such reductions will be thoroughly discussed in Chapter 11. Ra-
bin [148] showed decidability of monadic second-order logic using tree automata
which process infinite binary trees. The crucial part in his paper is a comple-
mentation theorem for nondeterministic finite-state automata on infinite trees.
The proof of this theorem implicitly entails determinacy of parity games. How-
ever, Büchi [21] observed that this proof can be much simplified when games
are applied explicitly. This approach was successfully implemented by numerous
authors, see for instance [77, 55]. Here, we present a game-theoretically based
proof of Rabin’s theorem according to Thomas [183] and Zielonka [203]. For this
purpose we use some results introduced in the previous chapters about infinite
games, especially the determinacy theorem for parity games.
Moreover, we consider the emptiness problem for finite-state automata on
infinite trees in terms of decidability and efficiency. These observations will be
useful in the subsequent chapter about monadic second-order logic.
The chapter is structured as follows. In Section 8.2 we introduce notations
and definitions. Section 8.3 introduces two tree automaton models which differ
in their acceptance conditions but recognize the same classes of tree languages.
We merely sketch the proof of equivalence between the two models. A game-
theoretical view on tree automata and their acceptance conditions, together
with the main results is given in Section 8.4. Then we are prepared to restate
the above-mentioned complementation theorem. The last section, Section 8.5,
discusses decidability questions of tree automata. We show that for a particular
class of tree automata it is decidable whether their recognized language is empty
or not.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 135-152, 2002.
Springer-Verlag Berlin Heidelberg 2002
136 Frank Nießner
8.2 Preliminaries
The infinite binary tree is the set T ω = {0, 1}∗ of all finite words on {0, 1}.
The elements u ∈ T ω are the nodes of T ω where ε is the root and u0, u1 are the
immediate (say) left and right successors of node u.
We restrict ourselves to binary trees, since they are sufficient for most appli-
cations, see, for instance, Chapter 12.
Let u, v ∈ T ω , then v is a successor of u, denoted by u < v, if there exists a
w ∈ T ω such that v = uw.
An ω-word π ∈ {0, 1}ω is called a path of the binary tree T ω . The set
P re< (π) ⊂ {0, 1}∗ of all prefixes of path π (linearly ordered by <) describes the
set of nodes which occur in π.
For sets Θ, Σ and a mapping µ : Θ → Σ, we define the infinity set Inf(µ) =
{σ ∈ Σ | µ−1 (σ) is an infinite set}.
We consider here trees where the nodes are labeled with a symbol of an
alphabet. A mapping t : T ω → Σ labels trees with symbols of Σ. The set of all
Σ-labeled trees is denoted by TΣω (or TΣ for simplicity, if no confusion occurs).
Sometimes we are only interested in the labeling of a path π through t. Hence
let t|π : P re< (π) → Σ denote the restriction of the mapping t to π.
For n an integer and 1 ≤ i ≤ n, the projection onto the i-th coordinate
is the mapping pi : Σ n → Σ such that pi ((σ1 , σ2 , . . . , σn )) = σi . We extend
projections to labeled infinite trees. For a Σ1 × Σ2 -labeled tree t ∈ TΣω1 ×Σ2 ,
let p1 (t) ∈ TΣω1 be the corresponding tree labeled exclusively with elements of
Σ1 . Projections can be applied to sets as well. Thus a projection p1 (Θ) of a set
Θ ⊆ TΣω1 ×Σ2 is defined as p1 (Θ) = {p1 (t) | t ∈ Θ}.
Example 8.1. Let Σ = {a, b}, t(ε) = a, t(w0) = a and t(w1) = b, w ∈ {0, 1}∗.
Tω : ε t: a
0 1 a b
00 01 10 11 a b a b
Exercise 8.1. Prove the above-mentioned assertion that binary trees suffice to
describe the general case, i.e., describe an encoding of trees with arbitrary finite
branching as binary trees.
8 Nondeterministic Tree Automata 137
The automata seen so far processed finite or infinite sequences of alphabet sym-
bols. They consume one input symbol at a time and thereby enter a successor
state determined by a transition relation. It is obvious that we somehow have to
modify the automaton models in order to make them running on infinite trees.
Since each position in a binary tree has two successors (rather than one successor
as in infinite words) it is natural to define for a state out of a set Q and an input
symbol from Σ two successor states in the transition relation, that is, transitions
are now elements of Q × Σ × Q × Q. Computations then start at the root of
an input tree and work through the input on each path in parallel. A transition
(q, a, q1 , q2 ) allows to pass from state q at node u with input-tree label a to the
states q1 , q2 at the successor nodes u0, u1. Afterwards there may be transitions
which allow to continue from q1 and from q2 , and so on. This procedure yields a
Q-labeled tree which we call the run of an automaton on an input tree. Such a
run is successful if all the state sequences along the paths meet an acceptance
condition similar to the types of acceptance conditions known already from
sequential ω-automata.
We start with the description of a Muller tree automaton.
t: a : qI
a b qd qa
a b a b qd qd qb qd
In a similar way we can define parity tree automata, that is, we adopt the
parity condition, introduced in [55], to tree automata. It will turn out that this
automaton model is particularly useful for the solution of the complementation
problem for automata on infinite trees.
Example 8.5. We consider the tree language T = {t ∈ T{a,b} | for each path π
through t holds t|π ∈ aω ∪ (a + b)∗ bω }. The language can be recognized by a
8 Nondeterministic Tree Automata 139
parity tree automaton A that checks simultaneously whether the labels of all
paths belong to aω ∪ (a + b)∗ bω or not. Hence there is no necessity to guess a
correct path, i.e, for each state the left and right successor states will be identical.
The automaton starts in the initial state qI and changes to successor states
qb , qb if an alphabet symbol b was read and remains in qI for a symbol a, respec-
tively. We observe that reading a symbol b means we cannot have a label aω
on the corresponding path. The following initial transitions (qI , b, qb , qb ), (qI , a,
qI , qI ) belong to the transition relation ∆ of A. The automaton remains in qb if
the corresponding input is a b, i.e., (qb , b, qb , qb ) ∈ ∆, otherwise it switches both
successor states and thus (qb , a, qa , qa ) ∈ ∆. A behaves symmetrically when its
current state is qa , that is, (qa , a, qa , qa ), (qa , b, qb , qb ) ∈ ∆.
While reading a’s, A labels the nodes of his run on t with qI . An alphabet
symbol b signals that from now on the automaton has to verify (a+b)∗ bω . This is
done by using the states qa and qb which indicate that the symbol last read was
a or b, respectively. On paths which labels belong to (a + b)∗ bω the automaton
remains, from some point of time, in state qb and consumes b’s exclusively. Thus,
if we index the states by c(qa ) = 1 and c(qb ) = 2 = c(qI ), we can ensure that
only the desired trees are accepted.
Exercise 8.3. Define a Muller and a parity tree automaton recognizing the lan-
guage T = {t ∈ T{a,b} | any path through t carries only finitely many b}.
Büchi, Rabin and Streett tree automata are defined analogously, i.e., we
provide the tree automata with a Büchi, Rabin or Streett acceptance condition.
For a thorough definition of these acceptance conditions see Chapter 1. Hence
a run of one of these automata is successful if and only if for each path of the
run the corresponding acceptance condition is satisfied. Büchi tree automata
differ from the other automaton models in terms of their generative capacity,
i.e., they differ in terms of the language class recognized. We state this fact in
the following theorem.
Theorem 8.6. Büchi tree automata are strictly weaker than Muller tree au-
tomata in the sense that there exists a Muller tree automaton recognizable lan-
guage which is not Büchi tree automaton recognizable [149].
Proof. The language T = {t ∈ T{a,b} | any path through t carries only finitely
many b} can obviously be recognized by a Muller tree automaton with transitions
(qI , a, qI , qI ), (q1 , a, qI , qI ), (qI , b, q1 , q1 ), (q1 , b, q1 , q1 ) and the designated set F =
{{qI }}. (This solves one part of the above exercise.) However, it can not be
recognized by any Büchi tree automaton.
Assume for contradiction that T is recognized by a Büchi tree automaton B =
(Q, Σ, ∆, qI , F ) such that card(Q) = n. Consider the input tree tn ∈ T{a,b} which
has a label b exactly at the nodes 1+ 0, 1+ 01+ 0, . . . , (1+ 0)n , i.e., at positions that
we reach by choosing the left successor after a sequence of right successors, but
only for at most n left choices. It is obvious that tn ∈ T . Thus there is a successful
run of B on tn . On path 1ω a final state is visited infinitely often, hence there
must be a natural number m0 so that (1m0 ) ∈ F . The same observation holds
140 Frank Nießner
for path 1m0 01ω with m1 and (1m0 01m1 ) ∈ F . Proceeding in this way we obtain
n + 1 positions 1m0 , 1m0 01m1 , . . . , 1m0 01m1 0 . . . 1mn on which runs through a
final state. This means that there must be positions, say u and v, where u < v
and (u) = (v) = f ∈ F . We consider the finite path πu in tn from u to v.
By construction this path performs at least one left turn and thus it contains a
node with label b. Now we construct another input tree tn by infinite repetition
of πu . This tree contains an infinite path which carries infinitely many b’s, thus
tn ∈ T , but we can easily construct a successful run on tn by copying the actions
of to πu infinitely often, hence getting a contradiction.
One can show that Muller, parity, Rabin and Streett tree automata all ac-
cept the same class of languages. The proofs are similar to those for sequential
automata from the first chapter. This is not a surprising fact because for tree
automata the appropriate acceptance condition is applied to each path of a run
separately, i.e., to a sequence of states.
Theorem 8.7. Muller, parity, Rabin and Streett tree automata all recognize the
same tree languages.
Proof. We sketch the transformations of tree automata according to those for
word automata described in Chapter 1.
We start with transforming Muller acceptance to parity acceptance. This
transformation reuses the modified LAR construction already introduced in
Section 1.4.2. Let A = ({1, 2, . . . , n}, Σ, 1, ∆, F ) be a Muller tree automaton.
The states of the parity tree automaton A are permutations of subsets of A’s
states together with a marker " that indicates the position of the last change in
the record. If (i, a, i , i ) ∈ ∆, then for all states u"v where i is the rightmost
symbol we have to add transitions (u"v, a, u "v , u "v ) to the transition relation
set of A . The states u "v and u "v are the successor states determined by the
rules described in Section 1.4.2. If the states out of
{ u"v | |u| = i ∧ { a ∈ Σ | a v } ∈ F }
F := { G ∈ P (Q) | ∀(E, F ) ∈ Ω . G ∩ E = ∅ ∨ G ∩ F = ∅ }.
Our final transformation transforms Rabin acceptance to Muller acceptance.
Let A = (Q, Σ, ∆, qI , Ω) be a Rabin tree automaton. We define an equivalent
Muller tree automaton by A = ({1, 2, . . . , n}, Σ, 1, ∆, F ) where
F := { G ∈ P (Q) | ∃(E, F ) ∈ Ω . G ∩ E = ∅ ∧ G ∩ F = ∅ }.
Exercise 8.4. Give an example that shows that the straight-forward conversion
of Muller ω-automata to Büchi ω-automata from Chapter 1 does not work for
tree automata.
It is not difficult to prove closure under union, intersection and projection for
finite tree automata languages. We leave this as an exercise.
Exercise 8.5. Prove closure under union, intersection and projection for the class
of Muller tree automaton recognizable languages.
t. Player 0 wins the play if this infinite state sequence satisfies the acceptance
condition of A, otherwise Player 1 wins. Player 0’s goal is it to show that the state
sequences for all paths of the corresponding run meet the acceptance condition,
i.e., that A accepts t. Player 1 tries to prevent Player 0 from being the winner,
his goal is to verify the existence of a path such that the corresponding state
sequence violates the acceptance condition of A, i.e., the rejection of t by A.
Example 8.8. For our input tree t and the parity tree automaton A introduced
in Example 8.5, Figure 8.3 shows the first moves in a play of GA,t . Each arrow
is labeled with that player whose decision determines the succeeding position.
ε, a, q
I
; ε, a, qI @
; @
; @
Player 0: Player 1:
a b −→ ; ;0, a, qI 1, b, qI@
@ −→
a b a b a b a b
; ε, a, qI @ ; ε, a, qI @
; @ ; @
; 1, b,@q @ Player 0: ;0,;a, q @ Player 1:
;
; a, qI
0, @ −→ ;
I I 1, b, qI@
A@ −→
A
A
a b a b a b 10, a, qb 11, bqbA
The positions from where on Player 0 or Player 1 have to react are called
game positions. Thus a play is an infinite sequence of game positions which
alternately belong to Player 0 or Player 1. A game can be considered as an
infinite graph which consists of all game positions as vertices. Edges between
different positions indicate that the succeeding position is reachable from the
preceding one by a valid action of Player 0 or Player 1, respectively. The game
positions of Player 0 are defined by
In a game position u = (w, q), Player 0 chooses a transition τ = (q, t(w), q0 ,
q1 ) and thus determines the states belonging to the successors of w. Further-
more, by this decision a game position v = (w, τ ) of Player 1 is established. The
edge (u, v) then represents a valid move of Player 0. Now Player 1 chooses a di-
rection i ∈ {0, 1} and determines from where to proceed, i.e., Player 1 determines
wi and thus establishes u = (wi, qi ) which is again a game position of Player 0.
The edge (v, u ) represents a valid move of Player 1. The usual starting position
of a play is (ε, qI ) and thus belongs to Player 0. Now we index the game posi-
tions with the colors of the states belonging to them, i.e., c((w, q)) = c(q) and
c((w, (q, t(w), q0 , q1 ))) = c(q). The games GA,t then meet exactly the definition
of min-parity games given in Chapter 4.
Furthermore the notions of a strategy, a memoryless strategy and a
winning strategy as defined in Section 2.4 apply to the games GA,t as well. A
winning strategy of a game GA,t and a successful run ∈ TQ of the corresponding
automaton A = (Q, Σ, ∆, qI , c) are closely related.
The run keeps track of all transitions that have to be chosen in order
to accept the input tree t. For any of the nodes (w, q), w ∈ {0, 1}∗, q ∈ Q,
where (w0, q0 ) and (w1, q1 ) are the immediate successors, we can derive the
corresponding transition τ = (q, t(w), q0 , q1 ) ∈ ∆. In other words, we know for
each node w in each path π through which transition to apply. Each of these
paths is an infinite sequence of states that corresponds to a particular play of
the game GA,t . This play is won by Player 0, since the infinite state sequence
is a path of the successful run . The decisions of Player 1 determine the path
generated by the current play. Since determines for each node and each path the
correct transition, Player 0 can always choose the right transition, independently
of Player 1’s decisions, i.e., Player 0 has a winning strategy. Thus if there exists
a successful run of A on t, then Player 0 has a winning strategy.
Conversely, we can use a winning strategy f0 for Player 0 in GA,t to construct
a successful run of A on t. For each game position (w, q) of Player 0, f0
determines the correct transition τ = (q, t(w), q0 , q1 ). Player 0 must be prepared
to proceed at game position (w0, q0 ) or at game position (w1, q1 ) since he can not
predict Player 1’s decision. However, for both positions the winning strategy can
determine correct transitions such that the play can be continued to a winning
play for Player 0. Hence in we label w by q, w0 by q0 and w1 by q1 . Proceeding
in this way we obtain the entire run which is successful since it is determined
by a winning strategy of Player 0. Thus, if Player 0 has a winning strategy in
GA,t , then there exists a successful run of A on t.
We summarize these observations in the following lemma.
144 Frank Nießner
Lemma 8.9. A tree automaton A accepts an input tree t if and only if there is
a winning strategy for Player 0 from position (ε, qI ) in the game GA,t .
Remark 8.10. Let A be a parity tree automaton and t be an input tree. There
exists a winning tree for Player 1 if and only if A does not accept t.
Lemma 8.12. The tree s is a winning tree for t if and only if L(s, t)∩L(M) = ∅.
The word automaton M accepts all sequences over Σ which satisfy A’s
acceptance condition. However, we are actually interested in a tree automaton
B which recognizes T (B) = TΣω \ T (A). Thus in order to construct B, we first
of all generate a word automaton S such that L(S) = Σ \ L(M). For this
we apply Safra’s determinization construction to M as described in Chapter 3.
Actually Safra’s algorithm applies to nondeterministic Büchi-automata hence, by
the methods specified in Chapter 1, we transform M to a Büchi-automaton. Now
Safra’s construction yields a deterministic Rabin automaton that accepts L(M).
Since a Streett condition is dual to a Rabin condition, we equip the outcome of
146 Frank Nießner
Even though the proof of closure under complement is somewhat lengthy due
to some technical details, it should be much easier to understand than the original
one presented by Rabin [148]. The proof given above highly benefits from a
game theoretical view, especially from the observation, that computations of tree
automata can be interpreted as parity games. Specifically, it is the determinacy
result for this class of games that induces the aforementioned simplification.
Exercise 8.6. Prove the above claim that the unique run of a deterministic input-
free automaton is a regular tree.
t: I
1 qb 0
d b
qI 1 1
0 qd 0 d b b d
Exercise 8.7. Prove the above claim that a tree is regular if and only if it is
generated by a deterministic finite automaton with output function as described
above.
Example 8.15. Figure 8.6 presents a run of the input-free tree automaton
B where {(qI , I), (qb , b), (qd , d)} is the state set, ∆ = (((qI , I), (qd , d), (qb , b)),
((qd , d), (qd , d), (qb , b)), ((qb , b), (qb , b), (qd , d))) and (qI , I) is the initial state.
: (qI , I) t: I
(qd , d) (qb , b) d b
With respect to the emptiness problem, we now prove the following crucial
lemma.
Lemma 8.16. For each parity tree automaton A there exists an input-free tree
automaton A such that Tω (A) = ∅ if and only if A admits a successful run.
qI
qI qI
qa qd qd qb
qb
qa
qd
qb
qa qa
qb qd
qa qI qd qa qd
qd qb
Since the state set of a tree automaton is finite, the game graph of GA is
finite as well and, according to Sections 6.3 and 6.4, the winning strategies for
both players are effectively computable. This allows us to solve the emptiness
problem.
Theorem 8.19. For parity tree automata it is decidable whether their recognized
language is empty or not.
150 Frank Nießner
Corollary 8.20. If the language of a parity tree automaton is not empty, then
it contains a regular tree.
Example 8.21. Consider the finite game graph GA depicted in Figure 8.7. We
observe the absence of second components in our illustration; just consider the
second entry to be the index of the corresponding state. Furthermore, assume the
coloring c(qI , I) = 1, c(qb , b) = 2, c(qa , a) = 3 and c(qd , d) = 4. Thus a winning
strategy could determine the subgraph emphasized by solid arcs in Figure 8.8.
The regular tree generated by the subgraph is the one depicted in Figure 8.6.
Corollary 8.22. (1) The emptiness test for parity tree automata can be carried
out in time d/2
rn
O d · r2 m
d/2
where d ≥ 2 is the number of priorities used in the coloring function.
(2) The emptiness test for parity tree automata is in UP ∩ co-UP.
qI
qI qI
qa qd qd qb
qb
qa
qd
qb
qa qa
qb qd
qa qI qd qa qd
qd qb
an algorithm that computes the winning regions and the winning strategy for
Player 0. Here we should apply the best algorithm for the problem known so
far (Jurdziński’s algorithm [93]) which is thoroughly discussed in Section 7.5.
Chapter 6 also presents tight time bounds for this problem, depending on the
number of edges, vertices and colors in the game graph. Using this, we get the
above bound.
Furthermore, in Chapter 6 it is shown that solving finite parity games lies in
the complexity theoretic class UP ∩ co-UP. This proves the second claim.
Exercise 8.8. Use the above corollary to provide upper bounds for the complexity
of the emptiness problem for Rabin tree automata.
8.6 Conclusions
In this chapter we have introduced finite-state automata that are able to consume
input trees instead of unidimensional structures. We have applied the acceptance
conditions presented in Chapter 1 to our tree automata and have obtained that
the resulting models are all equivalent with regard to their acceptance capa-
bilities. Büchi tree automata are an exception; they are weaker, even in their
nondeterministic version.
Subsequently we have identified a tree automaton and its input tree with an
infinite two-person game. This was significant, since it has allowed us to benefit
from various results about infinite games, especially in the proof of closure under
152 Frank Nießner
complementation for sets which are recognizable by finite tree automata. This
complementation result is essential to prove the decidability of monadic second-
order logic and thus demonstrates the importance of tree automaton concepts.
More about this will be presented in the Chapter 12.
We have next studied the algorithmic properties of finite tree automata and
have shown decidability of the emptiness problem for parity tree automata by
again utilizing results about infinite games on finite graphs.
9 Alternating Tree Automata and Parity Games
Daniel Kirsten
9.1 Introduction
Since Büchi’s work in 1960 [17], automata play an important role in logic. Nu-
merous different notions of automata provide decision and complexity results in
various kinds of logic. Often, one develops a method to translate some given for-
mula ϕ into an appropriate finite automaton A such that L(ϕ) = L(A). Such a
translation reduces the model checking problem and the satisfiability problem in
some logic to the word problem and the emptiness problem for finite automata.
Moreover, such a translation provides algorithms to solve the model checking
and the satisfiability problems on a computer. Consequently, one is interested
in the decidability and the complexity of the word and emptiness problems of
automata.
In this chapter, we introduce the notion of alternating tree automata. They
have been introduced in [202] to get a better understanding of the modal µ-
calculus. Thus, alternating tree automata work on transition systems (Kripke
structures). We state complexity results for the word problem, the emptiness
problem and complementation.
The notion of parity games and related results play a crucial role within the
whole chapter. Parity games provide three advantages:
(1) We use parity games to define the semantics of alternating tree automata,
i.e., we define whether an automaton accepts or rejects some transition sys-
tem by the existence of a winning strategy for Player 0 in an appropriate
parity game.
(2) Parity games provide a straightforward, convenient construction to comple-
ment a given alternating tree automaton; moreover, the fact that parity
games are determined is used to prove the correctness of this construction.
(3) We use parity games to show the decidability of the word problem and the
emptiness problem. By applying Jurdziński’s result[93], we achieve strong
complexity bounds.
The reader should be familiar with parity games as introduced in Chapter 2 and
5. To prove the decidability of the emptiness problem we use various notions
of automata on infinite words such as Büchi automata, Rabin- and Streett-
automata, parity automata, and transformations between them as introduced in
Chapter 1 of this book. We also apply Safra’s construction from Chapter 3.
The results from the present chapter will be used to examine the modal
µ-calculus in Chapters 10 and 11. In particular, the complexity results of the
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 153-167, 2002.
Springer-Verlag Berlin Heidelberg 2002
154 Daniel Kirsten
word problem and emptiness problem will be used to show complexity results
for model checking and satisfiability in the modal µ-calculus.
This chapter is organized as follows: In Section 9.2, we introduce some basic
notions. In Section 9.3, we introduce alternating tree automata and their seman-
tics formally. Sections 9.4, 9.5, and 9.6 are devoted to the three main results:
The decidability and complexity of the word problem, the complementation of
alternating tree automata, and the decidability and complexity of the emptiness
problem. Some remarks and exercises close this chapter.
The main ideas presented here are due to [202]. Our complementation of
alternating tree automata is based on an idea from [137] with some extensions
to our concept of alternating tree automata.
9.2 Preliminaries
We fix a set of propositional variables P during this chapter. A transition
system is a triple S = (S, R, λ) where
• S is a set called states,
• R ⊆ S × S is a relation, and
P
• λ : P → (S) is a mapping which assigns a set of states to every proposi-
tional variable.
P
Transition systems are also known as Kripke structures. If we consider the inverse
mapping γ −1 : S → (P ), then we can regard transition systems as labeled,
directed graphs. For every variable p ∈ P and every state s ∈ λ(p), we say that
p is true in s, and for s ∈ S \ λ(p), we say that p is false in s.
For every s ∈ S, we denote
every transition system and is convenient in some proofs, for instance, to show
the closure under complement. The second method uses a more compact arena,
in particular, a finite one if the transition system in question is finite, and is
used to examine the complexity of the word problem. In Proposition 9.2, we
show that these two ways to define the semantics are equivalent.
In Section 9.3.3, we show a small lemma which is used to show the complexity
of the emptiness problem.
In Section 9.3.4, we discuss a syntactic extension to our concept of alternating
tree automata.
Note that this definition does not allow transition conditions like q1 ∧ 2q2 or
p ∧ q for p ∈ P and q ∈ Q. Below, we will explain a method to allow these more
complex transition conditions without violating our definition. An alternating
tree automaton is a tuple A = (Q, qI , δ, Ω) where
In the beginning, the automaton is in the initial state qI and inspects the
state sI of the alternating tree automaton.
Now, assume that the automaton is in the state q and it inspects the state s,
i.e., the current instance is (q, s). The automaton tries to execute the transition
condition δ(q). If δ(q) ∈ {0, 1}, δ(q) = p, or δ(q) = ¬p for some p ∈ P , then the
automaton needs not to take any action.
If δ(q) = q ∈ Q then the automaton changes into the state q , but it does not
move to another state of the transition system, i.e., the new situation is (q , s). If
δ(q) = q1 ∧q2 or δ(q) = q1 ∨q2 , then the automaton splits itself into two instances
(q1 , s) and (q2 , s). If δ(q) = 2q or δ(q) = 3q , then the automaton parses the
relation R of S. The automaton splits into several instances. These instances are
in state q and inspect the successors of s in S, i.e., for every (s, s ) ∈ R we get
an instance (q , s ). Thus, the set of new instances is {q } × sR.
The result of this process is a possibly infinite parse tree with instances as
nodes. The main question is how does this tree determine whether A accepts or
rejects the pointed transition system (S, sI ). To answer this question, we try to
develop a notion of a “successful instance”. If δ(q) is a propositional variable p
and p is true in the state s, then the instance (q, s) is successful. Similarly, if
δ(q) = ¬p and s ∈ λ(p), then the instance is successful. Conversely, if δ(q) = p
but s ∈ λ(p) (or δ(q) = ¬p but s ∈ λ(p)), then the instance is not successful. If
δ(q) = 1, then the instance succeeds, but if δ(q) = 0, then it does not succeed.
If δ(q) = q , then we have seen above that the automaton changes its state
to q , i.e., the new situation is (q , s). Straightforwardly, we simply say that the
instance (q, s) is successful iff (q , s) is successful.
If δ(q) = q1 ∧q2 , then the instance (q, s) succeeds iff both the instances (q1 , s)
and (q2 , s) succeed. If δ(q) = q1 ∨ q2 , then the instance succeeds iff at least one
of the instances (q1 , s) and (q2 , s) succeeds.
The case δ(q) = 2q is very similar to the case δ(q) = q1 ∧ q2 , above. If
δ(q) = 2q , then the instance (q, s) succeeds iff for every s ∈ sR the instance
(q , s ) succeeds. Finally, if δ(q) = 3q , then the instance (q, s) succeeds iff there
is at least one s ∈ sR such that (q , s ) succeeds.
The automaton accepts the transition system (S, sI ) iff the initial instance
(qI , sI ) succeeds.
If we try to formalize this idea of the notion of a “successful instance” then
we will encounter problems:
Example 9.1. At first, we consider several very simple alternating tree automata
with Q = {qI }.
(1) Let δ(qI ) = 2qI and Ω(qI ) = 0. Let (S, sI ) be any pointed transition system.
Player 0 has not any location in G . However, Player 1 cannot win. He looses
every finite play. He also looses every infinite play, because the only priority
is 0. Hence, the automaton accepts every pointed transition system.
(2) Let δ(qI ) = 2qI and Ω(qI ) = 1. Again, Player 0 has no location. Let (S, sI ) be
any pointed transition system with some infinite path starting at sI . Player
1 can win the game by playing along the infinite path.
Conversely, let (S, sI ) be any pointed transition system in which every path
starting from sI is finite. There are just finite plays in G . Thus, Player 1
looses every play in G .
Consequently, the automaton accepts every pointed transition system (S, sI )
which has no infinite path starting at sI .
(3) Let δ(qI ) = 3qI and Ω(qI ) = 1. This automaton accepts not any pointed
transition system.
Exercise 9.1. Construct alternating tree automata for the following languages.
(1) The language of all pointed transition systems where p is true in the desig-
nated state.
(2) The language of all pointed transition systems that have an infinite path
starting in the designated state.
(3) The language of all pointed transition systems where on each infinite path
starting in the designated state p is true only finitely often.
Exercise 9.2. Let (S, ∫I ) and (S , ∫I ) be two pointed transition systems and as-
sume ρ is a bisimulation between the two systems, that is, ρ ⊆ S × S such that
the following holds true.
Show that for every alternating tree automaton A the following is true. A accepts
(S, sI ) iff A accepts (S , sI ).
If δ(q) ∈ {0, 1} or δ(q) ∈ {p, ¬p}, then (q, s) has no successor, and we have
|(q, s)[E]| = 0.
(q,s)∈[V ]
If δ(q) ∈ Q, then every location (q, s) ∈ [V ] has exactly one successor, i.e.,
|(q, s)[E]| ≤ |S | ≤ |R| + 1.
(q,s)∈[V ]
To sum up, we have (q,s)∈[V ] |(q, s)[E]| ≤ 2(|R| + 1) and |[E]| ≤ 2|Q|(|R| + 1).
Now, we can apply Jurdziński’s algorithm (Theorem 7.25, Section 7.5).
9.5 Complementation
Because parity games are determined (cf. Section 6.3), we have to show that
there is a winning strategy for Player 0 in G iff there is no winning strategy
for Player 1 in Ḡ. The argument is very similar to in the proof of Lemma 9.3.
Therefore, it is left as Exercise 9.4.
(1) Assume a winning strategy for Player 0 in G and construct a winning strategy
for Player 1 in Ḡ.
(2) Assume a winning strategy for Player 1 in Ḡ and construct a winning strategy
for Player 0 in G.
Exercise 9.5. Theorem 9.7 tells us that the languages recognizing by alternating
tree automata are closed under complementation. Show that they are closed
under intersection and union as well.
In this section, we show the decidability of the emptiness problem for alternating
tree automata. As a byproduct, we show that an alternating tree automaton A
accepts a finite pointed transition system if A accepts at least one transition
164 Daniel Kirsten
system. This result is used in Chapter 10 to show that the modal µ-calculus
has the finite model property which means that every satisfyable formula in the
modal µ-calculus has a finite model.
We fix some alternating tree automaton A = (Q, qI , δ, Ω). By Lemma 9.3, we
can assume that for every q ∈ Q the transition condition δ(q) is not inflated.
At first, we give the notion of a tile, which is a graph consisting of states
from A with various properties. We construct a parity game T from these tiles.
In the parity game T , Player 0 can use some arbitrary pointed transition system
in L(A) to construct a winning strategy. Conversely, if we assume some winning
strategy for Player 0 in T , we can construct some pointed transition system
which A accepts.
9.6.1 Tiles
A tile over Q is a graph ϑ = (Vϑ , Eϑ ) where Vϑ ⊆ Q, E ⊆ Vϑ × Vϑ and
(1) ∀q ∈ Vϑ : δ(q) = 0
(2) ¬ ∃q1 , q2 ∈ Vϑ ∃p ∈ P : δ(q1 ) = p ∧ δ(q2 ) = ¬p
(3) ∀q ∈ Vϑ : δ(q) = q1 −→ (q, q1) ∈ Eϑ
(4) ∀q ∈ Vϑ : δ(q) = q1 ∧ q2 −→ (q, q1 ) ∈ Eϑ ∧ (q, q2 ) ∈ Eϑ
(5) ∀q ∈ Vϑ : δ(q) = q1 ∨ q2 −→ (q, q1 ) ∈ Eϑ ↔ (q, q2 ) ∈ Eϑ
(6) For every cycle in (Vϑ , Eϑ ) the maximal priority of its states is even.
Note that (q, q1 ) ∈ Eϑ in (3) (and similarly in (4) and (5)) implies q1 ∈ Vϑ .
Further, note that in condition (5) it is possible that both q1 and q2 belong
to Vϑ as long as exactly one of the pairs (q, q1 ) or (q, q2 ) belongs to Eϑ . For
condition (5), it is useful that there are no inflated transition conditions in A.
A tile with port is a tuple (ϑ, q) where ϑ = (Vϑ , Eϑ ) is some tile and
q ∈ Vϑ ∩ Q3 . We denote the set of all tiles and all tiles with port by Θ and Θp ,
respectively.
We call a tile with port ϑ0 = (Vϑ0 , Eϑ0 , q0 ) and a tile ϑ1 = (Vϑ1 , Eϑ1 )
(similarly tile with port ϑ1 = (Vϑ1 , Eϑ1 , qϑ1 )) concatenable iff − →
q0 ∈ Vϑ1 and
−−−−−−→
Vϑ0 ∩ Q2 ⊆ Vϑ1 .
Let g = (ϑ1 , q1 ), (ϑ2 , q2 ), · · · ∈ Θω be an infinite sequence of tiles with port
where (ϑi , qi ) and (ϑi+1 , qi+1 ) are concatenable for every i ∈ ω. We define the
graph of g in a usual way:
• V := i∈ω {i} × Vi
• E := i∈ω (i, q
), (i, q
) (q , q ) ∈ Ei ∪
(i, qi ), (i + 1, −
→
qi )
i∈ω
∪ i∈ω (i, q), (i + 1, − →
q ) q ∈ Vi ∩ Q2
We call an infinite path π in (V, E) even iff the maximal priority which occurs
in π infinitely often is even. We call the sequence g even iff every infinite path
π in (V, E) is even.
There can be infinite paths π in (V, E) which get stuck in one tile, i.e., there
is some integer i such that vertices (i , q) for any i > i and any q ∈ Q do not
occur in π. These paths π are even, because of (6) in the definition of a tile.
9 Alternating Tree Automata and Parity Games 165
We denote the set of states of the automaton C by QC an its initial state by qIC .
We construct a parity game T over tiles.
The locations are V0 := QC × Θp and V1 := QC × Θ.
We define the set of moves E. For every state q C ∈ QC and every tile with
port (ϑ, q) ∈ Θp , there is a move from (q C , ϑ) ∈ V1 to (q C , ϑ, q) ∈ V0 .
Let (q C , ϑ, q) ∈ V0 , and let (q1C , ϑ1 ) ∈ V1 . There is a move from (q C , ϑ, q) to
(q1C , ϑ1 ) iff (ϑ, q) and ϑ1 are concatenable and C admits a transition from q C to
q1C via (ϑ, q). Consequently, a move of Player 0 means to construct a tile, the
state q1C is determined by the automaton C. We can imagine Player 0 and 1 as
“tile constructor” and “port selector”, respectively.
We define the priority ΩT of a location (q C , ϑ) (resp. (q C , ϑ, q)) as the priority
of the state q C in the parity automaton C.
166 Daniel Kirsten
Proof. (1) ⇒ (2) Let (S, sI ) be some pointed transition system which A accepts.
We consider the parity game G from the proof of Theorem 9.6. Let f : [V0 ] → [V ]
be a memoryless winning strategy for Player 0 in G . We construct a winning
strategy for Player 0 in T . The winning strategy which we construct is not
necessarily memoryless.
At first, we show a mechanism how Player 0 can construct tiles. He construct
tiles outgoing from some set V ⊆ Q w.r.t. some state s ∈ S. Player 0 starts his
construction with (V, ∅). He chooses some state q ∈ V , and adds new states and
edges in order to satisfy the closure properties (3), (4), (5) in the definition of a
tile. If for example δ(q) = q1 ∧ q2 , he adds two states q1 and q2 and two edges
(qI , q1 ) and (qI , q2 ) to the tile. Then, he has to take care about both q1 and q2 .
For example, let δ(q1 ) = q3 ∨ q4 . To satisfy property (5), Player 0 has to choose
between q3 and q4 . He simply calculates f (s, q1 ). If f (s, q1 ) = (s, q3 ), he adds
state q3 and the edge (q1 , q3 ) to his tile. Conversely, if f (s, q1 ) = (s, q4 ), he adds
q4 and (q1 , q4 ) to his tile.
Now, we explain a winning strategy for Player 0. At the beginning, Player
0 constructs a tile outgoing from {qI } w.r.t. sI . Let us call this tile ϑ1 . Player
0 chooses (qIC , ϑ1 ) as initial location. Next, Player 1 chooses some port, i.e., he
chooses a state from q ∈ Vϑ1 ∩ Q and moves to (qIC , ϑ1 , q).
Then, Player 0 has to move to a state/tile pair (q2C , ϑ2 ). It suffices to construct
ϑ2 , because q2C is determined by C. Let f (sI , q) = (s , q ). Player 0 constructs ϑ2
−−−−−−→
outgoing from Vϑ1 ∩ Q2 ∪ {− →q } w.r.t. s .
It is easy but technically involved to verify that this technique yields a win-
ning strategy for Player 0.
(2) ⇒ (3) Let f : V0 → V1 be a memoryless winning strategy for Player 0 in
the parity game T .
We construct a pointed transition system which A accepts. Its states are
Player 1’s locations V1 = QC × Θ. We can estimate |V1 | by |QC | |Θ|, which is
4 2 4
2O(|Q| log |Q|) · 2|Q|+|Q| , i.e., 2O(|Q| log |Q|) .
9 Alternating Tree Automata and Parity Games 167
Corollary 9.10. The problem whether some alternating tree automaton accepts
at least one pointed transition system is decidable in Exptime.
9.7 Acknowledgements
The author thanks Thomas Wilke for reading and improving a preliminary ver-
sion of this chapter.
10 Modal µ-Calculus and Alternating Tree
Automata
Júlia Zappe
10.1 Introduction
The modal µ-calculus is a logic that combines simple modal operators with
fixed point operators to provide a form of recursion. The modal µ-calculus—as
we use it today—was introduced in 1983 by Dexter Kozen [100]. It is well suited
for specifying properties of transition systems. For this reason, there is a great
interest in efficient solutions of the model checking and the satisfiability problem.
In this chapter these problems will be reduced to the corresponding prob-
lems for alternating tree automata, that is, to the problems of acceptance and
nonemptiness, respectively. This will be realised by giving a translation which
constructs for every formula ϕ an alternating tree automaton A(ϕ). Such an au-
tomaton accepts a pointed transition system iff the formula ϕ holds true in this
transition system. Solutions of the acceptance problem and the nonemptiness
problem were already given in Chapter 9.
The first comprehensive survey about the close connection between fixed
point calculi and alternating tree automata was given by Damian Niwiński in
[139]. The investigations presented in [139] are rather general and applicable to
a number of fixed point calculi. The translation described in this chapter was
introduced by Thomas Wilke in [202] and is specific to the modal µ-calculus.
P
Notation. As in the previous chapters ω denotes the set of the natural numbers.
The power set of a set S is denoted by (S). Further, we fix an enumerable set
P of propositional variables.
With regard to parity games we will use the notations and notions introduced
in Chapter 6. In addition, for a parity game G and a vertex v ∈ V G let G ↓ v
denote the subgame of G consisting of the vertices of G reachable from v, having
the same edges between them and the same priorities as in the game G.
When proving the correctness of our translation the following lemma about
parity games will be useful:
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 171-184, 2002.
Springer-Verlag Berlin Heidelberg 2002
172 Júlia Zappe
mA A
C = |{Ω (q) | q ∈ C}|
The index of the automaton is defined as the maximum of all these values, i.e.,
ind(A) = max({mA A
C | C ∈ C } ∪ {0})
10.2.1 Syntax
Definition 10.2. The set Lµ of formulas of the modal µ-calculus is inductively
defined as follows:
- ⊥, ∈ Lµ .
- For every atomic proposition p ∈ P : p, ¬p ∈ Lµ .
- If ϕ, ψ ∈ Lµ then ϕ ∨ ψ, ϕ ∧ ψ ∈ Lµ .
10 Modal µ-Calculus and Alternating Tree Automata 173
- If ϕ ∈ Lµ then 2ϕ, 3ϕ ∈ Lµ
- If p ∈ P , ϕ ∈ Lµ and p occurs in ϕ only positively then µp ϕ, νp ϕ ∈ Lµ .
Remark 10.3. Note that in the definition of Lµ negations can only be applied
to propositional variables. However, we will see that the negation of an arbi-
trary formula can easily be expressed, using de Morgan laws and the following
equivalences:
ψ1 ∨ ψ2 ↔ ¬(¬ψ1 ∧ ¬ψ2 ) , (10.1)
3ψ ↔ ¬2¬ψ , (10.2)
µp ψ ↔ ¬νp¬ψ[p/¬p] , (10.3)
where ψ[p/¬p] means that in ψ every occurrence of p is replaced by ¬p and vice
versa. We defined Lµ in this way (without using auxiliary fixed point variables),
because the translation of formulas into automata is simpler for formulas of this
form.
The operators µ and ν are called fixed point operators. They are viewed as
quantifiers. Accordingly, the set free(ϕ) of free variables of an Lµ formula ϕ is
defined inductively as follows:
- free(⊥) = free() = ∅,
- free(p) = free(¬p) = {p},
- free(ϕ ∨ ψ) = free(ϕ ∧ ψ) = free(ϕ) ∪ free(ψ),
- free(2ϕ) = free(3ϕ) = free(ϕ),
- free(µp ϕ) = free(νp ϕ) = free(ϕ) \ {p}.
The sets Fµ and Fν of µ- and ν-formulas, respectively, are defined as follows:
Fµ = {µp ψ | ψ ∈ Lµ } ,
Fν = {νp ψ | ψ ∈ Lµ } .
Formulas from the set Fη = Fµ ∪ Fν are called fixed point formulas.
10.2.2 Semantics
Formulas of the modal µ-calculus are interpreted on pointed transition systems.
The modal operators 3 and 2 have their usual meaning. In order to define the
semantics of fixed point formulas we need the Knaster–Tarski Theorem for the
special case of the power set lattice:
P
Proposition 10.4 (Knaster and Tarski). Let S be a set and g : (S) → (S) P
a function monotonic with respect to set inclusion. Then g has a least fixed
point µg and a greatest fixed point νg. These fixed points satisfy the following
equations:
µg = {S ⊆ S | g(S ) ⊆ S } ,
νg = {S ⊆ S | g(S ) ⊇ S } .
174 Júlia Zappe
The proof of this theorem and further results on fixed points are summarised in
the Appendix of this book, Chapter 20.
Definition 10.5. Let S = (S, R, λ) be a transition system. For a formula ϕ ∈ Lµ
the set ϕS ⊆ S is defined as follows:
- ⊥S = ∅, S = S ,
- pS = λ(p), ¬pS = S \ λ(p) for p ∈ P ,
- ψ1 ∨ ψ2 S = ψ1 S ∪ ψ2 S ,
- ψ1 ∧ ψ2 S = ψ1 S ∩ ψ2 S ,
- 2ψS = {s ∈ S | sR ⊆ ψS } ,
- 3ψS = {s ∈ S | sR ∩ ψS = ∅} ,
- µp ψS = {S ⊆ S | ψS[p→S ] ⊆ S } ,
- νp ψS = {S ⊆ S | ψS[p→S ] ⊇ S } .
Intuitively, ϕS denotes the set of states where ϕ holds true. For a pointed
transition system (S, s) and a formula ϕ ∈ Lµ we will write (S, s) |= ϕ for
s ∈ ϕS .
Note that µp ψS and νp ψS are the least and greatest fixed points, resp.,
of the following function:
Exercise 10.2. Prove the above claim about negating Lµ -formulas, see (10.1)–
(10.3).
Example 10.6. Let ϕ := νp0 (ψ ∧ 3p0 ). This formula can be read as follows: ϕ
holds true in the current state s0 iff ψ holds true in s0 and there is a successor s1
of s0 at which ϕ holds true. By unfolding the formula ϕ it follows that ψ holds
true at s1 and that s1 also has a successor s2 such that at s2 the formula ϕ holds
true. Since ϕ is a greatest fixed point, we may loop for ever and we obtain an
infinite path such that at each vertex on this path ψ holds true.
10 Modal µ-Calculus and Alternating Tree Automata 175
Exercise 10.3. Give an Lµ -formula ϕ such that the following holds: s ∈ ϕS iff
all paths in S starting in s are finite.
The attractor set Attr0 (G, X) for Player 0 (c. f. Section 6.2 in Chapter 6) can be
defined by the following formula:
that is, (S, v) |= ϕ iff v ∈ Attr0 (G, X). Here, we use the least fixed point operator,
because a vertex in X must be reached after a finite number of steps and thus,
the equation p = p ∨ ((p0 ∧ 3p) ∨ (¬p0 ∧ 2p )) may be applied only finitely
many times.
We give a formula ϕ that describes the set of winning positions for Player 0 in
the game G:
ϕ = ηpn η̄pn−1 ηpn−2 . . . νp0 (p ∧ pi ∧ 3pi ) ∨ (¬p ∧ pi ∧ 2pi )
i≤n i≤n
Remark 10.9. The interpretation of a formula does not depend on the interpreta-
tion of its bound variables, i.e., for a formula ϕ with p ∈
/ free(ϕ) and a transition
system S the following holds: ϕS = ϕS[p→S ] for all S ⊆ S.
For a formula ψ in G(ϕ) let SCCϕ (ψ) denote the strongly connected compo-
nent of G(ϕ) the formula belongs to.
Example 10.11. Let ϕ be the formula from Example 10.6, now letting ψ = p,
that is, ϕ := νp0 (p ∧ 3p0). By Definition 10.10, the alternation depth of this
formula is equal to 1.
10 Modal µ-Calculus and Alternating Tree Automata 177
Example 10.12. Let ϕ := νp1 (µp2 (p ∨ 3p2 ) ∧ 2p1 ). Then α(µp2 (p ∨ 3p2 )) = 1
/ free(µp2 (p ∨ 3p2 )) the alternation depth of ϕ is 1.
and because of p1 ∈
Example 10.13. Let ψ = µp1 ((p2 ∧p0 )∨p1 ) and ϕ := νp2 (3ψ). Clearly, α(3ψ) =
α(ψ) = 1. Since p2 ∈ free(ψ), it follows that α(ϕ) = 2.
A(ϕ) = (Q, qI , δ, Ω)
where
- Q := {ψ | ψ ≤ ϕ},
- qI := ϕ,
- δ : Q → TCQ is defined by:
δ(⊥) = 0, δ() = 1,
p if p ∈ free(ϕ),
δ(p) = δ(¬p) = ¬p,
ϕp if p ∈
/ free(ϕ),
δ(ψ ∧ χ) = ψ ∧ χ, δ(ψ ∨ χ) = ψ ∨ χ,
δ(2ψ) = 2ψ, δ(3ψ) = 3ψ,
δ(µp ψ) = ψ, δ(νp ψ) = ψ.
10.3.2 Correctness
We prove the correctness of the translation presented in the previous section and
start with a couple of lemma.
We will write G(ψ, S, s) for the game G(A(ψ), S, s).
Proof. We only prove the first assertion, the proof of the second assertion being
similar.
“⊆”: Let (S, s) ∈ L(A(ψ1 ∧ ψ2 )). By the definition of acceptance for alternat-
ing tree automata (cf. Chapter 9) there exists a memoryless winning strategy f
for Player 0 in the initialised parity game G(ψ1 ∧ψ2 , S, s). By Lemma 10.1, it fol-
lows that Player 0 has winning strategies for the games G(ψ1 ∧ψ2 , S, s) ↓ (ψi , s).
Since G(ψi , S, s) = G(ψ1 ∧ ψ2 , S, s) ↓ (ψi , s), the claim follows.
“⊇”: Let (S, s) ∈ L(A(ψ1 )) ∩ L(A(ψ2 )), that is, Player 0 has winning strate-
gies f01 and f02 for the games G(ψ1 , S, s) and G(ψ2 , S, s), respectively. Then
Player 0 can play in the game G(ψ1 ∧ ψ2 , S, s) as follows: the initial vertex
(ψ1 ∧ ψ2 , s) belongs to Player 1 and it has exactly two successors, namely
(ψ1 , s) and (ψ2 , s). If Player 1 chooses (ψ1 , s), then Player 0 can play in
accordance with f01 and he wins. Similarly, if Player 1 chooses the other successor
of the initial vertex, Player 0 can play in accordance with f02 and he also wins.
Therefore Player 0 wins the game G(ψ1 ∧ ψ2 , S, s). 2
Lemma 10.17. Let ψ be an Lµ formula in normal form. Then the following is
true:
L(A(2ψ)) = {(S, s) | ∀s ∈ sR : (S, s ) ∈ L(A(ψ)),
L(A(3ψ)) = {(S, s) | ∃s ∈ sR : (S, s ) ∈ L(A(ψ))}.
Since
(S, s) |= µp ψ iff s ∈ µg
and
“⊆”: Let Sµ abbreviate the right hand side of the above equation, i.e., Sµ is the
set of winning positions
of Player 0 in the game G(µp ψ, S, s).
Since µg = {S ⊆ S | g(S ) ⊆ S }, it suffices to show that g(Sµ ) ⊆ Sµ . Let
s ∈ g(Sµ ), that is (cf. 10.4), Player 0 has a memoryless winning strategy f in
the game G(ψ, S[p → Sµ ], s). We must prove (cf. definition of Sµ ) that Player 0
wins the game G(µpψ, S, s), too. In order to show this we study this game more
carefully.
The game G(µp ψ, S, s) has initial vertex (µp ψ, s) and this vertex has an
outgoing edge to the initial vertex of G(ψ, S[p → Sµ ], s), i.e., to (ψ, s).
10 Modal µ-Calculus and Alternating Tree Automata 181
Further, the game has all the vertices and edges of the game
G(ψ, S[p → Sµ ], s). In addition, every vertex of the form (p, s ) has an edge
to the initial vertex (µp ψ, s ) of the game G(µp ψ, S, s ). All vertices belong to
the same Player as in the original game, except for the vertices (p, s ) which
are now Player 0’s vertices (cf. Remark 10.15) .
Now we will show that Player 0 wins this game. At first Player 0 moves
the pebble to (ψ, s). Now he plays in accordance with his memoryless winning
strategy f for the game G(ψ, S[p → Sµ ], s) until the play reaches a vertex of
the form (p, s ) (if no vertex of this form will be reached, then Player 0 wins,
because in this case the resulting play is a play in the game G(ψ, S[p → Sµ ], s)
and f is a winning strategy for Player 0 in this game). Such a vertex was a dead
end in the game G(ψ, S[p → Sµ ], s). Since Player 0 played with f , this vertex
must have belonged to Player 1 in the game G(ψ, S[p → Sµ ], s), that is, s ∈ Sµ
(= λ[p → Sµ ](p)). By the definition of Sµ , Player 0 has a winning strategy
for G(µpψ, S, s ). Now, he can move the pebble to (µp ψ, s ) and then play in
accordance with this winning strategy and he wins.
“⊇”: For the converse we prove that every fixed point of g is a superset of
Sµ , that is,
for all S ⊆ S : if g(S ) = S then Sµ ⊆ S
Since Ω(µp ψ) is the maximum priority of the automaton A(µp ψ) and it is
odd, we have a contradiction.
Case ϕ = νp ψ. Let S and g be as in the previous case. We have to show:
νg = {s ∈ S | Player 0 wins the game G(νp ψ, S, s)} .
“⊇”: Let Sν abbreviate the set on the right hand side, i.e., let Sν be the set of
for Player 0 in the game G(νp ψ, S, s).
the winning positions
Due to νg = {S ⊆ S | S ⊆ g(S )}, it suffices to show that Sν ⊆ g(Sν ).
Let s ∈ Sν , that is, Player 0 has a memoryless winning strategy f in the game
G(νp ψ, S, s). We have to prove that Player 0 wins the game G(ψ, S[p → Sν ], s)
as well. We show that the restriction of f is a winning strategy for Player 0
in this game. Clearly, every infinite play consistent with the restriction is won
by Player 0, because such a play is a play consistent with f in the original
game. Let π be a finite play consistent with the restriction of f in the game
G(ψ, S[p → Sν ], s). The last vertex is a dead end in the game G(ψ, S[p → Sν ], s).
If this vertex is also a dead end in the original game then π is a play consistent
with f in the original game and therefore won by Player 0. If this vertex has a
successor in the original game, then it must be of the form (p, s ). As in the
previous case it follows that the restriction of f is a winning strategy for Player 0
in the game G(νp ψ, S, s) ↓ (νp ψ, s ) = G(νp ψ, S, s ). Hence, s ∈ Sν and so the
vertex (p, s ) belongs to Player 1 in the game G(ψ, S[p → Sν ], s). Therefore,
Player 0 wins also this play.
“⊆”: For the converse, let s ∈ νg. Because of νg = g(νg) for every s ∈
νg there exists a memoryless winning strategy fs for Player 0 in the game
G(ψ, S[p → νg], s ). We have to show that Player 0 wins the game G(νp ψ, S, s).
The initial vertex of the game is (νp ψ, s) and its unique successor is (ψ, s),
the initial vertex of the game G(ψ, S[p → νg], s). So, Player 0 moves the pebble
to this vertex and then he plays in accordance with fs unless a vertex (p, s1 )
is reached (in particular, the play is a play in G(ψ, S[p → νg], s) consistent with
fs ). This vertex is a dead end in the game G(ψ, S[p → νg], s). Since this play is
consistent with fs , the vertex belongs to Player 1, that is, s1 ∈ νg. The vertex
(p, s1 ) has exactly one outgoing edge that leads to the vertex (νp ψ, s1 ). This
vertex has also only one successor, namely (ψ, s1 ). Now Player 0 should play
in accordance with the strategy fs1 until a vertex (p, s2 ) is reached. After the
two trivial steps (to the vertex (νp ψ, s2 ) and then to (ψ, s2 )) he should play
as given by fs2 and so on. We want to prove that if he plays in this way, he wins.
Let π denote the resulting play.
We distinguish two cases.
First case:
∃i ∈ ω∀j > i∀s ∈ S : π(j) = (p, s )
In this case there exists a j > i such that the suffix π(j)π(j + 1) . . . of π is
an infinite play consistent with fs in a game G(ψ, S[p → νg], s ). Since fs is a
winning strategy for Player 0 in this game, he wins π.
Second case:
∀i ∈ ω∃j > i∃s ∈ S : π(j) = (p, s )
10 Modal µ-Calculus and Alternating Tree Automata 183
In this case vertices of the form (νp ψ, s ) occur infinitely often in π. Their
priority, Ω(νp ψ), is the maximum priority in the entire game, it is even and
so Player 0 wins. 2
Exercise 10.6. In the above proof, we dealt with the least fixed point operator
and the greatest fixed point operator separately. Use Theorem 9.7 to show that
it is enough to consider only one of the two fixed point operators.
and in space
O α(ϕ)|ϕ||S| log(|ϕ||S|)
where |ϕ| denotes the number of subformulas of ϕ.
The model-checking problem for the modal µ-calculus is in UP ∩ co-UP.
Observe that the time complexity bound is exponential in #α(ϕ)$/2. However,
since formulas which specify properties of transition systems are usually not too
long and have a rather small alternation depth, this complexity is practically as
good as polynomial in the size of the state space of the transition system.
The following Corollary is a straightforward consequence of Corollary 9.10.
Corollary 10.20. The satisfiability problem for the modal µ-calculus is in Ex-
ptime.
In the following chapter it will be shown that there also is a way to translate
every alternating tree automaton into an equivalent modal µ-calculus formula.
This will be used to prove that the alternation hierarchy of the modal µ-calculus
is strict. The deep connection between modal µ-calculus and monadic second-
order logic will be revealed in Chapter 14.
11 Strictness of the Modal µ-Calculus
Hierarchy
Luca Alberucci
11.1 Introduction
The two main results of this chapter are the translation of alternating automata
into the modal µ-calculus and the hierarchy theorem for the modal µ-calculus.
The first main result was initially proven by Niwinski in [139]. He introduces
automata on semi-algebras and shows the equivalence with certain fixed point
terms on so-called powerset algebras. By using the fact that on binary struc-
tures the µ-calculus corresponds to a certain powerset algebra, this result can
then be applied to the modal µ-calculus to obtain the equivalence of alternat-
ing automata and the calculus, on binary transition systems. We give a direct
translation of automata to µ-formulae, using the alternating automata model
introduced by Wilke in [202], and discussed in this volume in Chapter 9. The
translation will be such that for every automaton there is a µ-formula which is
equivalent to it on all transition systems. In this sense the new result we get is a
generalization of the original one by Niwinski, since we are not restricting ourself
to binary transition systems. By combining our result with Chapter 10, where
µ-formulae are translated to alternating automata, we get the equivalence on
all transition systems. The established equivalence is such that the alternation
depth of the µ-formulae corresponds to the index of the automata.
The second main result was first proved independently by Bradfield in [13]
and Lenzi in [112]. Our proof follows the one of Arnold in [5]. We first prove
a hierarchy theorem for alternating automata. Then, by using the equivalence
result established previously, we get the hierarchy theorem for the modal µ-
calculus.
This chapter contains five sections: Section 11.2 reviews the basic notions
which will be needed later. Section 11.3 introduces hierarchies, both for µ-
formulae and for alternating automata. In Section 11.4 we translate alternating
automata into the modal µ-calculus. And in Section 11.5 we prove the hierarchy
theorems for alternating automata and for the modal µ-calculus.
11.2 Preliminaries
With regard to alternating tree automata and the modal µ-calculus we use al-
most the same definitions and the same notation that were introduced in Chap-
ters 9 and 10.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 185-201, 2002.
Springer-Verlag Berlin Heidelberg 2002
186 Luca Alberucci
This justifies that we define that a run of an alternating tree automaton with
complex transition conditions is required to satisfy the second condition.
Exercise 11.1. Give a direct definition of run for automata with complex transi-
tion conditions, without using the notion of equivalent simple automaton. Show
that your automaton accepts the same runs as the equivalent simple automaton.
(Hint: Use the equivalence established in Remark 11.1.)
11.3 Hierarchies
In this section we introduce hierarchies both for alternating automata and for
µ-formulae.
Before we introduce a hierarchy on the formulae Lµ let us define two operators
µ and ν on classes of µ-formulae. Let Φ be a class of µ-formulae. We define µ(Φ)
to be the smallest class of formulae such that the following requirements are
fulfilled:
ν(Φ) is defined analogously to µ(Φ) with the only difference that (ii) is substi-
tuted by:
With the help of the previous definitions we introduce two modal µ-calculus
hierarchies. The first one is on the syntactical side, that is, it is a hierarchy of
classes of µ-formulae, and the second one is on the semantical side, that is, it is
a hierarchy of classes of transition systems.
For all natural numbers n we define the classes of µ-formulae Σnµ and Πnµ
inductively:
• Σ0µ and Π0µ are equal and consist of all fixed point free µ-formulae.
µ
• Σn+1 = µ(Πnµ ).
µ
• Πn+1 = ν(Σnµ ).
All Πnµ and Σnµ form the syntactic modal µ-calculus hierarchy.
To define the semantical counterpart of this hierarchy, we introduce the class
ϕ for all µ-formulae ϕ. It consists of all pointed transition systems (S, sI ) such
that sI ∈ ϕS . The semantical modal µ-calculus hierarchy consists of all
Σnµ TR and Πnµ TR , which are the following classes of pointed transition systems:
and that
This clearly shows that on the syntactical side we have a strict hierarchy of
formulae. Showing an equivalent result on the semantical side will be the second
main result of this chapter.
Proof. By Lemma 20.9 in Chapter 20 we have for all transition system S and
all formulae ϕ
For the semantical part, if A is the class of all pointed transition systems
accepted by an automaton A, we define
• ΣnTR = {A | A ∈ Σn },
• ΠnTR = {A | A ∈ Πn }.
We conclude this section by stating a lemma, which follows from the Com-
plementation Theorem 9.7.
 = TR − A,
There are µ-formulae τ1 , . . . , τk in ν(Φ) and ρ1 , . . . , ρk in µ(Φ) such that for all
transition systems S we have (where GFP(FS ) denotes the greatest fixed point
of FS and LFP(FS ) the least fixed point)
and
Proof. The proof goes by induction on the index n of the automaton. We assume
for all alternating automata A that the priority function is defined only on
strongly connected components of the transition graph. Moreover, we assume
that the cardinality of the range of the priority function of an automaton of
index n is also n. This is no real restriction, since all automata are equivalent to
one fulfilling these assumptions. There will be two cases for the induction step
(n > 0):
Case 1: If the maximal priority m is even, we will consider k auxiliary au-
tomata of index ≤ n − 1, in which the states of Ω −1 [m] are moved into variables.
Then we will apply the greatest fixed point operator.
Case 2: If the maximal priority m is odd, we consider the complement  of
our automaton A. By Lemma 11.3, Â can be chosen to have the same index as A,
but with maximal priority even. Thus, if we assume that the induction step for
Case 1 has been made, we have a Πnµ -formula τ representing the complement.
By Lemma 11.2 we know that there is a formula τA ∈ Σnµ which is equivalent to
¬τ . It is easy to check that τA is the Σnµ -formula fulfilling the requirements of
the theorem. So, only the induction step for Case 1 has to be carried out.
The informal description above shows that greatest fixed points capture the
automata with even maximal priority and the least fixed points, as negations
11 Strictness of the Modal µ-Calculus Hierarchy 191
of greatest fixed points, the automata with an odd maximal priority. Before we
carry out the induction, let us explain what means ‘moving states into variables’.
We need to define two transformations for automata:
The first takes an automaton A = (Q, P, δ, qI , Ω) and a set X ( Q such that
qI ∈ X and defines a new automaton
Af ree(X) = (Q − X, P ∪ X, δ , qI , Ω )
and
Claim. For all pointed transition systems (S, sI ) and for all i = 1. . . . , k we have
the two following facts:
To prove the ‘only if’ direction let us assume that is a qI -run on sI of the
automaton A on S. We want to convert it into a qI -run on sI of the automaton
Af ree(U ) on S . Let us do the conversion for every branch of . If we have a
branch where there is no state of U , then we do not change anything, otherwise,
when we meet the first qi ∈ U appearing in the branch, we cut off the rest.
The new end point we get is of the form (s, qi ), where by assumption (S, s) is
accepted by A with new initial state qi . Using the fact that this automaton is
equivalent to Astart(qi ) and that qi is now a variable, which by definition is true
in s ∈ S (under the valuation λ ), we get the desired result. The proof of the ‘if’
direction follows similar arguments.
Proof of (1). As before AS is the set of all points s in S such that (S, s) is ac-
cepted by A. By definition of ρi we have to prove that (ASstart(q1 ) , . . . , ASstart(qk ) )
is the greatest fixed point of FS , and so by Tarski-Knaster:
We first prove (i). Recall that the i-th component of the tuple
FS (ASstart(q1 ) , . . . , ASstart(qk ) )
is of the form
F2 : S → q1 S .
Putting all this together we obtain (Example 11.5 may be useful for a better
understanding):
• τ0 = q1 ,
3
• τ1 = q2 ∨ q1 ,
• τ2 = p1 ∧ q1 .
So we get
In the following, we only consider pointed transition systems which are rooted
trees with degree at most 2, which we will call binary transition systems. This
will be no restriction at all, because if formulas do not agree on such structures,
then they will definitely not agree on all structures.
We reduce the problem of acceptance of a binary transition system S by
a given automaton A ∈ Σn (resp. Πn ) to the acceptance of another binary
transition system GA,q (S) by TΣn (resp. TΠn ). This transition system will in
some sense be a representation of the game tree described in Chapter 9 and
which was used to define acceptance of a tree.
To define this transition system, let us introduce a more compact notation for
binary transition systems: In the sequel the symbols t1 , t2 , . . . stand for binary
trees, when no confusion arises we also use them to denote binary transition
systems. 1 is the trivial binary tree (or transition system), that is, the one with
no states. If t1 and t2 are two binary transition systems and a is a subset of the
propositional variables then a(t1 , t2 ) denotes a binary transition system with a
new root v such that exactly the variables in a are valid there and such that v
has two edges to the roots of t1 and t2 , respectively. (Observe that this means
that we do not distinguish between a(t1 , t2 ) and a(t2 , t1 ).) If v should only have
one son (resp. no son) we write a(t1 , 1) (resp. a(1, 1)). If a = {p} we also write
p(t1 , t2 ). Obviously, for any binary transition system there are a, t1 , t2 such that
it is of the form a(t1 , t2 ) (when v is chosen in the right way).
Let A = (Q, P, qI , δ, Ω) ∈ Σn (resp. Πn ) be an automaton and q a state of
A. With every binary transition system t we associate a new binary transition
system GA,q (t). The definition of this transition system is inductive, according
to the following rules.
p1
p1 p2
c1
d2 d2
∅ P ∅ ∅
c1
d2 d2
d1 ∅ d1 ∅
P P P P
Fig. 11.1. S, GA,q1 (S) and G2A,q1 (S)
Lemma 11.10. For any binary transition system S and any automaton A with
initial state q, we have:
• If A ∈ Σn :
S ∈ A ⇔ GA,q (S) ∈ TΣn .
• If A ∈ Πn :
S ∈ A ⇔ GA,q (S) ∈ TΠn .
The proof of the next lemma needs the notion of limit tree. Suppose we
have a sequence of trees (tn )n∈ω which is monotone, that is, the following holds:
For all m ∈ ω there is a n(m) ∈ ω such that for all n , n ≥ n(m) the trees tn
and tn are identical up to depth m.
In that case we can define the limit tree lim((tn )n∈ω ) of the sequence (tn )n∈ω
such that for all natural numbers m the limit tree is identical to tn(m) up to depth
m. Notice that lim((tn )n∈ω ) is well defined since (tn )n∈ω is monotone.
Lemma 11.11. Let A ∈ Σn (∈ Πn ) be an automaton. There is an automaton
A ∈ Σn (∈ Πn ) with initial state q and a transition system FA ,q such that
A = A and GA ,q (FA ,q ) = FA ,q .
11 Strictness of the Modal µ-Calculus Hierarchy 199
Proof. Let A be an automaton of the form (Q, P, qI , δ, Ω). For the semantically
equivalent automaton A we take a new state q and set
A = (Q ∪ {q }, P, q , δ , Ω )
By induction on n, with the help of the claim, we can easily prove that for all n
the trees tn and tn+1 are identical up to depth n. From that, the monotonicity
of (tn )n∈ω easily follows. We set
By definition of the limit tree we see that FA ,q is a fixed point of GA ,q , and
this completes the proof.
TR
Proof. 1. Proof by contradiction. Suppose Σn+1 = ΣnTR . By definition, it follows
that Πn ⊆ Σn . With Lemma 11.3 we get
TR TR
TR − TΣn ∈ ΣnTR .
formulae ψ1 , . . . , ψm , ¬ψm+1 , . . . , ¬ψm+k such that all ψi ∈ Πnµ and such that
3
ψ is obtained from these formulae using ∧, ∨, µ, , and substitution. Using
this representation of ψ we show that the formula is equivalent to a formula
ψ ∈ Σnµ . Hence we have ψ ∈ Σnµ TR , which is a contradiction to Corollary 11.13,
since we have ϕ = ψ.
So, let us show the equivalence of ψ to a ψ ∈ Σnµ . In the construction
of ψ we started from formulae ψ1 , . . . , ψm , ¬ψm+1 , . . . , ¬ψm+k such that all
ψi ∈ Πnµ . Since by assumption Σnµ TR = ΠnµTR , for all i ∈ {1, . . . , m} there are
formulae ψi ∈ Σnµ which are equivalent to ψi . Further, by Lemma 11.2 for all
i ∈ {m + 1, . . . , m + k} there are formulae ψi ∈ Σnµ equivalent to ¬ψi . Hence
ψ is equivalent to a formula constructed analogously starting from formulae
ψ1 , . . . , ψm
, ψm+1
, . . . , ψm+k , where all ψi ∈ Σnµ , that is ψ is obtained from
3
the ψi by using ∧, ∨, µ, , and substitution. Since Σnµ is closed under these
LTR
µ
s
; ; @@
Π3µTR ; =
s @Σ s
µTR
6@
I@
;
;6
3
@@ ; ;
;
;
; @@
Π2µTR ;; = @@ Σ
s s
µTR
6@
I@ 6
;
2
@@ ;;
@;
;; @@
Π1µTR ;; = @@ Σ
s s
µTR
I@
@
;
1
;
@@;; s
Σ0µTR = Π0µTR
Fig. 11.2. The modal µ-calculus hierarchy. Notice that the arrows stand for strict
inclusion and that LTR
µ = {ϕ | ϕ ∈ Lµ }.
12 Decidability of S1S and S2S
Mark Weyer
Overview
Inspired by an idea of Lullus, Leibniz hoped to develop a calculus for the decision
of truth in the most general sense, allowing to settle even philosophical ques-
tions by computation (his terminology was different, though). Many discoveries
of the 1930’s (e.g., Gödel’s Incompleteness Theorem [66]) implied that such a
calculus can not exist. However, this left open the undecidablility of truth in spe-
cial subtopics of mathematics. Charting the boundary between decidability and
undecidability remained and still remains an important question in many areas.
For instance, the undecidability result [121] for Hilbert’s Tenth Problem [81] was
not proved until 1970.
Proofs of undecidability of a logical formalism often work by encoding some
computationally complete calculus in the logic (which implies that the logic itself
is a computationally complete calculus). Suppose we choose Turing Machines.
A run of a Turing Machine can be described by a coloured two-dimensional
grid, where the configuration extends to the right and time extends downwards.
Therefore, whenever a logic is expressive enough to encode infinite or arbitrarily
large finite (coloured) grids, it is undecidable.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 207-230, 2002.
Springer-Verlag Berlin Heidelberg 2002
208 Mark Weyer
This applies to general first-order logic and first-order logic over finite struc-
tures because the vocabulary may be chosen to allow such an encoding rather
directly. In fact, a single binary relation suffices. On the other hand, if the class
of structures that the logic reasons about limits the interpretation of these re-
lations, decidability may become possible again. An important example is Pres-
burger arithmetic, i.e., the first-order theory of the structure (ω, +), which has
been proved to be decidable in 1929 [146]. What about second-order logic? It
is undecidable even over very limited classes of structures, because unrestricted
binary relations can be created by second-order quantification. This even holds
for universal and weak second-order logic. Hence logics with the expressiveness
of first-order logic or above can only be decidable in cases where second-order
quantification over binary relations is prohibited and no freely interpretable bi-
nary relations occur in the vocabulary. This leads to the investigation of monadic
second-order logic.
Let us gather some results.
(1) The first-order theory of arithmetic (ω, +, ·) is undecidable, which is basically
due to Gödel.
(2) The first-order theory of the field of reals (R, +, ·) is decidable, a result due
to Tarski [174].
(3) The monadic second-order theory of (R, +, ·) is undecidable, because the set
ω ( R is definable and hence arithmetic may be embedded.
(4) The first-order theory of (ω, +) is decidable, as mentioned above.
(5) The monadic second-order theory of (ω, +) is undecidable: An n × k grid
may be encoded by an appropriate subset of ω, using +1 and +2k to move
left and down.
This background leads us to monadic second-order logic over structures that
are weaker than (ω, +). An obvious choice is (ω, S0 ), where S0 is the successor
relation. We will show that the monadic second-order theory of this structure,
often called S1S, is decidable. We instead use the name S1S for monadic second-
order logic over ω-words. Decidability issues of both approaches are equivalent,
because ω-words may be regarded as colourings of the structure (ω, S0 ) and
monadic second-order logic is able to reason about colourings.
The structure (ω, S0 ) is free of rank one among the structures with one unary
function. An obvious generalization is to admit more than one unary function.
The structure that is free of rank one then is an infinite tree with fixed branching.
We show decidability in that case as well.1 Much of the research that is covered
in other chapters has been initiated by the decidability question of S1S and such
generalizations.
In Section 18.5 an extension of monadic second-order logic, called guarded
second-order logic, will be introduced. Guarded second-order logic over usual
1
The curious reader might ask about further generalizations. The monadic and weak
monadic second-order theories of infinite free structures are decidable if, and only
if, no higher arity than 1 occurs among the functions. The first-order theories of
arbitrary free structures are decidable. However, that is not among the topics of this
book.
12 Decidability of S1S and S2S 209
Suppose that E is the edge relation of an undirected graph. Then the subformula
∀x∀y P x ∧ Exy → P y
states that the set P (which is a subset of the set of vertices) is closed under
E. Hence it is the union of some of the graph’s connected components. The
subformula
∀x∀y P x ↔ P y
states that P does not separate any of the graph’s vertices. So, the whole formula
specifies that no union of connected components separates any vertices. This is
equivalent to connectivity of the graph.
In this example two different versions of quantification were used: (first-order)
quantification over vertices by ∀x . . . and ∀y . . . and (second-order) quantifica-
tion over sets of vertices by ∀P . . . . We may equivalently replace quantification
over individuals by quantification over sets containing exactly one individual
(singleton sets). This leads us to a normal form of monadic second-order logic,
which we use as a definition.
A vocabulary is a set of relational symbols, each of which has a specified
arity. A symbol R ∈ σ is called monadic if its arity is one, i.e., if it is used to
denote sets.
210 Mark Weyer
X=∅ for ∀Y X ⊆ Y
sing (x) for ¬x = ∅ ∧ ∀X (X ⊆ x → (x ⊆ X ∨ X = ∅))
x∈P for sing (x) ∧ x ⊆ P
P =Q for P ⊆Q ∧ Q⊆P
ϕ→ ψ for ¬ϕ ∨ ψ
ϕi for ϕi1 ∧ · · · ∧ ϕin
i∈I
ϕi for ϕi1 ∨ · · · ∨ ϕin
i∈I
∃x ∈ P ϕ for ∃x (x ∈ P ∧ ϕ)
∀x ∈ P ϕ for ∀x (x ∈ P → ϕ)
Note that A |= sing (P ) iff P A is a singleton set, i.e., contains exactly one element.
Moreover, we use set theoretical operations that are clearly (W)MSO-definable
such as P ∪ Q, P ∩ Q, P \ Q, and P ∪· Q.
12 Decidability of S1S and S2S 211
For a more precise treatment of what a logic is, cf. Definition 12.28 and the
subsequent remark and examples.
Definition 12.8. Define the orderings < and ≺ of word or tree positions by:
≤ := { (n, n + m) | n, m ∈ ω }
∗
:= { (w, wv) | w, v ∈ {0, 1} }
< := { (n, m) ∈ ≤ | n = m }
≺ := { (w, v) ∈ | w = v }
212 Mark Weyer
Thus < is the transitive closure of the successor relation of ω-words. It intuitively
specifies whether or not a position is closer to the initial position than another.
Similarly, ≺ is the transitive closure of the union of both successor relations. It,
in turn, specifies whether or not a position is closer to the root than another.
Exercise 12.1. Express in S1S that every occurrence of a is eventually followed
by an occurrence of b. More precisely: find an S1S-formula ϕ such that for ω-
words W we have W |= ϕ iff W has the mentioned property. You may use the
predicates < and ≤ (due to Lemma 12.11).
Lemma 12.9 (Being initially closed is (W)S1S-expressible). There is a for-
mula Incl1 (P ) ∈ MSO such that for all extended ω-words W the following are
equivalent.
(1) W |= Incl1 (P ),
(2) W |=W Incl1 (P ),
(3) y ∈ P W implies x ∈ P W for all word positions x ≤ y.
2
Lemma 12.10. Being initially closed is (W)S2S-expressible
Proof. This time, choose
Incl2 (P ) = ∀x∀y sing (x) ∧ (S0 xy ∨ S1 xy) ∧ y ∈ P → x ∈ P .
2
Lemma 12.11. The relations ≤ and < are (W)S1S-expressible, the relations
and ≺ are (W)S2S-expressible.
Proof. Observe that a ≤ b iff a is contained in all initially closed sets that contain
b iff a is contained in all initially closed finite sets that contain b. The same holds
for a b. Therefore we set
x ≤ y := sing (y) ∧ ∀P Incl1 (P ) ∧ y ∈ P → x ∈ P ,
x y := sing (y) ∧ ∀P Incl2 (P ) ∧ y ∈ P → x ∈ P ,
x < y := x ≤ y ∧ ¬x = y,
x ≺ y := x y ∧ ¬x = y.
2
Lemma 12.12 (Lexicographic ordering is (W)S2S-expressible). There is a for-
mula Lex (x, y) such that for all extended infinite binary trees T where xT and y T
are singletons, say xT = {a} and y T = {b}, the following holds. T |= Lex (x, y)
iff a precedes b in the lexicographic ordering of tree positions (viewed as words
with alphabet {0, 1}).
12 Decidability of S1S and S2S 213
The lexicographic ordering is the union of this relation and the prefix relation.
Hence we assemble
Lex (x, y) := ϕ ∨ (x ≺ y) .
2
Lemma 12.13 (Infiniteness is S1S- and S2S-expressible). There are formulae
Inf 1 (P ) ∈ MSO and Inf 2 (P ) ∈ MSO such that for ω-words W respectively in-
finite binary trees T we have W |= Inf 1 (P ) respectively T |= Inf 2 (P ) iff P W
respectively P T is infinite.
There are also formulae Fin1 (P ) ∈ MSO and Fin2 (P ) ∈ MSO expressing
finiteness of P W respectively P T .
2
Lemma 12.14. Being a path is S2S-expressible.
Proof. Observe that being a path is equivalent to being minimal among the
infinite initially closed sets. Therefore we set
Path (P ) := Inf 2 (P ) ∧ Incl2 (P ) ∧ ∀Q Inf 2 (Q) ∧ Incl2 (Q) ∧ Q ⊆ P → Q = P .
Theorem 12.15 (Büchi [18]). Büchi word automata and S1S are expressively
equivalent. Moreover, the equivalence is effective.
Theorem 12.16 (Rabin [148]). Muller tree automata and S2S are expressively
equivalent. Moreover, the equivalence is effective.
These theorems might need some explanation (for a precise formulation, see
the next four lemmata). Automata and formulae both define languages of ω-
words (infinite binary trees) over a certain alphabet. Automata do so by recogni-
tion, formulae by the satisfaction relation. Such languages are generally thought
of as properties of ω-words (infinite binary trees) that are expressed by the au-
tomata or formulae. ‘Expressive equivalence’ means that the same languages
may be defined by the two formalisms, and ‘effectiveness’ means that automata
and formulae can be translated effectively into each other.
Example 12.17. Consider the language T from Example 8.3 which is recognizable
by a Muller tree automaton. We can also give a formula ϕ ∈ S2S such that
(cf. Lemma 12.10, Lemma 12.13 and Lemma 12.14). Rabin’s Theorem states that
this is no coincidence. Whenever there is an automaton describing some property
of infinite binary trees there is also a formula describing the same property and
vice versa.
The proofs of Theorem 12.15 and Theorem 12.16 are split into the following
four lemmata. Instead of Büchi acceptance conditions we use Muller conditions.
Lemma 12.18. There is an algorithm that upon input of a Muller word au-
tomaton A produces a formula ϕA ∈ S1S such that for all ω-words W we have
W |= ϕA iff A accepts W.
Lemma 12.19. There is an algorithm that upon input of an S1S-formula ϕ
produces a Muller word automaton Aϕ such that for all ω-words W we have
W |= ϕ iff Aϕ accepts W.
Lemma 12.20. There is an algorithm that upon input of a Muller tree automa-
ton A produces a formula ϕA ∈ S2S such that for all infinite binary trees T we
have T |= ϕA iff A accepts T.
Due to the the obvious similarities we only prove the latter two lemmata.
Proof (of Lemma 12.20). Let A = (Q, Σ, qI , ∆, F ) be the given Muller tree
automaton. We construct an equivalent MSO-formula ϕA .
Note that the acceptance condition ‘there is a tree of states satisfying the
following conditions: . . . ’ already is in the form of existential second-order quan-
tification. Monadic second-order symbols R̄ = (Rq )q∈Q are used to encode the
‘tree of states’ of A. What remains is to express the ‘conditions’.
The overall shape of ϕA is
ϕA = ∃R̄ (Part ∧ Init ∧ Trans ∧ Accept) .
A tree of states that contains state q at position x is represented by a structure
T iff x ∈ RqT and x ∈ RqT for all q = q. This is formalized by
Stateq (x) := x ∈ Rq ∧ ¬x ∈ Rq .
q ∈Q\{q}
Part expresses that the Rq form a partition, i.e., that the R̄ indeed encode a tree
of states.
Init formalizes the initial condition, i.e., that the root is in state qI .
Init := ∃x StateqI (x) ∧ ∀y sing (y) → x y
The fact that the one and only run is in state q at position x is expressed by the
formula Occq (x).
Occq (x) := ∃I Incl1 (I) ∧ x ∈ I ∧ ∃R̄ Part ∧ Init ∧ Trans ∧ Stateq (x)
Theorem 12.25 (Rabin [149]). For languages L of infinite binary trees the
following are equivalent.
Using this theorem, it can easily be verified that the inclusion WS2S ⊆ S2S
is strict. The language defined by ‘there is a path containing infinitely many
b’, is Büchi-recognizable but its complement is not, cf. Chapter 8. Hence this
language is S2S-expressible but not WS2S-expressible.
These examinations are also relevant in the finite case, where we have to
replace the concept of
Theorem 12.26 (Büchi [17], Elgot [51] and Trakhtenbrot [186]). For lan-
guages L of finite words the following are effectively equivalent.
Theorem 12.27 (Thatcher and Wright [176], Doner [45]). For languages L
of finite binary trees the following are effectively equivalent.
The only part that needs further explanation is the notion of bottom-up tree
automata. It differs from Muller tree automata in the following ways.
emptiness of automata
WS1S WS2S
S1S S2S
Presburger arithmetic MSO over countable orders
SnS
12.4 Decidability
An overview over the present section is given by Figure 12.1 that charts the
decidability results ahead.
Remark 12.29. Despite the indexing, we will usually regard C and L assingle
classes and |= as a single relation. Furthermore we will assume the union L [σ]
σ
to be disjoint and the index vocabulary to be computable from the formula. This
can be achieved by replacing each ϕ ∈ L [σ] by (σ, ϕ).
Example 12.30. In the case of WS1S, C is the class of ω-words, |= is |=W , and L
is the syntax of MSO for word vocabularies. (For other vocabularies C [σ], L [σ]
and |=σ are empty.)
Example 12.31. In the case of modal µ-calculus, C is the class of pointed transi-
tion systems (S, s), |= is as in Definition 10.5, and L is the set Lµ .
Note. Section 12.1 mentions that S1S and S2S sometimes are considered as the-
ories rather than logics. Theories of structures may be regarded as the special
cases of logics in which the class C contains only one structure. Then, decidability
of tautology becomes decidability of truth in (or satisfaction by) that structure.
Theorem 12.33. (1) (W)MSO over finite words is decidable [17, 51].
(2) WS1S is decidable [18].
(3) S1S is decidable [18].
(4) (W)MSO over finite binary trees is decidable [176, 45].
(5) WS2S is decidable [148].
(6) S2S is decidable [148].
Proof. Part (2) and Part (5) are proved after Lemma 12.36 as an application
of Lemma 12.23. For the other parts, suppose a formula ϕ is given. By one
of Theorem 12.26, Theorem 12.15, Theorem 12.27 or Theorem 12.16 we may
effectively construct an automaton A such that A accepts A iff A |= ¬ϕ. The
question whether or not A |= ϕ always holds can be reduced to the question
whether or not the language of A is empty. But emptiness of all these languages
is decidable.
How do we decide emptiness of these automata?
For (1) and (4), there are simple algorithms for determining the reachable
states. It suffices to check whether or not some final state is reachable.
For (6), emptiness of the automata is covered in Chapter 8. Let us recall this
in more detail. In Section 8.3 it is indicated that a Muller tree automaton can
be transformed into a parity tree automaton. In the proof of Theorem 8.19 it is
shown how the emptiness question of this automaton can be rewritten into the
winning question of some finite parity game. Decision of the latter is covered in
Chapter 6 and Chapter 7.
For (3), we may proceed completely analogously. Some parts have already
been established in other chapters: The transformation of the automata is indi-
cated in Chapter 1. The decision of the winning question is the same as for (6)
and is covered in Chapter 6 and Chapter 7. 2
These decidability results may be further exploited by reducing decidability
questions of other logics to them.
Definition 12.34. Let (C1 , |=1 , L1 ) and (C2 , |=2 , L2 ) be logics. An effective
translation from (C1 , |=1 , L1 ) to (C2 , |=2 , L2 ) is a tuple (v, R, f ), where
Remark 12.35. As above for logics, we will usually neglect the indexing by vo-
cabularies. We will regard R as a single relation and f as a single mapping.
Furthermore the mapping v can be constructed from f and is omitted from the
notation.
Lemma 12.36. If L2 is decidable and if there is an effective translation (v, R, f )
from L1 to L2 , then L1 is decidable.
Proof. As a decision procedure for L1 , given a formula ϕ ∈ L1 [σ], apply the
decision procedure for L2 to fσ (ϕ).
For the correctness of this decision procedure, recall that fσ (ϕ) is a tautology
iff A2 |=2 fσ (ϕ) for all A2 ∈ C2 [v (σ)]. Since Rσ is onto on the right-hand side,
this is the case iff A2 |=2 fσ (ϕ) for all (A1 , A2 ) ∈ Rσ . By Condition (5) for
effective translations this is the case iff A1 |=1 ϕ for all (A1 , A2 ) ∈ Rσ . Since Rσ
is onto on the left-hand side, this is the case iff A1 |=1 ϕ for all A1 ∈ C1 [σ]. 2
So far, the proofs of Part (2) and Part (5) of Theorem 12.33 have been
omitted. Now we are able to prove these parts as a trivial application of the
lemma.
Proof. We use the effective translation (R, f ), where each R is the identity and
f is as in the proof of Lemma 12.23. 2
Definition 12.37. Let n > 2. The logic SnS for infinite trees with branching
of fixed arity n is defined in a completely analogous way to S2S.
The same holds for SωS, in which case there is one successor function Si for
every natural number i and hence ω-ary branching.
The logic SωB is different. It is MSO over structures T = (T, S T , PaT a∈Σ )
that are ω-ary branching unordered trees (of height ω), i.e., S T is the one suc-
cessor relation (the union of all successor relations in former cases). Formally we
might set
T = ω∗
S T = { (v, vα) | v ∈ ω ∗ , α ∈ ω }
Proposition 12.38. SωS is decidable.
Proof. Note that in an ω-ary tree, the n-th child of a node x is the n-th right
sibling of the 0-th child of x. Instead of 0th child, first child, second child, etc. we
may use the notions of leftmost child and next sibling. Instead of infinitely many
successor relations we only need two relations to describe tree positions. These
are identified with S0 and S1 respectively and the ω-ary tree can be embedded
in a binary tree. This lifts to an embedding of SωS-structures in S2S-structures.
cf. Figure 12.2 and Figure 12.3 for a visualization of an example with Σ = {a, b}.
More specifically, we state an effective translation (R, f ) from SωS to S2S.
For the relation R we give a bijection β from positions in ω-ary trees to the set
B of binary tree positions not starting with 1. Note that B is S2S-expressible.
β (ε) := ε
β (vn) := β (v) 01n
12 Decidability of S1S and S2S 223
0 1 2 3
b a b b
0 1 2
a b a
a
0
b a b b
1 1 1 1
0
a b a
1 1 1
P ⊆ Q → P ∩ B ⊆ Q ∩ B
P X → P (X ∩ B)
S0 XY → S0 (X
∩ B) (Y ∩ B)
Sn+1 XY → ∃z̄ zi ∈ B ∧ S0 Xz0 ∧ S1 zi zi+1 ∧ S1 zn Y
0≤i≤n 0≤i<n
224 Mark Weyer
Proof. We may use nearly the same effective translation as in the case of SωS.
The only change is that SXY has to be translated to a formula stating ‘There
are x ∈ X ∩ B and y ∈ Y such that y is of the form x01n ’. This can be done by
∀Z ∀x∀x ((x ∈ X ∩ B ∧ S0 xx ) ∨ (x ∈ Z ∧ S1 xx )) → Zx → ZY .
2
Proposition 12.40. SnS is decidable for all n > 2.
Proof. It is clear that the concept of binary tree automaton can be adapted in
such a way that all the results from Chapter 8 and Section 12.3 also hold for
the case of SnS (for finite n). A proof of Proposition 12.40 might explicate this
fact. More simply, one can find an effective translation (R, f ) from SnS to SωS,
because there is a trivial injection ι from positions of n-ary trees to positions
of ω-ary trees. The proof works similar to the one of Proposition 12.38; here, ι
assumes the role that β did formerly. 2
Definition 12.41. A structure O = O, <O is called a dense linear order
without endpoints if
Lemma 12.42 (Cantor [25]). All countable dense linear orders without end-
points are isomorphic.
Proof. Suppose O, <O and P, <P are dense linear orders without endpoints
and β is a monotone bijection from a finite subset O ( O to a finite subset
P ( P . Suppose further that o ∈ O\O is given. Due to the fact that <P is dense
and without endpoints there is an element p ∈ P \ P such that β ∪· {o → p} is a
monotone bijection between O ∪· {o} and P ∪· {p}. Similar arguments succeed,
if p ∈ P \ P is given.
Let η : ω → O ∪· P be an enumeration of O ∪· P . By a naı̈ve induction we can
construct an increasing sequence (βn )n∈ω of monotone bijections between finite
subsets of O and P , such that for all n ∈ ω it holds that η (n) is in the domain
or range of βn+1 : Let β0 be the empty bijection. Let βn+1 = βn if η (n) already
is in the domain or range of βn . Use the above extension otherwise. Then βn
n∈ω
is a monotone bijection from O to P , i.e., an isomorphism from O to P.
For a set-theoretical foundation of the naı̈ve induction we have to use the
Lemma of Zorn. 2
12 Decidability of S1S and S2S 225
Lemma 12.43. MSO is decidable over countable dense linear orders without
endpoints.
Proof. Since all countable dense linear orders without endpoints are isomorphic,
it suffices to prove that ({O} , |=, MSO) is decidable
for one countable dense
linear order without endpoints O = O, <O . We will choose as O the set of
binary tree positions and as <O the ordering from left to right. This ordering is
S2S-definable because x <O y iff x1 precedes y1 in the lexicographic ordering of
Lemma 12.12. It is straightforward to translate MSO over this order to S2S. It
remains to prove that this order is dense and without endpoints.
For lack of endpoints let x ∈ O be given. We have to construct y, z ∈ O such
that y <O x <O z. Choose y = x0, z = x1. For denseness let x, y ∈ O be given
such that x <O y. We have to construct z ∈ O such that x <O z <O y. Suppose
first that x is at a larger depth in the tree than y. In this case choose z = x1,
otherwise choose z = y0. 2
Proposition 12.44. MSO is decidable over countable linear orders.
Proof. Again, we use an effective translation (R, f ), this time to WS1S. R is the
total relation. Consequently only the definition of f remains.
We will use binary number representation to interpret natural numbers in
WS1S. A finite set N ⊆ ω of word positions encodes the natural number n given
by
n= 2i .
i∈N
P
Let V be a set of variables (for natural numbers). An ω-word w with alphabet
(V ) encodes a family (nv )v∈V of natural numbers via the sets Nv := { i < ω |
v ∈ wi }. For example, the ω-word
encodes
Na = {0, 2} na = 5
Nb = {1, 2} nb = 6
Nc = {0, 1, 3} nc = 11
Addition of numbers is specified most easily by an automaton implementing the
blackboard addition algorithm. The automaton
∅ {a}
{a,c} {b}
{b,c} {a,b,c}
{a,b}
q0 q1
{c}
expresses that na + nb = nc . Let us see how it works. If the automaton is, e.g.,
in state q1 , then a carry of 1 has to be considered in addition to the digits of
na and nb . If these digits are 0 and 1 respectively, then nc will have a digit 0 at
the considered position. Hence, the position is labelled with b and neither with
a nor c. Therefore the corresponding transition is labelled with {b} and leads
back to q1 because the next carry is again 1. The automaton can be translated
into a WS1S-formula. The remainder of the construction of f provides no more
difficulties. 2
This technique extends to the addition of real numbers:
Proposition 12.47. FO over the structure (R, +) (additive group of real num-
bers) is decidable.
Proof. We identify real numbers r with pairs (n, x), where n is an integer,
0 ≤ x < 1, and r = n + x. By writing n in two’s complement binary repre-
sentation we can identify integers with {0, 1}-words not containing both 0 and
1 infinitely often. x can be written in normal binary representation, identifying
fractional parts with {0, 1}-words containing 0 infinitely often. Examples: The
real number π is represented by the two words 110000 . . . (for 3, the usual direc-
tion is . . . 000011) and 001001000011111 . . . (for 0.1415 . . . ). The real number
−6.25 is represented by the two words 1001111 . . . (for −7) and 1100000 . . .
(for 0.75). Therefore a variable for (R, +) may be transformed into two monadic
symbols of S1S. Specifying addition using this encoding can be done similar to
the case of Presburger arithmetic. 2
Using a different proof technique, Tarski [174] proved in 1948 that even FO
over (R, +, ·) is decidable.
countable models. Therefore, we say that FO has the countable model property.
We are interested in similar special model properties for other logics. These
are closely related to the fact that the logics cannot identify every structure (up
to isomorphism). Often the decidability of a logic can be established via a special
model property.
The automata theoretic nature which underlies the proofs of many theorems
in this chapter will yield special model properties for the logics that appeared in
the previous section. It is known from Theorem 8.19 that every tree automaton
recognizing a non-empty language accepts a regular tree. (Recall that an infinite
binary tree is called regular if its labelling can be generated by a deterministic
tree automaton with output (Chapter 8), or equivalently if it contains, up to
isomorphism, only finitely many subtrees (Chapter 15).) The analogon for ω-
words is treated in the following exercise.
Exercise 12.3. Prove that every word automaton (say with parity condition)
with non-empty language accepts an eventually periodic ω-word. An ω-word v
is called eventually periodic, if it is of the form v1 v2ω . You may want to look
up the proof of Theorem 8.19.
By the equivalence of word and tree automata with S1S and S2S, we obtain
special model properties of these logics.
Proposition 12.48 (Eventually periodic model property). Each satisfiable S1S
formula is satisfied by an eventually periodic ω-word.
Proposition 12.49 (Regular model property). Each satisfiable S2S formula is
satisfied by a regular infinite binary tree.
Exercise 12.4. Prove that model checking S2S over regular trees is decidable: On
input of an S2S-formula ϕ and a regular binary tree T (given by its generating
automaton) it can be decided whether or not T |= ϕ.
Proof. For an FO-formula ϕ, let ϕ denote the formula that is obtained from ϕ
by replacing all quantifiers ∃x ( · ) and ∀x ( · ) by ∃x ∈ Q ( · ) and ∀x ∈ Q ( · ). Since
Q is not FO-definable in (R, +), ϕ is a priori not equivalent over (R, +) to any
FO-formula. Therefore we will interpret ϕ in the structure (R, Q , +). Obviously
for ā ∈ Q we have (R, Q , +, ā) |= ϕ iff (Q , +, ā) |= ϕ. As a consequence it suffices
to show that (R, +, ā) |= ϕ iff (R, Q , +, ā) |= ϕ for ā ∈ Q . This will be shown
by induction on the formula ϕ.
The claim is trivial for quantifier-free ϕ because in that case ϕ = ϕ. The
induction steps ϕ1 ∧ ϕ2 , ϕ1 ∨ ϕ2 , or ¬ϕ1 are easy as well. Let us prove the case
ϕ = ∃xψ. The case ϕ = ∀xψ can be handled as ¬∃x¬ψ, since these two formulae
are equivalent.
Suppose (R, +, ā) |= ϕ with ā ∈ Q . Then there is an a ∈ R such that
(R, +, ā, a) |= ψ. Using Lemma 12.54 there is also a b ∈ Q such that (R, +, ā, b) |=
ψ. By the induction hypothesis we may conclude that (R, Q , +, ā, b) |= ψ . Then
(R, Q , +, ā) |= ∃x ∈ Q ψ which is (R, Q , +, ā) |= ϕ .
For the converse suppose (R, Q , +, ā) |= ϕ . Then there is an a ∈ Q such
that (R, Q , +, ā, a) |= ψ . With the induction hypothesis we may conclude that
(R, +, ā, a) |= ψ. Since a ∈ R holds anyway we have (R, +, ā) |= ∃xψ which is
(R, +, ā) |= ϕ. 2
The modal µ-calculus. It was proved already in Chapter 10 that the modal µ-
calculus Lµ is decidable. In fact, this is closely related to special model properties
of Lµ and to the embedding of Lµ into MSO (for details, see also Chapter 14).
Countable model property: Each satisfiable formula in Lµ has a countable model.
Tree model property: Each satisfiable formula in Lµ has a tree model.
The countable model property holds for fixed point logics in general (for a
proof, see Chapter 18). The tree model property follows from the invariance of Lµ
under bisimulation (see Chapter 14) and the possibility to unravel any transition
system to a bisimilar tree model. Together, the two results imply that Lµ has the
countable tree model property. Since every formula in Lµ can be translated into
an equivalent MSO-formula, the decidability of Lµ follows from the decidability
of SωB.
For a related result concerning the more powerful guarded fixed point logic
µGF, see Chapter 18. Finally we prove that Lµ has the finite model property.
Proof. The essence of the proof is that the generating automaton of a regular
tree already is a pointed finite transition system.
230 Mark Weyer
{ (q, v) ∈ QU × S | q = δ (qI , v) }.
Hence (S0 , qI ) |= ϕ. 2
Obviously, this proposition also establishes the finite model property of every
modal or temporal logic that can be embedded into Lµ (such as LTL, CTL,
CTL∗ etc.) We finally remark that using the translation of Lµ into alternating
tree automata one obtains a stronger version of the tree model property and
better complexity bounds (see Chapters 9 and 10).
Tree model property with bounded degree: Each satisfiable formula ϕ ∈ Lµ has
a tree model of degree ≤ |ϕ|.
Complexity: The satisfiability problem for Lµ is Exptime-complete.
For related results in the context of guarded logics, see Chapter 19.
13 The Complexity of Translating Logic to
Finite Automata
Klaus Reinhardt
13.1 Introduction
The aim of this chapter is to show a non-elementary lower bound for the com-
plexity of translating logic to finite automata.
Here a function is elementary if it is O(hk (n)) for one of the k-fold exponential
functions hk with h0 (n) = n and hk+1 (n) = 2hk (n) .
The non-elementary lower bound is established directly by constructing a
sequence of formulas which describe counters of binary numbers whose length is
of non-elementary growth. This will show that the growth rate for corresponding
finite automata is also non-elementary.
The main motivation for this result is the question how efficiently the decid-
ability in Theorem 12.33 of Chapter 12 can be accomplished. The construction
of a finite automaton using Lemma 12.19 can lead to an exponentiation in each
step of recursion, where a negation forces to make the automaton deterministic
(nondeterminism is caused by existential quantifiers). The following Section 13.2
shows that this blowup cannot be avoided for monadic second-order logic over
finite words. This can be seen as an exercise for Section 13.3 which shows the
same even for first-order logic with < over finite words.
Furthermore, using the counters mentioned above, configurations of Turing
machines are admeasured in Section 13.4, showing that the satisfiability problem
(over word models) for first-order logic with < and also for monadic second-order
logic is complete for a non-elementary complexity class, which means that there
is no principally better method of proving decidability of WS1S (Theorem 12.33
of Chapter 12), or the satisfiability of first-order logic with < over words, than
by the standard construction of the corresponding automata.
The kind of construction used in this chapter appears in [122, 123] and [170]
in connection with picture languages and regular expressions.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 231-238, 2002.
Springer-Verlag Berlin Heidelberg 2002
232 Klaus Reinhardt
Theorem 13.1. The formula ϕA n defined above has size O(n) and defines the
language {0∗ 10F (n)−1 10∗ }, for which a finite automaton must have at least F (n)
states.
Example:
The language 0∗ 102047 10∗ is described by ϕA
4 , where the existentially quantified
C contains all binary representations of numbers from 0 to 255 having length 8.
In the figure below, a gap is inserted between these blocks to support readability.
To check the correctness of C and the block-marks in B, the formula recursively
∗ 7 ∗
uses ϕA3 describing 0 10 10 , where the corresponding C contains all binary
representations of numbers from 0 to 3 having length 2. This recursively uses
∗ ∗
ϕA A
2 describing 0 1010 , the corresponding C containing 0 and 1 finally using ϕ1 .
we use Σ>k (x) as abbreviation for the formula $k+1 (x) ∨ 0k+1 (x) ∨ ... ∨ 1n (x)
meaning that the symbol at position x is $k+1 or 0k+1 or ... or 1n .
The representations of the counters are defined inductively starting with
c1,0 := $1 01 , c1,1 := $1 11 , representing 0 and 1 on the first alphabet. Then for
example on the second alphabet c2,0 := $2 02 c1,0 02 c1,1 , c2,1 := $2 12 c1,0 02 c1,1 ,
c2,2 := $2 02 c1,0 12 c1,1 and c2,3 := $2 12 c1,0 12 c1,1 represent the numbers from 0
to 3. On the k-th alphabet ck+1,0 := $k+1 0k+1 ck,0 0k+1 ck,1 ...0k+1 ck,G(k)−1 repre-
sents 0 and in general we have
as follows: First we use the formula ϕk−1 to describe the necessary condition,
that the word contains the counters on level k − 1 in correct order. Now we can
use these counters to identify corresponding positions in the counter on level k.
This allows to define the equality of two counters starting on positions x and
y by the identity of the digit before each sub-counter representation starting
on position x in the first counter with the digit before the equal sub-counter
representation starting on position y in the second counter:
Equalk (x, y) :=
∀ x > x(($k−1 (x ) ∧ ¬∃u x < u < x ∧ $k (u)) →
∃y > y( $k−1 (x ) ∧ Equalk−1 (x , y ) ∧ ¬∃u y < u < y ∧ $k (u)∧
(0k (x − 1) ↔ 0k (y − 1)))).
Two counters are equal if the digit before equal sub-counter representations
are equal, because they are ordered by recursion, the induction starts with
Equal1 (x, y) := (01 (x + 1) ↔ 01 (y + 1)))).
Now we can define the neighbor relation Nextk (x, y) expressing that the
counter starting on position y contains the by one incremented number con-
tained in the counter starting on position x. We proceed as follows (see the
13 The Complexity of Translating Logic to Finite Automata 235
formula presented below): The first (least significant) bit always changes (line
2). For every but the first sub-counter starting on position x (line 3) there is
a corresponding sub-counter starting on position y , which represents the same
number and and which is in the second counter (line 4). The previous bits (fol-
lowed by sub-counters on the position x and y in line 5 such that there is no
other sub-counter on position u described in line 6 or 7 between them) cause a
change of the bit iff it changes from 1k to 0k (and thus causes a carry described
in line 8).
The formula
Initializek (x) := $k (x) ∧ ∃y > x($k (x) ∧ ¬∃z(x < z < y ∧ (1k (z) ∨ $k (z))))
(1) ϕk := ϕk−1 ∧
(2) ∃x(Initializek (x) ∧ ∀y < x Σ>k (y))∧
(3) ∀x($k (x) ↔ Initializek−1 (x + 2))∧
(4) ∀x($k (x) → (∃y > x( (∀u(x < u ≤ y → (Σ<k (u) ∨ 0k (u) ∨ 1k (u)))∧
(5) ∃z > y( $k (z) ∧ Nextk (x, z)∧
(6) ∀u(y < u < z → Σ>k (u))))∨
(7) (∀u(x < u ≤ y → (Σ<k (u) ∨ 1k (u)))∧
(8) ∀u > y Σ>k (u)))))∧
(9) ∀x((0k (x) ∨ 1k (x)) → $k−1 (x + 1)).
The length of the formula Equalk and thus also the formula Nextk grows
linear with k. Thus the length of ϕn is in O(n2 ). (If we count the representation
of a variable indexed by n as having length log n, we even have O(n2 log n). )
On the other hand a finite automaton recognizing the language described by ϕn
236 Klaus Reinhardt
needs at least one state for each of the G(n) counters. This means we have a
sequence of formulas ϕn , where the growth rate for the size of equivalent finite
automata is non-elementary.
Theorem 13.2. The formula ϕn defined above has size O(n2 log n) and defines
∗ ∗ ∗ ∗
the language (Σ>n cn,0 Σ>n cn,1 ...Σ>n cn,G(n)−1 )+ Σ>n , for which a finite automa-
ton must have at least G(n) states.
Exercise 13.3. Give a better lower bound for the number of states in Theorem
13.2.
since even a single increment in the input size already allows an exponential
increase in time to simulate the NSPACE-machine.
Theorem 13.5. The satisfiability problem for first-order logic with < over finite
words is complete for DSPACE(G(cn)) under polynomial time reductions.
c
Proof. For containment in the class see the proof of Lemma 12.21 (and the
following Remark 12.22) in Chapter 12, where the given formula is translated to
a finite automaton. The worst case for one step of recursion in this translation is
an exponential blowup, which occurs when the automaton is made deterministic
in order to translate negation by recognizing the complement.
To show hardness, we use the following reduction: Let L be recognized by
a deterministic one-tape Turing machine M = (Σ, Q, δ, b, q0, qf ) using G(cn)
space, with the blank symbol b ∈ Σ. A word w = w1 w2 · · · wn is accepted by
M if there is a sequence w = $C0 $C1 $ · · · $Cf of configurations over Σ ∪ (Σ ×
Q) ∪ {$} with the initial configuration C0 = (w1 , q0 )w2 w3 · · · wn b · · · b, a final
configuration Cf starting with (b, qf ) (w.l.o.g M moves to the beginning, when
it accepts,), |$Ci | = g(m) with m := cn for i ≤ f and Ci ⇒M Ci+1 for i < f .
Since the k-th symbol in $Ci+1 depends only on the k − 1-th, k-th, and k + 1-
th symbol, we can construct a first-order formula ϕδ (x, y, z, y ), which is true
iff the symbol ∈ Σ ∪ (Σ × Q) ∪ {$} at position y is the correct consequence
of (x, y, z) in the previous configuration (respecting the separation marker $).
Here y corresponds to position y in the previous configuration. For example if
(q, a)(x) and d(y) and δ(q, a) = (q , e, R), which means that the machine is in
13 The Complexity of Translating Logic to Finite Automata 237
state q on the symbol a and the consequence is that it enters state q , writes a e
and goes right, then ϕδ (x, y, z, y ) is true iff (q , d)(y ), which means that in the
following configuration the machine is in state q on symbol d. Or if (q, a)(y) and
δ(q, a) = (q , e, R), then ϕδ (x, y, z, y ) is true iff e(y ), which means that in the
following configuration there is the e, which was written by the machine (but
the machine has moved away). Since δ is finite, ϕδ is a finite formula as well.
Now we extend the construction in the previous m+1 section in the following way:
Let Σm+1 := Σ ∪ (Σ × Q) ∪ {$} and Σ>k = i=k+1 Σi . Instead of describing
where each symbol is followed by a counter containing its position. We use the
following formula:
(1) ϕM(w) := ϕm ∧ $(1) ∧ InitializeCw ∧
(2) ∀x($(x) ↔ Initializem (x + 1))∧
(3) ∀x(Σm+1 (x) ↔ $m (x + 1))∧
(4) ∀x, y, z( (Σm+1 (x) ∧ Σm+1 (y) ∧ Σm+1 (z)∧
(5) ¬∃u(x < u < z ∧ u = y ∧ (Σm+1 (u)) →
(6) (∃y > z( Equalm (y + 1, y + 1) ∧ ϕδ (x, y, z, y )∧
(7) ¬∃u(z < u < y ∧ Equalm (y + 1, u)))∨
(8) (¬∃y > z(Equalm (y + 1, y + 1))∧
Here line 2 says that the separation marker $ is exactly at those positions which
are followed by the counter representation cm,0 . Line 3 says that each symbol of
the configuration in followed by a counter, line 4 says that for all triples x, y, z
of symbols of a configuration, which are (line 5) subsequent in the configuration,
which means there are only symbols of the counter in-between, there is (line
6) a position y followed by the same counter as y with the symbol, which is
determined by δ. Line 7 makes sure that it is indeed the following configuration.
The alternative of line 8 is that there is no following configuration and (line 9)
the current configuration is a final configuration Cf . Line 1 makes sure that the
counters work in the correct way according to the previous section and the first
configuration is $C0 , which is expressed by
InitializeCw := ∃ x1 < x2 < ... < xn < y
( (w1 , q0 )(x1 ) ∧ w2 (x2 ) ∧ w3 (x3 ) ∧ ...wn (xn ) ∧ $(y)∧
∀u < y(∃i u = xi ∨ Σ≤m (u) ∨ (b(u) ∧ xn < u)))
where line 1 and 2 define the positions occupied by the input symbols and
line 3 says that all other symbols are either symbols of the counter or blank
symbols filling the tape after the input w (this excludes the $). Thus the size
of InitializeCw is linear. According to the previous section, the formula ϕm and
238 Klaus Reinhardt
thus also ϕM(w) has a size of O(m2 log m) = O(n2 log n) and can on input w be
written in polynomial time. The machine M accepts w iff ϕM(w) is satisfiable.
2
Corollary 13.6. Satisfiability of first-order formulas with < over finite words
is in no elementary space-bounded complexity class.
Philipp Rohde
14.1 Introduction
We consider monadic second order logic (MSO) and the modal µ-calculus (Lµ )
over transition systems (Kripke structures). It is well known that every class
of transition systems which is definable by a sentence of Lµ is definable by a
sentence of MSO as well. It will be shown that the converse is also true for an
important fragment of MSO: every class of transition systems which is MSO-
definable and which is closed under bisimulation – i.e., the sentence does not
distinguish between bisimilar models – is also Lµ -definable. Hence we obtain the
following expressive completeness result: the bisimulation invariant fragment of
MSO and Lµ are equivalent. The result was proved by David Janin and Igor
Walukiewicz. Our presentation is based on their article [91]. The main step is the
development of automata-based characterizations of Lµ over arbitrary transition
systems and of MSO over transition trees (see also Chapter 16). It turns out that
there is a general notion of automaton subsuming both characterizations, so we
obtain a common ground to compare these two logics. Moreover we need the
notion of the ω-unravelling for a transition system, on the one hand to obtain
a bisimilar transition tree and on the other hand to increase the possibilities of
choosing successors.
We start with a section introducing the notions of transition systems and
transition trees, bisimulations and the ω-unravelling. In Section 14.3 we repeat
the definitions of MSO and Lµ . In Section 14.4 we develop a general notion of
automaton and acceptance conditions in terms of games to obtain the charac-
terizations of the two logics. In the last section we will prove the main result
mentioned above.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 239-257, 2002.
Springer-Verlag Berlin Heidelberg 2002
240 Philipp Rohde
R
• •
• • • • • •
Exercise 14.1. Show that ∼ is an equivalence relation on the class of all transition
systems.
The main property of the ω-unravelling for our purpose is, that we always
have enough possibilities to choose a different r-successor for finitely many r-
In other words: Let t be an r-successor of s in
successors of an element in M.
M and let u be an ω-path to s. Then there are infinitely many r-successors of
which are bisimilar to t.
u in M
M
r1
s1
M s0 M
r3 M
r4
M
r2 M
s2 r0
•···
5
• 0 •
5 3
•···
Mc •
• 9 •···
2 1
• •···
9
5 •
0
•···
Fig. 14.2. A transition system M with source s0 = srM and a part of its ω-unravelling
c
M (we suppressed the labelling of the nodes). Notice that in fact every node has
infinitely many sons.
In fact we have:
The proof can be found in [27]. Notice that one direction is the statement of
the last exercise. N can be seen as quotient of M1 and M2 under bisimulation
relation, i.e., the minimal representative of the equivalence class [M1 ]∼ .
In the countable case we obtain:
Exercise 14.4. Let M and N be transition systems such that S M and S N are
countable. Show that:
• M is an extension of M,
and N
• If M and N are bisimilar then M are isomorphic.
14 Expressive Power of MSO and Lµ 243
These two logics will be interpreted over transition systems. There are several
ways to define MSO over transition systems, for example by using two types
of variables (first-order and second-order variables) or by introducing a new
predicate sing(X) for singleton sets. We use the following definition:
Definition 14.6. The signature of monadic second order logic (MSO) over
transition systems contains unary predicate symbols from Prop, binary predicate
symbols from Rel, the constant symbol sr and variables from Var. Formulae of
MSO are defined inductively by the following grammar. Let p ∈ Prop, r ∈ Rel
and X, Y ∈ Var:
• sr(X),
• p(X),
• r(X, Y ),
• X ⊆Y,
• ¬ϕ for any formula ϕ,
• ϕ ∨ ψ for any formulae ϕ and ψ,
• ∃X.ϕ(X) for any formula ϕ.
where β[X := T ] denotes the assignment such that β[X := T ](X) = T and
β[X := T ](Y ) = β(Y ) for Y = X.
For a MSO-sentence ϕ we write M |= ϕ if (M, β) |= ϕ is true for an arbitrary
assignment. A MSO-sentence ϕ defines a class of transition systems by
The sentence ϕ states that there is exactly one r-successor of the source. The class
C MSO (ϕ) cannot be bisimulation closed because a bisimulation relation cannot
fix any number of r-successors, i.e., if there is a r-successor in one transition
system then there is one in all bisimilar systems, but there could be arbitrary
many.
In the following we repeat the definition of the µ-calculus (cf. Chapter 10).
Definition 14.8. The signature of the modal µ-calculus L over transition
systems contains only unary predicate symbols from Prop, binary predicate sym-
bols from Rel and variables from Var. Formulae are defined inductively by the
following grammar. Let p ∈ Prop, r ∈ Rel and X ∈ Var:
• X,
• p,
• ¬ϕ for any formula ϕ,
• ϕ ∨ ψ for any formulae ϕ and ψ,
• rϕ for any formula ϕ,
• µX.ϕ(X) for any formula ϕ(X) where X occurs only positively, i.e., under
an even number of negations.
The dual of the modality r is denoted by [r] and defined by [r]ϕ := ¬r¬ϕ.
P
A formula resp. sentence of Lµ is called a Lµ -formula resp. Lµ-sentence.
For a
given transition system M and an assignment β : Var → S M we define
inductively the set ϕM
β in which the Lµ -formula ϕ is true:
XM
β := β(X),
pM M
β := p ,
¬ϕM
β := S
M
− ϕM
β ,
ϕ ∨ ψM M M
β := ϕβ ∪ ψβ ,
M
rϕM
β := s ∈ S sccM M
r (s) ∩ ϕβ = ∅ ,
µX.ϕ(X)Mβ := T ⊆ S M ϕ(X)M β[X:=T ] ⊆ T .
14 Expressive Power of MSO and Lµ 245
C Lµ (ϕ) := M M is a transition system and M, srM |= ϕ .
s ∈ ϕM N
β ⇐⇒ t ∈ ϕβ ∗ ,
ϕ := ¬(rX ∧ r¬X).
L
Let C := C∗ µ (ϕ). For a transition system M and an arbitrary assignment β
we have srM ∈ ϕM β iff either the set β(X) or the complement of β(X) does
not contain any r-successor of srM , i.e., for all M in C we have that either
246 Philipp Rohde
M M
sccM
r sr ∩ β(X) is empty or sccM r sr is a subset of β(X) for every as-
signment β(X). In particular we obtain for the special case β(X) M:= {s} with
s ∈ S M that there is at most one s ∈ S M such that s ∈ sccM r sr for every
M ∈ C. But there are transition systems without this property although they
are bisimilar to M, so C is not bisimulation closed (cf. Remark 14.7).
Proof. For every Lµ -formula ϕ there is a MSO-formula ϕ∗ (X) where the variable
X does not occur in ϕ and such that for every transition system M and every
assignment β with β(X) = {s} for some s ∈ S M :
It is easy to check that ϕ∗ satisfies the property above. We obtain for any as-
signment β:
M, srM |= ϕ ⇐⇒ srM ∈ ϕM β
⇐⇒ M, β[X := {srM }] |= ϕ∗ (X)
⇐⇒ (M, β) |= ∃X. sr(X) ∧ ϕ∗ (X) .
For an arbitrary Lµ -sentence ϕ the formula ϕ̃ := ∃X. sr(X) ∧ ϕ∗ (X) is a MSO-
sentence. Hence it follows C Lµ (ϕ) = C MSO (ϕ̃).
In the sequel we consider structures of the form (T, { m(p) | p ∈ U }), i.e., a
structure with carrier T where each predicate p ∈ U is interpreted as m(p). If a
sentence ϕ ∈ Sent(U) is satisfied in this structure we write as usual
(T, { m(p) | p ∈ U }) |= ϕ.
(Q, ΣP , ΣR , qI , δ, Ω)
an µ-automaton A.
In fact this is the definition of an alternating parity automaton. Observe
that the µ-automaton has two alphabets ΣP and ΣR , the first is for checking
properties of states and the second is for checking the labels of taken transitions.
We will define the acceptance of arbitrary transition systems by the µ-
automaton in terms of games. But we introduce first some abbreviations: for
a given transition system M and a state s ∈ S M let
LM (s) := p ∈ ΣP (M, s) |= p = p ∈ Prop s ∈ pM ∩ ΣP
The initial position is (s0 , q0 ) = srM , qI . If the current position is (si , qi )
then Player 0 is to move. Player 0 chooses a marking mi+1 of SCCM (si ) – i.e.,
a function mi+1 : ΣR × Q → P SCCM (si ) – such that:
• for every r ∈ ΣR and every q ∈ Q the elements of mi+1 (r, q) are r-successors
of si ,
• the structure N := SCCM (si ), { mi+1 (r, q) | r ∈ ΣR , q ∈ Q } is a model of
the first order sentence δ qi , LM (si ) :
N |= δ qi , LM (si ) .
We define for every n ∈ ω the formula diff of first order logic as follows:
diff(x1 , . . . , xn ) := xj .
xi =
1≤i<j≤n
The formula “diff” states that the values of x1 , . . . , xn are pairwise different.
The main tool for our purpose is the following correspondence of µ-automata
and formulae of Lµ and MSO respectively which was proved by Janin and
Walukiewicz.
14 Expressive Power of MSO and Lµ 249
Theorem 14.16.
(1) A class C of transition systems is L -definable iff C = L(A) for a µ-
automaton A = (Q, ΣP , ΣR , qI , δ, Ω) such that Sent(ΣR × Q) contains only
disjunctions of sentences of the form:
∃x1 , . . . , xm . pki (xi ) ∧ ∀y. pki (y) ,
1≤i≤m 1≤i≤m
A := (Q, ΣP , ΣR , q1 , δ, Ω) ,
where the parity function is defined as Ω(q1 ) = 1, Ω(q2 ) = 0 and the transition
function as
∃x1 , x2 . i=1,2 pi (xi ) ∧ ∀y. i=1,2 pi (y) if q = q1 and P = ∅,
δ(q, P ) :=
∀y.⊥ ∨ ∃x. p2 (x) ∧ ∀y.p2 (y) otherwise.
Notice that since we have ∅ = and ∅ = ⊥ the sentence
∃x1 , . . . , xk . pi (xi ) ∧ ∀y. pi (y) (14.1)
1≤i≤k 1≤i≤k
so the move is legal as well. Since m(r, q1 ) is empty Player 1 can only respond
with a position (t, q2 ) where t is a r-successor of s. To this position we can
apply the same strategy again. If the resulting play is infinite then only q2 is
encountered infinitely often. So we have min Inf(Ω(π)) = 0 and therefore Player
0 wins the game. (Claim)
Now we prove that C Lµ (ϕ) = L(A).
(⊆) Let M be a transition system with M, srM |= ϕ, i.e., there is a se-
quence s0 = srM , s1 , . . . , sn with si+1 ∈ sccM
r (si ) for i < n such that (M, sn ) |=
p. We may assume that (M, si ) |= p for i < n.
14 Expressive Power of MSO and Lµ 251
If the current position is (si , q1 ) with i < n then Player 0 plays the mark-
ing mi+1 defined by mi+1 (r, q1 ) = {si+1 } and mi+1 (r, q2 ) = sccM r (si ). Since
LM (si ) = ∅ we have
M
sccr (si ), {mi+1 (r, q1 ), mi+1 (r, q2 )} |= δ(q1 , ∅),
so the move is legal. Then Player 1 must respond with the position (si+1 , q1 ),
since otherwise he would loose the game by the claim above.
So eventually the position (sn , q1 ) is reached. Player 0 then plays the marking
with mn+1 (r, q1 ) = ∅ and mn+1 (r, q2 ) = sccM M
r (sn ). Now we have L (sn ) = {p}
and the move is legal by
M
sccr (sn ), {mn+1 (r, q2 ), mn+1 (r, q1 )} |= δ(q1 , {p}).
If Player 1 can make a move at all he can only respond with the position (t, q2 )
for an r-successor t of sn , so by the claim above he looses the game. This means
that the strategy for Player 0 presented above is a winning strategy in the game
G(M, A) and therefore we obtain M ∈ L(A).
(⊇) Let M be a transition system with M, srM |= ϕ. Let (s, q1 ) be the
current position in the game G. Since we have ΣR = {r} the states si of any
prefix of a play in G form an r-path of M starting in srM . By the assumption we
have (M, s) |= p and therefore LM (s) = ∅. Player 0 has to satisfy δ(q1 , ∅) in the
structure with carrier sccM r (s), so he must play two non-empty subsets m(r, q1 )
and m(r, q2 ) of sccM
r (s) such that the union is the whole set (cf. Example 14.13).
Otherwise he would loose the game. If he can make a move at all then let
t ∈ m(r, q1 ) be an r-successor of s. Player 1 responds with the position (t, q1 ).
By the assumption we have (M, t) |= p as well, so we can apply the same
strategy again. With this strategy either Player 0 cannot make a move or an
infinite game is played, where only q1 is encountered infinitely often. Because
Ω(q1 ) is odd Player 1 wins the game. So we obtain a winning strategy for Player
1 and therefore we have M ∈ L(A).
C is MSO-definable ⇐⇒ C is Lµ -definable.
Before proving the lemma let us show how it implies the theorem:
A∗ = (Q, ΣP , ΣR , qI , δ ∗ , Ω) .
14 Expressive Power of MSO and Lµ 253
is
Claim. Let M be a transition system. Then M is accepted by A∗ iff M
accepted by A.
Before proving the claim let us show how it implies the lemma. By defi-
nition of the function δ ∗ the µ-automaton A∗ has the required form of The-
orem 14.16(2). Hence there is a Lµ -sentence ϕ such that L(A∗ ) = C Lµ (ϕ).
By
Exercise 14.2 the ω-unravelling M is a transition tree for every transition system
M, hence we obtain by the claim
So we have
|= ϕ ⇐⇒ M, srM |= ϕ.
M
In particular we have
sn ∈ pM ⇐⇒ un ∈ pM ,
c (14.5)
c
in particular LM (sn ) = LM (un ) and therefore
c ∗
δ ∗ qn , LM (sn ) = δ qn , LM (un ) . (14.6)
Next we define abbreviations for the two first order structures which occur in
the rules of the games:
N := SCCM (sn ), { m∗n+1 (r, q) | r ∈ ΣR , q ∈ Q }
and
c(un ), { mn+1(r, q) | r ∈ ΣR , q ∈ Q } .
:= SCCM
N
By the fact that m∗n+1 is a legal move of Player 0 in the game G ∗ we have
N |= δ ∗ qn , LM (sn ) . (14.7)
Let ψ ∗ be some satisfied disjunct of δ ∗ qn , LM (sn ) of the form (14.3). We will
show that
|= ψ,
N
where ψ has the original form given by (14.2). By (14.4) the ‘existential part’
of ψ is satisfied by the structure N as well. Because of the ω-indexing there are
infinitely many elements in mn+1 (r, q) corresponding to each single element in
m∗n+1 (r, q). Hence we can always choose pairwise different witnesses in N , i.e.,
the formula diff(x1 , . . . , xm ) is additionally satisfied.
Next we check that N is a model of ∀y.χ(y) as well, in particular the re-
striction ∀y. diff(y, x1 , . . . , xm ) −→ χ(y) and therefore ψ is satisfied by N . To
c
see this let v = un (a, r, t) be an arbitrary element of SCCM (un ). Then t is an
r-successor of sn and by (14.7) we have N |= χ(t), i.e., N is a model of some ap-
propriate predicates p(t) occurring in χ. Since each p is interpreted as m∗n+1 (r, q)
for some q ∈ Q it follows that t ∈ m∗n+1 (r, q) and therefore v ∈ mn+1 (r, q) by
the definition of mn+1 . This means that N is a model of the same predicates
p(v) and therefore a model of χ(v).
In summary this means that taking mn+1 is indeed a legal move of Player 0
in the game G. So we define the value of the strategy f0 for the current position
by mn+1 . From this position Player 1 chooses some rn+1 ∈ ΣR , qn+1 ∈ Q and a
state un+1 ∈ mn+1 (rn+1 , qn+1 ) with un+1 = un (a, rn+1 , t). The pair (un+1 , qn+1 )
becomes the next position in the game G. Now we let sn+1 := t and continue the
game G ∗ by the move (sn+1 , qn+1 ) of Player 1. We arrive at prefixes of plays in
G and G ∗ satisfying our initial assumption.
14 Expressive Power of MSO and Lµ 255
It is clear that if Player 1 gets stuck in the game G ∗ then he cannot make
a move in the game G as well. On the other hand by the inductive definition of
the strategy f0 Player 0 can always make a move in G. Hence he cannot lose in
a finite number of rounds. For an infinite play the result is the sequence
M
sr , qI , m1 , (u1 , q1 ), . . . , mn , (un , qn ), . . .
be the corresponding prefix in the game G where we have: if si+1 ∈ sccM r (si )
holds for r ∈ ΣR then ui+1 = ui (a, r, si+1 ) for some a ∈ ω. The markings mi are
played according to the strategy f0 .
Let mn+1 : ΣR × Q →
P c
SCCM (un ) be the marking suggested by f0 . We
define the marking m∗n+1 : ΣR × Q →
P
SCCM (sn ) by
m∗n+1 (r, q) := t ∈ S M ∃a ∈ ω. un (a, r, t) ∈ mn+1 (r, q) .
Again we have m∗n+1 (r, q) ⊆ sccMr (sn ) by the definition of the ω-unravelling.
Since mn+1 is a legal move of Player 0 in the game G we have
c(un ).
|= δ qn , LM
N (14.8)
c
Let ψ be some satisfied disjunct of δ qn , LM (un ) . We have to check that m∗n+1
is indeed a legal move of Player 0 in the game G ∗ . By (14.6) it suffices to show
that
N |= ψ ∗ , (14.9)
256 Philipp Rohde
where ψ ∗ is the formula defined by (14.3). We may assume that the occurring
predicates are pki = (ri , qi ) with ri ∈ ΣR and qi ∈ Q for every i ∈ {1, . . . , m}.
First we check that
and therefore
N |= ∃x1 , . . . , xm . pki (xi ). (14.10)
1≤i≤m
|= ∃x1 , . . . , xm .
By (14.8) it follows that N |=
pki (xi ), in particular N
1≤i≤m
∃xi .pki (xi ) for every i ∈ {1, . . . , m}. Hence there is some v ∈ SCCM (un ) such
c
that v ∈ mn+1 (ri , qi ). By the definition of successors in M and the fact that
mn+1 (ri , qi ) contains only ri -successors of un in M we have v = un (a, ri , t) for
some a ∈ ω and t ∈ S M . Hence t ∈ m∗n+1 (ri , qi ) by the definition of m∗n+1 . Next
we check
N |= ∀y.χ(y). (14.11)
Let t ∈ sccM r (sn ) for some r ∈ ΣR . We use again the property of the ω-
unravelling that there are infinitely many different r-successors of un in M cor-
responding to each r-successor of sn in M. Hence there exists an a ∈ ω such that
for v = un (a, r, t) we have N |= diff(v, x1 , . . . , xm ). Therefore N
|= χ(v) holds
by (14.8). Since χ is monotone in the predicates we obtain N |= χ(t). To see this,
notice that χ(v) has the form χ(v) = w w pw,w (v) with pw,w ∈ ΣR × Q. So
we have N |= pw,w (v) for some appropriate pairs (w, w ), i.e., v is an element
of mn+1 (r, qw,w ). We obtain t ∈ m∗n+1 (r, qw,w ) by the definition of m∗n+1 and
therefore N |= pw,w (t) for the same predicates, i.e., N |= χ(t) is true. By (14.10)
and (14.11) and the definition of ψ ∗ we have N |= ψ ∗ . This proves (14.9).
Taking m∗n+1 is therefore a legal move of Player 0 in the game G ∗ . We define
the value of the strategy f0∗ for the current position by m∗n+1 and arrive at the
prefix
M
sr , qI , m1 , (u1 , q1 ), . . . , mn , (un , qn ), mn+1
Then (un+1 , qn+1 ) is a legal move of Player 1 in the game G and we continue
it by this move. Again we arrive at prefixes of plays in G and G ∗ satisfying our
initial assumptions.
We have to check that f0∗ is indeed a winning strategy for Player 0 in the
game G ∗ . By the inductive definition of f0∗ Player 0 can always make a move and
hence he cannot lose in a finite number of rounds. As in the first case the played
automaton states in any infinite play of G and in the corresponding infinite play
of G ∗ are the same and the parity functions of both automata are identical. Since
the play in G is according to the winning strategy f0 of Player 0 the value of
min Inf(Ω(π)) is even. It follows that Player 0 wins the game G ∗ as well. Hence f0∗
is indeed a winning strategy for Player 0 and M is accepted by the µ-automaton
A∗ .
Martin Leucker
15.1 Introduction
In 1969, Rabin [148] showed that the monadic second-order theory (MSO-theory)
of infinite binary trees is decidable (see Chapter 12 of this volume or [183]). Ever
since, it has been an interesting goal to extend this result to other classes of
objects.
Muller and Schupp [135] showed that the class of pushdown graphs has a
decidable MSO-theory. This class is obtained by considering the configuration
graphs of pushdown machines. The result was later extended to the class of
regular graphs introduced by Courcelle [42], which are defined as solutions of
graph-grammar equations.
Prefix-recognizable graphs were introduced by Caucal in [28]. They extend
the pushdown graphs of Muller and Schupp and the regular graphs of Courcelle.
Originally, Caucal introduced this class of graphs via transformations on the
complete infinite binary tree. The decidability result of their MSO-theory was
obtained by showing that these transformations are definable by MSO-formulas.
Hereby, the decidability result of the MSO-theory of trees was transferred to
the class of prefix-recognizable graphs. The approach can also be understood
as interpreting prefix-recognizable graphs in the infinite binary tree by means of
MSO-formulas. Barthelmann [6] and Blumensath [12] showed independently that
Caucal’s class of graphs coincides with the class of graphs MSO-interpretable in
the infinite binary tree. In simple words, prefix-recognizable graphs provide a
decidability proof of their MSO-theory via MSO-interpretations in the infinite
binary tree.
The aim of this chapter is to present prefix-recognizable graphs and to show
several of their representations. In contrast to Caucal’s original outline, we start
with graphs that are MSO-interpretable in the binary tree. In this way, we
obtain a natural class of graphs which trivially have a decidable MSO-theory
(see Section 15.3). We then provide several representations of these graphs in
Section 15.3 and Section 15.5. We learn that prefix-recognizable graphs can be
represented by means of prefix-transition systems, whose prefixes form regular
languages, justifying the name of this class. Furthermore, we introduce Caucal’s
transformations on the binary tree and show that they induce the same class of
graphs.
Supported by European Research Training Network “Games”.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 263-283, 2002.
Springer-Verlag Berlin Heidelberg 2002
264 Martin Leucker
15.2 Preliminaries
0 1
00 01 10 11
graphs will be represented in the form G = (V, (Ea )a∈Σ ). For convenience we
allow graphs to be represented also in the form G = (V, E) for V ⊆ N ∗ and
E ⊆ N ∗ ×Σ ×N ∗ . Sometimes, we simply use a ternary relation E ⊆ N ∗ ×Σ ×N ∗
for a graph G, in which case we assume VG (the set of nodes of ) to be G
implicitly defined by VG = {s | ∃a ∈ Σ, ∃t ∈ N ∗ .(s, a, t) ∈ E or (t, a, s) ∈ E}. It
is obvious that our notion of graphs subsumes that of trees.
Another key feature of our notion of graphs is that their nodes can be asso-
ciated with words over some alphabet Σ. Hence, the nodes of graphs constitute
languages. It is a traditional task to deal with finite representations of infinite
languages by means of automata. Taking languages as the domain for our node
sets, we provide the framework of automata theory for defining, characterizing,
and modifying the corresponding graphs.
Let G = (V, (Ea )a∈Σ ) be a graph. An edge (s, t) ∈ Ea is denoted by s−→t,
a
G
or, if G is fixed, by s−→t. A path from s to t in G via a word u = a1 . . . ak ∈ Σ ∗
a
is a sequence
a1 a k
s = p1 −→ · · · −→ pk+1 = t
G G
u
for s, t ∈ V , and appropriate nodes pi ∈ V . We write s=⇒t iff there is path from
G
s to t via u. Again, we may omit the subscript G if it is clear from the context.
L
We write s=⇒t to denote that there is path from s to t via a word u which is
in L. A root of a graph is a node from which all other nodes are reachable, i.e.,
r ∈ V is a root iff for all s ∈ V there is a u ∈ Σ ∗ such that there is path from r
to s via u.
Given a sequence of edges, its sequence of labels is a word over the alphabet
Σ. Given two nodes s and t of a graph G, we define the language L(G, s, t) to
be the union of all words which are obtained on paths from s to t in G in the
way described above. The union of L(G, s, t) for arbitrary nodes s and t of G is
abbreviated by L(G).
a
A letter a can be associated with the set of a-labelled edges {p−→q}, while
a word w = a1 . . . an over Σ can be associated with the w-paths from a node p
a1 an
to a node q, p = p0 −→ · · · −→p n = q.
a
Given sets of words W, U , and V we denote by W (U −→V ) the set of edges
a a
{wu−→wv | w ∈ W, u ∈ U, v ∈ V }. In a similar manner, we define U −→V . A
graph G is called recognizable iff there are a natural number n and a1 , . . . , an ∈
Σ, U1 , V1 , . . . , Un , Vn ∈ REG(Σ ∗ ) such that G = U1 −→V
a1 an
1 ∪ · · · ∪ Un −→Vn .
Let us recall the automata theoretic notations from Chapter 1. We denote
finite automata over words by tuples A = (Q, Σ, ∆, q0 , F ) with a set of states
266 Martin Leucker
= (δ B (x), ϕB
A∼ B
R (x̄), . . . , ϕR
1 n
(x̄))/εB (x, y)
The successor relation is simply defined by ϕS0 (x, y) = S1 xy and every set
of labels Pa can be defined by ϕPa (x) = Pa x. It is now easy to see that I =
(δ(x), ε(x, y), ϕS0 , (ϕPa )a∈Σ ) is an MSO-interpretation of the word A in B where
ε is assumed to express the identity relation of δ A .
Exercise 15.1. A structure A = (A, <A ) is called a dense open order if <A is a
total order on A, if for all x ∈ A there are y, z ∈ A such that y <A x <A z, and
if for all x, y ∈ A such that x <A y there is a z ∈ A such that x <A z <A y.
Show that a dense open order can be interpreted in the infinite binary tree.
Theorem 15.3. If A ≤MSO B and B has a decidable MSO-theory then A has
a decidable MSO-theory.
Proof. We give a sketch of the proof. The details are left as an exercise for the
reader. Let I = (δ(x), ε(x, y), ϕR1 (x̄), . . . , ϕRn (x̄)) be an MSO-interpretation of
A in B. Consider a formula ϕ. Let ϕ be obtained from ϕ in the following way:
Replace every relational symbol R in ϕ by its defining formula ϕR . Furthermore,
relativize every quantifier to δ(x), i.e. substitute ∃xϕ by ∃x(δ(x) ∧ ϕ) and ∀xϕ
by ∀x(δ(x) → ϕ). Now it is easy to see that A |= ϕ iff B |= ϕ .
Since we are interested in interpreting graphs with labelled edges in struc-
tures, we deal with interpretations of the form I = (δ(x), ε(x, y), (ϕRa (x, y))a∈Σ ).
The decidability of the monadic second-order theory of TB was established
by Rabin in [148]. Thus, if we consider MSO-interpretations in the binary tree
we get structures with a decidable MSO-theory.
Corollary 15.4. Every graph which is MSO-interpretable in the infinite binary
tree TB has a decidable monadic second-order theory.
Let us now give a representation of the graphs which are MSO-interpretable
in TBin terms of prefix-transition graphs having regular prefixes.
Definition 15.5. Let Σ be an alphabet. A graph G = (V, (Ea )a∈Σ ) is called
prefix-recognizable iff it is isomorphic to a graph of the form
n
a i
Wi (Ui −→Vi)
i=1
We will show in Section 15.5 that we can choose an arbitrary alphabet with
at least two elements instead of B .
Example 15.6. Let us consider the graph with edge labels a and b given by
a b
B ∗ ((ε−→B ) ∪ B ∗ .(B + −→ε)). It is depicted in Figure 15.2. Note that this graph
is (A∗ , Ra , Rb ) is isomorphic to (ω, succ, >), where succ is the successor relation
on the natural numbers.
b b
b b b
a a a
ε A A2 A3
This example shows that prefix-recognizable graphs may have nodes with
infinite out-degree. The class of prefix-recognizable graphs is a strict extension
of the class of regular graphs, since the latter have only a finite out-degree [42].
Every prefix-recognizable graph can be represented by a finite collection of
a
prefix-recognizable rewrite rules w.u−→v where w, u, and v are regular
expressions. This way of representing prefix-recognizable graphs will be employed
in Chapter 17.
Let us proceed to show that prefix-recognizable graphs coincide with graphs
that are MSO-definable in TB . In our constructions, we need to code tuples of
sets as labelled trees.
Furthermore, we employ Rabin’s tree theorem which gives the relation be-
tween tree automata and MSO-logic: (see also Chapter 12 and [179])
Theorem 15.8. For each ϕ(X̄, x̄) ∈ MSO there is a tree-automaton A such
that L(A) = {TX̄ x̄ | TB |= ϕ(X̄, x̄)}.
If the desired state q is reached, we have to look for a node labelled by [1, ] for
an element of Uq and for a node labelled by [ , 1] for an element of Vq . Hence,
we let Uq := L((Q∪·{qf }, B , ∆Uq , q, {qf })) where
q0 [0, 0] ε
w
q w
u v
[1, 0] wu = x
qf [0, 1] wv = y
qf
Proof. Denote the E-class of x by [x], define p[x] := inf [x] and sx := (p[x] )−1 x.
Let ϕp (x, y) be an MSO-definition of the function x → p[x] . Finally, let s be the
number of states of the tree automaton associated with E. We claim that each
class [x] has an element of length less than |p[x] | + s. Thus, one can define
where ≤ is the lexicographic ordering which is definable since the length of the
words is bounded so that we only need to consider finitely many cases.
To prove the claim, choose x0 , x1 ∈ [x] such that x0 x1 = p[x] . Since
(x0 , x1 ) ∈ E there are regular languages U, V, and W such that x0 = wu, x1 = wv
for u ∈ U , v ∈ V , and w ∈ W with w p[x] . If |wu| ≥ |p[x] | + s then, by a
pumping argument, there exists some u ∈ U such that |p[x] | ≤ |wu | ≤ |p[x] | + s.
Hence, (wu , x1 ) ∈ E is an element of the desired length.
Corollary 15.11.
(1) PRG is closed under prefix-recognizable congruences.
(2) Each graph MSO-interpretable in the binary tree has an injective MSO-
interpretation in TB .
ϕRa (x, y) = ∃z (ϕWi (ε, z) ∧ ϕUi (z, x) ∧ ϕVi (z, y)) .
{i|a=ai }
We will employ these transformations in the next section to obtain further rep-
resentations of prefix-recognizable graphs.
The idea of the first transformation is to collapse paths within a given graph
to a single edge with a new label in the new graph. To be able to deal with
inverse edges of a graph, we introduce the notion of an inverse alphabet.
Definition 15.15. Let Σ be an alphabet. The inverse alphabet of Σ is the
set Σ := {a | a ∈ Σ} which is a disjoint copy of Σ. The extended alphabet of
Σ is the union of Σ and its inverse alphabet and is denoted by Σ̂.
Words over the extended alphabet of Σ may correspond to paths with inverse
edges. For example, the word abb may be understood as the set of pairs of nodes
a b b
(p, q) such that there are p1 and p2 with p1 −→p, p1 −→p2 , and p2 −→q.
We extend the notion of inverse letters to inverse words by defining for every
u = x1 . . . xk ∈ Σ̂ ∗ the inverse u of u by u = xk . . . x1 . Here, every xi is an
element of Σ̂ and for xi = a, a ∈ Σ, xi is identified with a.
Given a word u over Σ̂, we assign to u a normal form u↓ which is obtained
by removing all pairs aa or aa in u. Formally, we could define for Σ a rewrite
system ↓Σ ⊆ Σ̂ ∗ × Σ̂ ∗ by ↓Σ := {(aa, ε), (aa, ε) | a ∈ Σ} and show that it is
terminating and confluent. Hence, we can speak also about the normal form of
u.
Let us now define our first transformation. It is based on the notion of an
extended substitution.
Definition 15.16. Let Σ and Γ be two alphabets. An extended substitution
from Γ to Σ is a homomorphism from Γ into the power set of words over the
extended alphabet Σ̂. More precisely, h is a mapping such that for every b ∈ Γ
h(b) ∈ P (Σ̂ )
∗
P
Sometimes, we silently assume an extended substitution to be extended to a
mapping from Γ̂ ∗ to (Σ̂ ∗ ) by h(b) = h(b) for b ∈ Γ .
Now we are ready to make precise the notion of an inverse substitution of a
graph.
Definition 15.17. Let G = (V, E) be a graph with edge labels from a given
P
alphabet Σ. Furthermore, let Γ be an alphabet, and let h : Γ → (Σ̂ ∗ ) be an
extended substitution. We define the inverse substitution h−1 (G) to be the
graph G = (V, E ) such that
b u
s−→ t iff ∃u ∈ h(b) s=⇒t
G G
for all s, t ∈ V . The inverse substitution is called regular (respectively finite)
iff h is a regular (respectively finite) extended finite substitution.
15 Prefix-Recognizable Graphs and Monadic Logic 273
•
0 1
• 0̄1 • • a •
a
0 1 0 1
• 0̄1 • • 0̄1 • • a • • a •
For the graphs under consideration, we may assume without loss of generality
that their nodes are words over some alphabet Σ. Hence, the nodes of our graphs
constitute languages. A natural operation on languages is restriction. We will
consequently also consider a second transformation on the binary tree called
restriction.
(V ∩ L, E ∩ (L × Σ × L))
L
where LG := {s | r=⇒s}.
G
Proof. Let ϕ be an MSO-formula. Observe that an a-successor of h−1 (G)|LG
corresponds to an h(a)-path in G. Furthermore, an element (a node) x exists
in h−1 (G)|LG iff it is the starting point or end point of some path in G and is
not removed because of the restriction with respect to L. The latter means that
274 Martin Leucker
the element is reached by some L-path from the root z of the graph. Hence, we
define the formula ϕL,h,z inductively:
Sa xy L,h,z = Path h(a) (x, y)
(x ∈ X)L,h,z = x∈X
(¬ϕ)L,h,z = ¬(ϕL,h,z )
(ϕ ∧ ψ)L,h,z = ϕL,h,z ∧ ψ L,h,z
(∃Xϕ)L,h,z = ∃XϕL,h,z
(∃xϕ)L,h,z = ∃x (Path L (z, x) ∧ ∃y(Path M (x, y) ∨ Path M (y, x))
∧ϕL,h,z
where M = h(a) and Path L (x, y) is as in Lemma 15.13. It is easy to see that
Instead of Σ̂, we sometimes consider Σ̂# . All definitions extend to this case
in the obvious way.
Let us collect some properties and interrelations of the transformations men-
tioned above.
u ∈ L(#w−1 L (TN )), and x = u↓# = u↓# . Similarly, v ∈ L(#w−1 L (TN )) and
a
y = v↓# . Thus, s = wx = wu↓# and t = wy = w(v↓# ). Finally, we have s−→t.
G
Let us now show the converse direction, i.e. G ⊆ h−1 (#L (TN )). Consider
s−→t. There are uv ∈ h(a) and w ∈ N ∗ , such that u, v ∈ L(#w−1 L (TN )), s =
a
G
t. Since u ∈ L(#w−1 L (TN )),
a
wu↓# , t = w(v↓# ). We must show that s −→
h−1 (#L ( TN ))
u u u
we have ε =⇒ u↓# = u↓# . So, u↓# =⇒ ε, s = wu↓# =⇒ w. In a
#w−1 L ( TN ) #w−1 L ( TN ) #L ( TN )
v uv h(a)
similar manner, we show that w =⇒ w(v↓# ) = t. Thus, s =⇒ t, s =⇒ t,
#L ( TN ) #L ( TN ) #L ( TN )
a
and s −→ t.
h−1 (#L ( TN ))
G.L := {uw−→vw
a a
| u−→v and w ∈ L}
G
Similarly, we define their left concatenation L.G. For the sake of brevity, we
also write GL and LG instead of G.L and L.G, respectively.
∗
Corollary 15.30. For any substitution h : Σ → N̂# and L ∈ REG(N ∗ ), we
have that h−1 (#L (TN )) equals
W {u↓#−→v↓# | uv ∈ h(a) and u, v ∈ L(#W −1 L (TN ))}.
a
W ∈[L]
q0
q0
p1 a0
q0 p1
pk ∈ F ak
. . . pk
#
a0 a2
0 1 a1
a0
• a1 • • • •
0 1 0 1 a0 a2 a0
a0 a2 a1 a1
• a1 • • • • • • • •
0 1 0 1 0 1 0 1
a0
a2
• • • • • • • •
Note that g(P ) is stable for (f ◦ g)−1 (TB ), and, using Lemma 15.25, G is
isomorphic to:
h−1 (ι−1 (TN )|B∗ )|L = (h−1 (ι−1 (TN ))|B∗ )|L by Lemma 15.25
= (h ◦ ι)−1 (TN )|Vι−1 (T ) ∩B∗ ∩L by Lemma 15.23
= (h ◦ ι)−1 (TN )|B∗ ∩L
N
Theorem 15.34 ([28]). Given an alphabet N with at least two letters, the fol-
lowing properties are equivalent:
(1) G is interpretable in TB .
(2) G ∈ PRGN (Σ)| .
(3) G is isomorphic to (N ∗ H)|L for some recognizable H ⊆ N ∗ × Σ × N ∗ and
∗
L ∈ REG(N ).
n
(4) G is isomorphic to i=1 Wi (Ui −→V
ai
i ) for some n ≥ 0; a1 , . . . , an ∈ Σ;
U1 , V1 , W1 , . . . , Un , Vn , Wn ∈ REG(N ∗ ).
15 Prefix-Recognizable Graphs and Monadic Logic 281
Proof. (2) ⇒ (3): Assume G ∈ PRGN (Σ)| . So G is isomorphic to h−1 (TN )|L
for an appropriate extended substitution h from Σ to N and L ∈ REG(N ∗ ). By
Corollary 15.31, we can write h−1 (TN ) as N ∗ H for
∗
H = {uv ∈ h(a)↓ ∩ N N ∗ , a ∈ Σ}.
where
and
Vi , respectively. Assume that these automata have pairwise disjoint state sets.
Let # be a new symbol and construct the following rewriting system R:
ε U
# =⇒ q0i
ε
pa =⇒ q if q ∈ δiU (p, a)
ε
p =⇒ q if p ∈ FiU and q ∈ FjU
ε
q =⇒ pa if q ∈ δiV (p, a)
V ai
q0i =⇒ #
So G is equal to the restriction to #L of the ε-closure of the prefix transition
graph of R.
15.7 Conclusion
In this chapter we introduced the class of prefix-recognizable graphs, originally
introduced by Caucal (cf. [28]). We have shown that this class of graphs is the
largest class of graphs providing a decidable MSO-theory provable by interpre-
tation in the infinite binary tree.
Several further representations of prefix-recognizable graphs were given in
the literature. Let us sum up (some) known results in the following theorem.
Whenever the formal notions are not clear, we refer the reader to the citations
given.
Theorem 15.36. Let G be a graph. The following statements are equivalent:
(1) G = h−1 (TB )|C for aregular substitution h and a regular language C.
(2) G is isomorphic to ni=1 Wi (Ui −→Va i
i ) for some n ≥ 0; a1 , . . . , an ∈ Σ;
U1 , V1 , W1 , . . . , Un , Vn , Wn ∈ REG(N ∗ ).
3
private communication
15 Prefix-Recognizable Graphs and Monadic Logic 283
(3) G = h−1 (#C (TB )) for a regular substitution h and a regular marking C.
(4) G is MSO-interpretable in the binary tree TB .
(5) G is VR-equational.
(6) G is a prefix-transition graph of Type-2.
(7) G is the configuration graph of a pushdown automaton with ε-transitions.
The equivalence of (1) – (3) was obtained by Caucal in [28]. (4) and (5) are
shown in [6] and [12]. The last two characterizations are due to Stirling [168]. In
this chapter, we have shown the equivalence of (1) – (4) and (7).
Two-player games for push-down graphs and prefix-recognizable are studied
Chapter 17. A different natural class of objects providing a decidable MSO-
theory is presented in Chapter 16.
16 The Monadic Theory of Tree-like Structures
16.1 Introduction
Initiated by the work of Büchi, Läuchli, Rabin, and Shelah in the late 60s,
the investigation of monadic second-order logic (MSO) has received continuous
attention. The attractiveness of MSO is due to the fact that, on the one hand, it
is quite expressive subsuming – besides first-order logic – most modal logics, in
particular the µ-calculus. On the other hand, MSO is simple enough such that
model-checking is still decidable for many structures. Hence, one can obtain
decidability results for several logics by just considering MSO.
For these reasons it is an important task to classify those structures for which
MSO model-checking is decidable. So far, only partial results are known and it
seems doubtful whether a complete characterisation can be obtained.
On the one hand, a useful tool to prove undecidability is the result that MSO
model-checking for the grid ω × ω is undecidable. On the other hand, Rabin’s
famous tree theorem states that, for the complete binary tree, model-checking
is decidable. Since many structures can be interpreted in the binary tree this
provides a wide range of decidability results. Furthermore, we often only need
to consider trees, as many modal logics have the tree-model property.
In this chapter we present a generalisation of Rabin’s Tree Theorem. Given
a structure A we construct its iteration A∗ which is a tree whose vertices are
finite sequences of elements of A. For each relation R of A its iteration has the
relation
R∗ := { (wa0 , . . . , war ) | ā ∈ R, w ∈ A∗ }.
Additionally, we include the successor relation son containing all pairs (w, wa)
for w ∈ A∗ , a ∈ A, and the clone relation cl consisting of all elements of the
form waa. Muchnik’s Theorem states that model-checking is decidable for A if
and only if it is so for A∗ . The first published proof appears in Semenov [162]. It
generalises an unpublished result of Stupp [173] described in Shelah [163] where
the clone relation was left out. Our presentation follows Walukiewicz [200].
For the proof we employ the usual technique of translating formulae into
automata and vice versa. Since, in general, we are operating on trees of infinite
degree, a new type of automaton is needed where the transition function is
defined by MSO-formulae. Furthermore, in order to handle the clone relation,
the transition function has to depend on the current position in the input tree.
In the next section we introduce the kind of automaton we will use to prove
Muchnik’s Theorem but in a more general version than needed, and we prove
that these automata are closed under boolean operations and projection.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 285-301, 2002.
Springer-Verlag Berlin Heidelberg 2002
286 Dietmar Berwanger and Achim Blumensath
In Section 16.3 we will restrict the class of automata to those with MSO-
definable transition function and the translation between automata and MSO-
formulae is presented.
Finally, Section 16.4 contains the proof of Muchnik’s Theorem.
16.2 Automata
To fix our notation, let [n] := {0, . . . , n − 1}. By B + (X) we denote the set
of (infinitary) positive boolean formulae over X, i.e., all formulae constructed
from X with disjunction and conjunction. An interpretation of a formula ϕ ∈
B + (X) is a set I ⊆ X of atoms we consider true. A Σ-labelled A-tree is a
function T : A∗ → Σ which assigns a label T (w) to each vertex w ∈ A∗ .
The main tool used for the investigation of MSO are automata on A-trees.
Since A is not required to be finite we need a model of automaton which can
work with trees of arbitrary degree. In addition the clone relation cl makes it
necessary that the transition function depends on the current position in the
input tree. Thus, we define a very general type of automaton which we will
restrict suitably in the next section.
A = (Q, Σ, A, δ, qI , W )
where the input is a Σ-labelled A-tree, Q is the set of states, qI is the initial
state, W ⊆ Qω is the acceptance condition, and
∗
δ : Q × Σ → B + (Q × A)A
is the transition function which assigns to each state q and input symbol c a
function δ(q, c) : A∗ → B + (Q × A). Frequently we will write δ(q, c, w) instead of
δ(q, c)(w).
Note that the transition function and acceptance condition of these automata
are not finite. To obtain finite automata we will represent the transition function
by an MSO-formula and consider only parity acceptance conditions in the next
section. For simplicity all results in this section are stated and proved for the
general model.
The language accepted by an automaton is defined by way of games. Recall
that a game G = (V0 , V1 , E, W ) consists of sets V0 and V1 of vertices associated
to the respective players, an edge relation E, and the set of winning plays W .
V0 consists of all pairs (q, w) ∈ Q × A∗ and all pairs of the form (ϕ, w) where
ϕ is either atomic or a disjunction, and V1 consists of all pairs where ϕ is a
conjunction.
(b) The initial position is (qI , ε).
(c) Each node (q, w) has the successor δ(q, T (w), w), w . The successors of
nodes (ϕ ◦ ψ, w) are (ϕ, w) and (ψ, w) where ◦ is either ∧ or ∨. Finally, the
successor of nodes (q, a), w with atomic formulae is (q, wa).
(d) Let (ξi , wi )i<ω be a play. Consider the subsequence (ξik , wik )k<ω of po-
sitions where ξik = qk is a state. The play is winning if the sequence q0 q1 . . . is
in W .
The language L(A) recognised by A is the set of all trees T such that player 0
has a winning strategy for the game G(A, T ).
In the remainder of this section we will prove that automata as defined above
are closed under union, complement, and projection. This property is needed in
the next section in order to translate formulae into automata. We start with the
union.
where
Here ϕ denotes the dual of ϕ, i.e., the formula where each ∧ is replaced by ∨
and vice versa.
Proof. Let G(Ā, T ) = (V̄0 , V̄1 , Ē, W̄ ). Note that in G(Ā, T ) the roles of player
0 and 1 are exchanged. V̄0 consists of all former V1 -nodes, and V̄1 contains all
V0 -nodes except for the atomic ones. Since the latter have exactly one successor
it is irrelevant which player they are assigned to. Thus, each choice of player 0
in the old game is made by player 1 in the new one and vice versa. Hence, each
winning strategy σ for player 0 in G(A, T ) is a strategy for player 1 in G(Ā, T )
which ensures that the resulting play induces a sequence in W = Qω \ W̄ . Thus,
σ is winning for 1. The other direction follows by symmetry.
The closure under projections is the hardest part to prove. The projec-
tion Π(L) of a tree-language L is the set of all trees T : A∗ → Σ such that
there is a tree T : A∗ → Σ × {0, 1} in L with T (w) = (T (w), iw ) for some
iw ∈ {0, 1} and all w ∈ A∗ .
The proof is split into several parts. We prove closure under projection for
non-deterministic automata, and show that each alternating automaton can be
transformed into an equivalent non-deterministic one.
Proof. (⊇) Let σ be a winning strategy for player 0 in G(A, T ). G(AΠ , Π(T ))
contains additional vertices of the form (ϕ0 ∨ ϕ1 , w) where ϕi = δ(q, (c, i), w).
By defining
σ(ϕ0 ∨ ϕ1 , w) := ϕi for the i with T (w) = (c, i)
we obtain a strategy for player 0 in the new game. This strategy is winning since,
if one removes the additional vertices from a play according to the extended
strategy, a play according to σ in the original game is obtained which is winning
by assumption.
(⊆) Let σ be a winning strategy for player 0 in G(AΠ , T ). We have to define
a tree T ∈ L(A) with T = Π(T ). Since AΠ is non-deterministic the game has
the following structure: At each position ((q, a), w) with
δ(q, T (w), w) = i k (qik , aik )
player 0 chooses some conjunction k (qik , aik ) out of which player 1 picks a
successor (qik , aik ). Thus, for each word w ∈ A∗ there is at most one state q such
that a play according to σ reaches the position (q, w). Let σ(ϕ0 ∨ϕ1 , w) = (ϕi , w)
where ϕ0 ∨ ϕ1 = δ(q, T (w), w). We define T by T (w) := (T (w), i).
It remains to show how to translate alternating automata to non-determin-
istic ones. To do so we need some notation to modify transition relations.
Definition 16.11. Let ϕ ∈ B + (Q × A).
(a) The collection of ϕ is defined as follows. Let i k (qik , aik ) be the
disjunctive normal-form of ϕ.
collect(ϕ) :=
Qi (a), a ∈ B + ( (Q) × A)P
i a∈A
An :=
P
(Q × Q), Σ, A, δn , {(qI , qI )}, Wn
P
A sequence (qi )i<ω ∈ Qω is called a trace of (Si )i<ω ∈ (Q × Q)ω if (qi , qi+1 ) ∈
P
Si for all i < ω. Wn consists of all sequences (Si )i<ω ∈ (Q × Q)ω such that
every trace of (Si )i<ω is in W .
P P
a parity acceptance condition, there is some parity automaton on infinite words
B = (P, (Q × Q), δ, p0 , Ω) which recognises Wn ⊆ (Q × Q)ω . Let Ap be the
product automaton of An and B, that is,
P
Ap = P × (Q × Q), Σ, A, δp , (p0 , qnI ), Σp
where
where (Φ, w) = σ(S, w) and p = δ(p, S). That way every play
(p0 , S0 ), w0 Φ0 , w0 (p1 , S1 ), w1 Φ1 , w1 . . .
in Ǧ(An , T ) according to σ. Further, (pi )i<ω is the run of B on (Si )i<ω . Since
the second play is winning, the first one is so as well, by definition of B. Hence,
σ is a winning condition. The other direction is proved analogously.
For simplicity we will use a variant of monadic second-order logic where all
first-order variables are eliminated. That is, formulae are constructed from atoms
of the form X ⊆ Y and RX0 . . . Xr by boolean operations and set quantification.
Using slightly non-standard semantics we say that RX̄ holds if ā ∈ R for some
elements ai ∈ Xi . Note that we do not require the Xi to be singletons. Obviously,
each MSO-formula can be brought into this form.
A |= ϕ(∅, S̄) ,
ϕA (wa) := { (q, b) | b ∈ Sq } S0 , . . . , Sn−1 ⊆ A such that
A |= ϕ({a}, S̄) .
In order to translate formulae into automata, the latter must be closed under
all operations available in the respective logic.
Proposition 16.20. MSO-automata are closed under boolean operations and
projection.
Proof. By Theorem 16.15 it is sufficient to show closure under or, and, dual,
shift, and collection. To do so we will frequently need to convert between inter-
pretations I ⊆ Q × A of boolean formulae ϕA (w) ∈ B + (Q × A) and sets Q̄
such that A |= ϕ(C, Q̄). Given I ⊆ Q × A define
Qi (I) := { a ∈ A | (qi , a) ∈ I }
I |= ϕA (w)
iff Q × A \ I |= ϕA (w)
iff J |= ϕA (w) implies J ∩ I = ∅
iff A |= ϕ(C, P̄ ) implies Pi ∩ Qi (I) = ∅ for some i
iff A |= ∀P̄ ϕ(C, P̄ ) → i<n Pi ∩ Qi = ∅
(shift) For a shift we simply need to renumber the states. If the pair (qi , qk )
is encoded as number ni + k we obtain
P
if L(A) consists of those structures whose labelling encode sets Ū such that
ϕ(Ū ) holds. The encoding of Ū is the ([m])-labelled tree T such that
T (w) = { i ∈ [m] | w ∈ Xi }
P
Proof. (⇒) By induction on ϕ(X̄) we construct an equivalent MSO-automaton
A := (Q, ([m]), δ, q0 , Ω). Since or corresponds to union, negation to comple-
ment, and existential quantifiers to projection, and MSO-automata are closed
under all of those operations we only need to construct automata for atomic
formulae.
(Xi ⊆ Xj ) We have to check for every element w of the input tree T that
i∈/ T (w) or j ∈ T (w). Thus, we set Q := {q0 } with Ω(q0 ) := 0 and define the
transition function such that
a∈A (q0 , a) if i ∈
/ c or j ∈ c,
δA (q0 , c, w) =
false otherwise.
Note that this is the only place where the transition function actually depends
on the current vertex.
(⇐) Let A = (Q, Σ, δ, 0, Ω) be an MSO-automaton and fix an input struc-
ture A∗ . W.l.o.g. assume that A is non-deterministic. A∗ is accepted by A if
there is an accepting run 1 : A∗ → Q of A on A∗ . This can be expressed by an
MSO-formula ϕ(X̄) in the following way: we quantify existentially over tuples Q̄
encoding 1 (i.e., Qi = 1−1 (i)), and then check that at each position w ∈ A∗
a valid transition is used and that each path in 1 is accepting.
The next lemma provides a normalform for MSO-automata over the empty sig-
nature.
Lemma 16.23. Every monotone MSO-formula ϕ(X̄) over the empty signature
is equivalent to a disjunction of FO-formulae of the form
∃ȳ diff(ȳ) ∧ ϑi (yi ) ∧ ∀z diff(ȳ, z) → ϑ (z)
i<n i<m
where the ϑi and ϑi are the positive part of some type.
16 The Monadic Theory of Tree-like Structures 297
each of which defines one of those n-equivalence classes where ϕ holds. If ϕ(X̄) is
monotone we can drop all negative atoms of the τi , τi .
Analogously, one can show the claim also for MSO-formulae
Q0 Y0 · · · Qn−1 Yn−1 ϕ(X̄, Ȳ )
with ϕ ∈ FO, since the effect of set quantifiers amounts to splitting each type
into two.
Similarly the decidability of SωS can be obtained directly without the need to
interpret the infinitely branching tree into the binary one.
Example 16.27. Let A := (ω, ≤). The iteration A∗ has universe ω∗ and relations
≤∗ = { (wa, wb) | a ≤ b, w ∈ ω ∗ },
son = { (w, wa) | a ∈ ω, w ∈ ω ∗ },
cl = { waa | a ∈ ω, w ∈ ω ∗ }.
The proof of Muchnik’s Theorem is split into several steps. First, let A =
(Q, Σ, δ, qI , Ω) be the MSO-automaton equivalent to ϕ. W.l.o.g. assume that
Ω(i) = i for all i ∈ Q = [n]. Note that the input alphabet Σ = {∅} of A is
unary since ϕ is a sentence. We construct a formula ϕ̂ stating that player 0 has
a winning strategy in the game Ǧ(A, A). Hence,
A |= ϕ̂ iff A∗ ∈ L(A) iff A∗ |= ϕ.
A µ-calculus formula defining the winning set is given in Example 10.8 of Chap-
ter 10. Translated into monadic fixed point logic it looks like
LFPZn ,x · · · GFPZ1 ,x ηi (x, Z̄)
i≤n
with
The game structure. In order to evaluate the above formula we need to embed
Ǧ(A, A) in the structure A. First, we reduce the second component of a position
(X, w) from w ∈ A∗ to a single symbol a ∈ A. Let G (A, A) be the game obtained
from Ǧ(A, A∗ ) by identifying all nodes of the form (q, wa) and (q, w a), i.e.:
(a) Let V0 := Q × A. The vertices of player 0 are V0 ∪ {(q0 , ε)}, those of
P
player 1 are V1 := (Q × A).
(b) The initial position is (q0 , ε).
(c) Let δ(q, ∅)A (a) = i Φi for a ∈ A∪{ε}. The node (q, a) ∈ V0 has the
successors Φi for all i. Nodes Φ ∈ V1 have their members (q, a) ∈ Φ as successors.
(d) A play (q0 , a0 ), Φ0 , (q1 , a1 ), Φ1 , . . . is winning if the sequence (qi )i<ω sat-
isfies the parity condition Ω.
Lemma 16.28. Player 0 has a winning strategy from the vertex (q, wa) in the
game Ǧ(A, A∗ ) if and only if he has one from the vertex (q, a) in the game
G (A, A).
Proof. The unravelings of Ǧ(A, A∗ ) and G (A, A) from the respective vertices are
isomorphic.
Note that these relations only contain elements of V0 . Let G(A, A)|V0 denote the
restriction of G(A, A) to V0 .
Finally, we can embed G(A, A)|V0 in A via an interpretation.
Definition 16.29. Let A = (A, R0 , . . . , Rr ) and B be structures. An interpre-
tation of A in B is a sequence
I := k, (ϑR
ı̄ )R,ı̄
where, given R of arity r, the indices ı̄ range over [k]r , such that
I(A) |= ϕ iff A |= ϕI
for every structure A.
To construct ϕI one simply replaces each relation in ϕ by its definition.
Lemma 16.31. There is an interpretation I with G(A, A)|V0 = I(A) for all
structures A.
Proof. Let I be defined by
eq
ϑik 2 (X, Y ) := X = Y,
Si true if i = k,
ϑk (X) :=
false otherwise,
RX̄ if k0 = · · · = kr ,
ϑR i
(X̄) :=
k̄
false otherwise.
In order to speak about all of G(A, A) in its restriction to V0 we treat elements
P
Φ ∈ V1 = (V0 ) as sets Φ ⊆ V0 . All we have to do is to define the edge relation.
We split E into three parts
Lemma 16.32. There are formulae ε0 (x, Y ), ε1 (X, y), and ε2 (Y ) defining the
edge relations E0 , E1 and E2 respectively.
Proof. Since Φ, (q, a) ∈ E1 iff (q, a) ∈ Φ we set
ε1 (Y, x) := Y x.
The definition of ε0 is more involved. Let δq (C, Q̄) := δ(q, ∅)A . We have
((q, a), Φ) ∈ E0 iff A |= δq ({a}, Q̄)
where Qi := { b | (i, b) ∈ Φ }. In order to evaluate δq we need to define A in-
side G(A, A). Since the latter consists of |Q| copies of A with universes (Sq )q∈Q ,
we pick one such copy and relativise δq to it. For simplicity we choose Sq corre-
sponding to the first component of (q, a).
((q, a), Φ) ∈ E0 iff G(A, A)|V0 |= δqSq {(q, a)}, Q̄
where Qi := { (q, b) | (i, b) ∈ Φ }. This condition can be written as
G(A, A)|V0 |= ∃C∃Q̄ δqSq (C, Q̄) ∧ C = {(q, a)}
∧ Qi = { (q, b) | (i, b) ∈ Φ } .
i∈Q
Thus, we define
ε0 (x, Y ) := Sq x ∧ εq0 (x, Y )
q∈Q
where
εq0 (x, Y ) := ∃C∃Q̄ δqSq (C, Q̄) ∧ C = {x} ∧ Qi = { (q, b) | (i, b) ∈ Y } .
i∈Q
Thierry Cachat
17.1 Introduction
Parity games (where the winner is determined by the parity of the maximal
priority appearing infinitely often) were presented in Chapter 2 and algorithms
solving parity games for the case of finite graphs in Chapter 7. In this paper we
study parity games on a simple class of infinite graphs: the pushdown (transition)
graphs. In [106], Kupferman and Vardi have given a very powerful method for the
µ-calculus model checking of these graphs: the formalism of two-way alternating
tree automata. This is a generalization of the (one-way) tree automata presented
in Chapters 8 and 9.
The transition graph of the pushdown automaton defines the arena: the graph
of the play and the partition of the vertex set needed to specify the parity winning
condition. We know from Chapter 6 that such games are determined and that
each of both players has a memoryless winning strategy on his winning region.
The aim of this paper is to show how to compute effectively the winning region
of Player 0 and a memoryless winning strategy. The idea of [106] is to simulate
the pushdown system in the full W -tree, where W is a finite set of directions,
and to use the expressive power of alternating two-way tree automata to answer
these questions. Finally it is necessary to translate the 2-way tree automaton into
an equivalent nondeterministic one-way tree automaton, with the construction
from [190].
In the next section we define two-way alternating automata and the effective
construction from [190] of equivalent one-way nondeterministic automata. In
Section 17.3 we apply these results to solve parity games over pushdown graphs
and to compute winning strategies. Section 17.4 presents an example. Some
extensions and modifications are discussed in Section 17.5.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 303-317, 2002.
Springer-Verlag Berlin Heidelberg 2002
304 Thierry Cachat
x q’
q
xA xA
xAB q’’
Fig. 17.1. Example of a transition δ(q, A) = (↑, q ) ∧ (B, q ), with the convention that
the label is equal to the last letter of the node
17 Two-Way Tree Automata Solving Pushdown Games 305
The idea of a transition δ(q, l1 ) = (↑, q )∧(d, q ) is the following: if the automaton
A is in state q on the node x of the labeled tree T, l and reads the input
l1 = l(x), it will send a “copy” of A in state q to the parent node of x and
another copy in state q to xd. See Figure 17.1. After that the two copies are
running independently. They may come again to the same node with two different
states.
More precisely a run of an alternating two-way automaton A over a labeled
tree W ∗ , l is another labeled tree Tr , r in which every node is labeled by an
element of W ∗ × Q. The latter tree is like the unfolding of the run, its structure
is quite different from W ∗ . A node in Tr , labeled by (x, q), describes a “copy”
of the automaton that is in state q and is situated at the node x of W ∗ . Note
that many nodes of Tr can correspond to the same node of W ∗ , because the
automaton can come back to a previously visited node. The label of a node
and its successors have to satisfy the transition function. Formally, a run Tr , r
is a Σr -labeled Γ -tree, for some (almost arbitrary) set Γ of directions, where
Σr := W ∗ × Q and Tr , r satisfies the following conditions:
Remember that x.d can be x. or x.↑, and the latter is defined only if x = . So
the run cannot go up from the root of the input tree. Note that it cannot use a
transition δ(q, l(x)) = false since the formula false cannot be satisfied.
A run Tr , r is accepting if all its infinite paths satisfy the acceptance con-
dition Acc (the finite paths of a run end with a transition θ = true , which is
viewed as successful termination). We consider here only parity acceptance con-
ditions (see previous chapters): Acc is given by a priority function Ω : Q −→ [m].
An infinite path β ∈ Tr ω satisfies the acceptance condition iff the smallest prior-
ity appearing infinitely often in this path is even: min Inf(Ω(r(β))) is even. Such
a path in the run consists of following only one “copy” of the automaton. An
automaton accepts a labeled tree if and only if there exists a run that accepts it.
The tree language accepted by an automaton A is denoted L(A). Two automata
are equivalent if they accept the same tree language.
The automaton A = (Q, Σ, δ, qI , Ω) and the input tree T, l are now fixed
for the rest of the Section 17.2. In the next subsections we will study how the
automaton A can accept the given tree. The strategy for A will give information
about the transitions used by A (because A is not deterministic). Then the
annotation will store the priorities seen during the detours of A. With all these
auxiliary tools, it is possible to construct a one-way tree automaton that checks
whether A accepts a tree.
306 Thierry Cachat
17.2.2 Strategy
In the same way as in Chapters 6, 4 and 8 of this book, A itself (as an alternating
automaton) is equivalent to a two-player parity game. The initial configuration
of this game is (, qI ) (= r()). From a configuration (x, q), x ∈ T, q ∈ Q, Player
0 chooses a set Y ⊆ ext(W ) × Q that satisfies δ(q, l(x)), then Player 1 chooses
d, q ∈ Y , the new configuration is (x.d, q ) and so on. If x.d is not defined or
δ(q, l(x)) = false then Player 1 wins immediately. If Y is empty (δ(q, l(x)) =
true ) then Player 0 wins immediately. If the play is infinite, then Player 0 wins
iff the parity condition is satisfied. So Player 0 is trying to show that A accepts
the input tree, and Player 1 is trying to challenge that.
Player 0 has a memoryless winning strategy iff A has an accepting run (see
Chapter 6). In other words, if A has an accepting run, then it has an accepting
run using a memoryless winning strategy: choosing always the same “transitions”
from the same node and state. We decompose the run of A using this strategy.
Definition 17.1. A strategy for A and a given tree is a mapping
τ : W ∗ −→ P (Q × ext(W ) × Q).
Intuitively (q, d, q ) ∈ τ (x) means that if A is in state q on the node x, it has to
send a copy in state q to node xd. It is memoryless because it depends only on
x. Note that the strategy does not read the labels, but it is defined for a fixed
tree T, l . See an example on Figure 17.2.
q1 q3 {(q1,A,q2),(q3,B,q4)}
^
{(q2,|,q3),
q2 q4
(q2,B,q5)}
q5
Considering St := P
(Q × ext(W ) × Q) as an alphabet, a (St × Σ)-labeled
tree defines a memoryless strategy on the corresponding Σ-labeled tree. We
will construct a one-way automaton B that checks that this strategy is correct
according to the previous requirements. For (q, d, q ) ∈ τ (x), if d ∈ W it has just
to check in the direction d downwards that the strategy is well defined for q ,
but if d =↑, he must have remembered that the strategy was defined for q in the
P P
parent-node. The states of B are pairs Q1 , Q2 ∈ (Q) × (Q), where q ∈ Q1
means that B has to check (down) that the strategy can be followed for q , and
q ∈ Q2 means that q is already allowed at the parent node.
P P
B := ( (Q) × (Q), St × Σ, δB , {qI }, ∅ , true ) where (17.4)
δB (Q1 , Q2 , τ1 , l1 ) :=
IF ∀q ∈ Q1 , { (d2 , q2 ) | (q, d2 , q2 ) ∈ τ1 } satisfies δ(q, l1 ), and (17.5)
∀(q , , q) ∈ τ1 , { (d2 , q2 ) | (q, d2 , q2 ) ∈ τ1 } satisfies δ(q, l1 ), and (17.6)
∀(q, ↑, q ) ∈ τ1 , q ∈ Q2 (17.7)
THEN d, { q | ∃ (q, d, q ) ∈ τ1 }, Q2 (17.8)
d∈W
with Q2 := { q | ∃ d1 , q1 , (q , d1 , q1 ) ∈ τ1 or ∅ satisfies δ(q , l1 ) }, (17.9)
ELSE false . (17.10)
The acceptance condition is easy to enunciate: it just requires that each path of
B is infinite (i.e., the transition is possible at each node). Note that although we
have used the formalism of alternating automata, B is a deterministic one-way
automaton: B sends exactly one copy to each son of the current node. It has 4|Q|
states.
In condition (17.5) there is no requirement on the q ∈ Q1 , that’s why the
condition (17.1) above is stronger. This is not a problem for the following, as we
are searching some winning strategy (one could define the minimal valid strategy
as in [190]). If A follows the strategy, its run is “deterministic” on the input tree
labeled by St × Σ.
A path β in a strategy (tree) τ is a sequence (u0 , q0 ), (u1 , q1 ), (u2 , q2 ), · · ·
of pairs from W ∗ × Q such that (u0 , q0 ) = (, qI ) and for all i > 0, there is
some ci ∈ ext(W ) such that (qi , ci , qi+1 ) ∈ τ (ui ) and ui+1 = ui ci . Thus, β just
follows (nondeterministically) the “transitions” of τ . The parity condition for β
is defined exactly the same way as for a path of (a run of) A. We say that τ is
accepting if all infinite paths in τ are accepting.
Proposition 17.2. A two-way alternating parity automaton accepts an input
tree iff it has an accepting strategy tree over the input tree.
With the help of a so called annotation, we will check in the following subsections
whether a strategy is accepting.
308 Thierry Cachat
17.2.3 Annotation
The previous automaton B just checks that the strategy can be followed (ad
infinitum) but forgets the priorities of A. To check the acceptance condition, it
is necessary to follow each path of A up and down, and remember the priorities
appearing. Such a path can be decomposed into a downwards path and several
finite detours from the path, that come back to their origin (in a loop). Because
each node has a unique parent and A starts at the root, we consider only down-
wards detour (each move ↑ is in a detour). That is to say, if a node is visited
more than once by a run β, we know that the first time it was visited, the run
came from above. To keep track of these finite detours, we use the following
annotation.
Definition 17.3. An annotation for A and a given tree is a mapping
q1 q3 {(q1,2,q3)}
q2 q4
Fig. 17.3. Part of a run and the corresponding annotation, assuming that Ω(q2 ) =
2, Ω(q3 ) = 3
Intuitively (q, f, q ) ∈ η(x) means that from node x and state q there is a
detour that comes back to x with state q and the smallest priority seen along
this detour is f . Figure 17.3 presents an example. By definition, the following
conditions are required for the annotation η of a given strategy τ :
∀ q, q ∈ Q, x ∈ W ∗ , d ∈ W, f, f ∈ [m] :
(q, , q ) ∈ τ (x) ⇒ (q, Ω(q ), q ) ∈ η(x), (17.12)
(q1 , f, q2 ) ∈ η(x), (q2 , f , q3 ) ∈ η(x) ⇒ (q1 , min(f, f ), q3 ) ∈ η(x), (17.13)
(q, d, q1 ) ∈ τ (x), (q1 , ↑, q ) ∈ τ (xd) ⇒ (q, min(Ω(q1 ), Ω(q )), q ) ∈ η(x),
(17.14)
(q, d, q1 ) ∈ τ (x), (q1 , f, q2 ) ∈ η(xd), (q2 , ↑, q ) ∈ τ (xd)
⇒ (q, min(Ω(q1 ), f, Ω(q )), q ) ∈ η(x). (17.15)
P
Considering An := (Q × [m] × Q) as an alphabet, the aim is to construct a
one-way automaton C on (An×St)-labeled trees that checks that the annotation
satisfies these requirements. The conditions 17.12 and 17.13 above can be checked
in each node (independently) without memory. For the last two, the automaton
17 Two-Way Tree Automata Solving Pushdown Games 309
has to remember the whole η(x) from the parent node x, and the part of τ (x)
leading to the current node.
C := (An × P (Q × Q), An × St, δ , ∅, ∅ , true ),
C
where
δC (η0 , α , η1 , τ1 ) :=
IF conditions 17.12 and 17.13 hold for η1 and τ1 AND
∀(q, q1 ) ∈ α, (q1 , ↑, q ) ∈ τ1 ⇒ (q, min(Ω(q1 ), Ω(q )), q ) ∈ η0
∀(q, q1 ) ∈ α, (q1 , f, q2 ) ∈ η1 , (q2 , ↑, q ) ∈ τ1
⇒ (q, min(Ω(q1 ), f, Ω(q )), q ) ∈ η0
THEN d, η1 , { (q, q1 ) | ∃ (q, d, q1 ) ∈ τ1 }
d∈W
ELSE false .
2 2
Similarly to B, C is a deterministic one-way automaton with 2|Q| m · 2|Q| =
2
2|Q| (m+1) states, and the acceptance condition is very simple: each path has to
be infinite. Note that if a part of the tree is not visited by the original automaton
A, the strategy and annotation can be empty on this part. The automaton C does
not check that the annotation is minimal, but this is not a problem. With the
help of the annotation one can determine if a path of A respects the acceptance
condition or not, as showed in the next subsection.
At each step D either goes down following the strategy, or simulates a detour
with an -move and the corresponding priority. The second component ([m])
of the states of D just remembers the last priority seen. We can transform D
into a nondeterministic one-way automaton D without -moves with the same
state space. Note that D can possibly stay forever in the same node by using
-transitions, either in an accepting run or not. This possibility can be checked
by D just by reading the current annotation, with a transition true or false .
We will use D and D to find the invalid paths of the run of A, just by
changing the acceptance condition: Ω0 (q, i ) := i + 1.
310 Thierry Cachat
These definitions are essentially the same for finite and infinite arena. We con-
sider now pushdown graphs: (V, E) is the (possibly infinite) transition graph of
a pushdown system, which is an unlabeled pushdown automaton.
Definition 17.7. A pushdown system (PDS) is a tuple Z := (P, W, ∆)
where:
A stack content is a word from W ∗ . Unlike standard notation we write the top
of the stack at the right of the word (we are considering suffix rewriting as in
Chapter 15). A configuration is a stack content and a control state: (w, p),
shortly wp, where w ∈ W ∗ , p ∈ P . The transition graph of Z is (V, E) where
V = W ∗ P is the whole set of configurations and ∀u, w ∈ W ∗ , ∀a ∈ W, ∀p, p ∈ P
(uap)E(uwp ) ⇔ (p, a, p , w) ∈ ∆.
This defines a vertex labeled graph: each vertex is labeled by his name, the
edges have no label. We use the name pushdown system, like in [61] because
the transitions are not labeled: we are not interested in the language recognized
by the pushdown automaton but in the underlying transition graph. To obtain
a parity game, it remains to define the sets V0 and V1 , associating the vertices
to the two players, and the priorities of the configurations. One fixes a disjoint
union P = P0 ∪· P1 , then V0 = W ∗ P0 and V1 = W ∗ P1 . The mapping ΩG is first
defined on P , then ΩG (wp) = ΩG (p), ∀w ∈ W ∗ and p ∈ P . So the player and the
priority only depend on the control states of Z, like in [196] and [198]. These
restrictions will be discussed later in Section 17.5.1.
The pushdown game is completely defined if we also fix an initial configu-
ration vI ∈ V : vI = wI pI .
W -tree, guess nondeterministically the best moves of Player 0 and follow each
possible move of Player 1 using alternation.
As an example, the transition (p, a, p , bc) ∈ ∆ from a configuration uap of
the pushdown system can be simulated by a two-way automaton over the full
W -tree from the node ua by the following sequence of moves: ↑, b, c because
ua↑bc = ubc. We have chosen suffix rewriting rather than prefix to conform with
the notation of the tree. The control states of Z are represented in the states of
A.
For our particular application, we simplify the definition of two-way automata
a little. The full W -tree will not be labeled by an input alphabet Σ, and the
automaton will “read” the last letter of the node, almost the same way as a
pushdown automaton (as remarked in [106], another solution is to check that
each label is equal to the last letter of its node).
To simulate with many steps a transition of Z, A has to remember in its
states the letters it has to write (see Figure 17.4). Let
<q,EB>
x
<q,B>
<p,ε >
xA xE
xEB
<q,ε >
Fig. 17.4. Transition (p, A, q, EB) of the PDS (long arrow) simulated by the two-way
automaton
executes all the possible moves of Player 1, to ensure that Player 0 can win after
each of these moves. But if p ∈ P0 , A chooses nondeterministically a move of
Player 0 and tries to make Player 0 win.
The “winning” condition of A is almost the same as the one of G: Ω(p, x ) =
ΩG (p). The initial state of A causes it to go “deterministically” to the initial
configuration of the game.
Theorem 17.8. Player 0 has a winning strategy in (G, vI ) iff A accepts the full
infinite tree W ∗ .
These strategies are not necessarily memoryless as presented, but the result of
Chapter 6 holds for both formalisms: there is a memoryless winning strategy if
there is a winning strategy.
17.4 Example
⊥ap1 /0 ⊥ap0 /1
⊥aap1 /0 ⊥aap0 /1
··· ···
As a shortcut we consider
Now we can see that D can not find an accepting path, i.e., a winning path for
Player 1. Player 0 win the game from qI , provided he never moves to the vertex
⊥p3 . Unlike finite graphs, we can have here arbitrary long paths with (minimal)
priority 1, but no infinite path winning for Player 1.
Exercise 17.1. Complete the solution of this example, and compute the strategy
according to section 17.3.3.
We have assumed that the priority of a configuration depends only on the con-
trol state. Another possibility is to define regular set of states for each priority,
or equivalently a finite automaton with output (over the alphabet W ∪· P ) that
accepts each configuration and outputs its priority. That wouldn’t be more gen-
eral: this automaton can be simulated by the states of the one-way automaton
E (or by A with new labels on the tree). Otherwise it can be simulated by Z by
extending the stack alphabet. The same ideas apply for the definition of V0 and
V1 in V .
A usual convention for an arena (V0 , V1 , E) is that E ⊆ V0 × V1 ∪ V1 × V0 ,
i.e., Player 0 and 1 alternate. This convention may clarify the situation but is
not essential for us. If a pushdown system Z does not satisfy it, we can add
“dummy” states to obtain a new pushdown game Z which is equivalent to Z
and satisfies the condition that in the new states there is only one possible move
(choice).
The usual convention is also that a player who cannot move has lost. This is
convenient with our formalism if we consider (see equations 17.17 and 17.18) that
an empty disjunction is false and an empty conjunction is true (analogously it
agrees with the definitions of and in µ-calculus). With pushdown games we
can simulate another convention. We know in which configuration no transition
is possible: if the stack is empty, or if there are no q , u with (q, a, q , u) ∈ ∆.
316 Thierry Cachat
We can add new transitions to a particular state for the second case, and for
the first case we can use a new symbol as the “bottom” of the stack, that can
neither be put nor removed, and new transitions for this symbol.
17.5.2 Extensions
One can easily extend the results presented in this paper to any suffix (resp.
prefix) rewrite system, either by simulating it with a pushdown automaton (up
to bisimilarity) or by adapting our construction to allow A to go up along a
fixed word (stored in its memory). In contrast one could restrict the pushdown
system so that a transition consists just of pushing or popping one letter, which
is equivalent to the general model.
In [106] other results are obtained with the help of two-way automata: the
model checking procedure is extended to any µ-calculus formula (Theorem 2)
over any context-free or even prefix recognizable graph (Theorem 5). In the
present paper we have just considered the problem of solving parity games. On
the other hand, each µ-calculus formula on a pushdown system is equivalent to
a parity game. To simulate the prefix recognizable rewrite rules (see Chapter
15), the two-way automaton simulates the finite automata that recognize the
different parts of the word (the prefix, the old suffix and the new suffix) using
alternation and guessing the correct rule.
17.6 Conclusion
After some technical work to make the two-way automata “usable”, it was pos-
sible to compute winning regions and winning strategies. This formalism is very
powerful and hopefully comprehensible.
Its expressive power is the same as the µ-calculus on trees and on the tran-
sition systems that can be simulated on trees: pushdown systems and prefix
recognizable graphs.
Chapter 15 of this book deals with another result about model checking: it
is shown that Monadic Second Order logic (MSO) is decidable on prefix rec-
ognizable graphs. It is well known (see Chapter 14) that MSO is at least as
expressive as µ-calculus, so implicitly the model-checking problem for µ-calculus
on prefix-recognizable graphs was already solved by Caucal in [28]. It is natural
to define parity games on prefix-recognizable graphs the same way as we have
done: a configuration (node of the game graph) is a word, for clarity we suppose
that the priority and the player (V0 and V1 ) are given by the first letter of the
word. In fact we can define in MSO the winning region of Player 0 (resp. Player
1).
But if we compare both approaches in more detail, important differences
show up: the MSO formula describes the whole winning region: the decision
procedure gives a finite automaton that recognizes the whole set of winning
vertices of Player 0 (resp. 1). On the contrary, the construction presented in the
present chapter just checks one at a time if a given “initial” configuration is in
17 Two-Way Tree Automata Solving Pushdown Games 317
the winning region. On the other hand, it is proved in [106] that his technique
generates a winning strategy for this initialized game, represented by a finite
automaton.
A similar result could be obtained by the methods introduced in Chapter
15, if a strategy could be defined in MSO. Unfortunately, this is not possible
over the given arena. Indeed, a strategy is a binary relation, or a function from
the vertices to the vertices, and it is not allowed in MSO to quantify about
(non monadic) relations. Note that a strategy provides more information than
a winning region does. It is possible to stay forever in the winning region and
not win (never reach the “goal”). One cannot quantify about paths: they are not
uniquely defined by their set of nodes. Finally, if the prefix-recognizable graph
is a directed tree, and the game played from the root (top-down), the situation
is much simpler: the strategy is a subtree with some good conditions, and is
MSO-definable. (This gives an answer to a question of [180].) But in general
the unraveling tree of a prefix-recognizable graph from a given vertex is not a
prefix-recognizable graph (it is an algebraic tree, but this is outside the scope of
this book).
18 Introduction to Guarded Logics
Thoralf Räsch
18.1 Introduction
Guarded logics are a family of logical formalisms that generalize certain desir-
able properties of modal logics from transition systems to the setting of arbitrary
relational structures. Modal logics are widely used in a number of areas in com-
puter science, particularly for the specification and verification of hardware and
software systems, for knowledge representation, in databases, and in artificial
intelligence. The most basic modal logic is propositional modal logic ML, which
is just the fixed-point free part of the µ-calculus (see Chapter 10). But modal
logics, broadly conceived, form a family of many different formalisms, including
temporal logics, description logics, process logics, etc, many of which are also
closely related to automata. An important reason for the successful applications
of modal logics is their good balance between expressive power and computa-
tional complexity. This means that on the one hand, the relevant statements for
many applications are expressible in these logics, and on the other hand, the
usual reasoning tasks are decidable and admit reasonably efficient algorithms.
In particular, the satisfiability problem is decidable for most modal logics,
including the modal µ-calculus (for instance via automata based methods). This
in sharp contrast with first-order logic where even quite simple fragments are un-
decidable. Vardi [189] explicitely formulated the question, to find reasons for the
robust decidability properties of modal logics. Note, however, that satisfiability
problems tend to be of relatively high complexity. Even for plain propositional
logic satisfiability is NP-complete and for more powerful logics it quickly be-
comes Pspace- or Exptime-hard. For instance the satifiability problems for
propositional modal logic ML and the modal µ-calculus are Pspace-complete
[109] and Exptime-complete [54].
There is a standard translation of ML into first-order logic (FO), inductively
taking a modal formula ϕ to a first-order formula ϕ∗ . This translation takes aϕ
to ∃y( Ea xy ∧ ϕ∗ (y) ) and [a]ϕ to ∀y( Ea xy → ϕ∗ (y) ). The first-order formulae
that correspond to ML-formulae under this translation form what in [4] is called
the modal fragment of first-order logic. It has been shown that this fragment has
interesting properties, in particular a semantic characterisation via bisimulation:
The properties definable in ML are precisely the properties that are first-order
definable and invariant under bisimulation (see Chapter 14). Another important
feature is the so-called tree model property which was already considered in [189].
We will look at a generalisation of this property in Section 18.4.
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 321-341, 2002.
Springer-Verlag Berlin Heidelberg 2002
322 Thoralf Räsch
Both bisimulation invariance and the tree model property are important for
the analysis of modal logics. In the following we will use these ideas for investi-
gating guarded logics.
To understand guarded logics let us come back to the question of Vardi
mentioned above: What are the reasons for the good algorithmic properties of
modal logics? Looking at a typical formula of the modal fragment, say ∃y(Ea xy∧
ϕ∗ (y) ), we note that we only need at most binary predicates and two variables.
This strong restriction might be an answer but there is another observation: We
only use quantifiers that are somehow restricted or guarded by the predicate E.
Let us look at this property in a more general context.
In [4], Andréka, van Benthem and Németi have introduced the guarded frag-
ment GF of first-order logic which we will define in the next section. Roughly,
the idea is to consider arbitrary first-order languages but only allow guarded
quantifiers. The concept of a guarded quantifier needs to be defined and we will
see that there are several different ways to do it. Even a rather liberal notion
of guardedness, clique-guarded quantification, leads to decidable extensions of
propositional modal logic and the µ-calculus. Moreover, these clique-guarded
logics will give us the possibility to prove general statements about guarded log-
ics. In fact, we prove in Theorem 18.23 a general decidability result for various
logics, particularly for guarded logics.
With this goal in mind we will define in Section 18.2 the guarded logics GF
and CGF, and their fixed point extensions µGF and µCGF (that correspond
to thew modal µ-calculus). In Section 18.3 we shall prove some basic properties
of these logics, e.g., that GF can define precisely the model classes that are
first-order definable and closed under guarded bisimulation. We will prove that
even our the most liberal guarded fixed point logic µCGF has a (generalised) tree
model property (Theorem 18.17). In Section 18.5 we define guarded second-order
logics, GSO and CliqueSO, which generalize monadic second-order MSO. We
shall then be prepared to prove the general decidability result, Theorem 18.23),
in Section 18.6.
Note that this decidability result is based on the decidability of monadic the-
ory of trees and therefore does not provide good complexity bounds for guarded
logics. For complexity issues and decision procedures for guarded logics that are
based automata-theoretic methods we refer to Chapter 19. Further information
on guarded logics can be found in [65, 67, 69, 68, 71, 72, 73, 74, 75, 86]. The
exposition in this chapter is based on [67, 72, 73, 74].
∃ȳ( α(x̄, ȳ) ∧ ϕ(x̄, ȳ) ) or ∀ȳ( α(x̄, ȳ) → ϕ(x̄, ȳ) ),
Let us start with the classical guarded logic, the so-called guarded fragment (GF)
defined by induction as follows:
Let us stress that all free variables of the formula must be contained in the
guard. Note that we just have generalised the modal concept we have spoken
about in the first section. Clearly, the aforementioned translation of modal logics
into first-order logic uses only guarded quantification, so we see immediately that
the modal fragment is contained in GF. The guarded fragment generalises the
modal fragment by dropping the restrictions to use only two variables and only
monadic and binary predicates, and retains the restriction that quantifiers must
be guarded.
We next consider a more liberal notion of guarded quantification and define the
clique-guarded fragment CGF of first-order logic. Although we will have more
freedom to take formulae, CGF still has the nice properties we are expecting
from a suitable generalisation of ML.
324 Thoralf Räsch
A set of nodes in a graph is called a clique if every node of that set is adjacent
to any other node of it.
(c”) If ϕ(x̄, ȳ) is in CGF, then ∃ȳ( clique(x̄, ȳ) ∧ ϕ(x̄, ȳ) ) and ∀ȳ( clique(x̄, ȳ) →
ϕ(x̄, ȳ) ) belong to CGF where free(ψ) ⊆ free(clique) = {x̄, ȳ}.
Definition 18.6. The guarded fixed point logics µGF and µCGF are obtained
by adding to GF and CGF, respectively, the following rules for constructing fixed
point formulae.
Let W be a k-ary relation symbol, x̄ = (x1 , . . . , xk ) a k-tuple of distinct
variables and ϕ(W, x̄) be a guarded formula that contains only positive oc-
currences of W , no free first-order variables other than x1 , . . . , xk and where
W is not used in guards.
Then we can build the formulae LFPW,x̄ (ϕ)(x̄) and GFPW,x̄ (ϕ)(x̄).
18 Introduction to Guarded Logics 325
The semantics of the fixed point formulae is the following. Given a structure
A providing interpretations for all free second-order variables in ϕ except W , let
ϕA be the operator on k-ary relations W ⊆ Ak defined by
The relations W α and W̃ α are called the stages of the LFP- or GFP-induction,
respectively, of ϕ(W, x̄) on A. Since the operator ϕA is monotone, we have W 0 ⊆
W 1 ⊆ · · · ⊆ W α ⊆ W α+1 ⊆ · · · and W̃ 0 ⊇ W̃ 1 ⊇ · · · ⊇ W̃ α ⊇ W̃ α+1 ⊇ · · · and
there exist ordinals α, α such that W α = LFP(ϕA ) and W̃ α = GFP(ϕA ). These
Countable Models
While the finite model property fails for guarded fixed point logics we recall that
the so-called Löwenheim-Skolem property holds even for the (unguarded) least
fixed point logic (FO + LFP), i.e., every satisfiable fixed point sentence has a
countable model.
326 Thoralf Räsch
Lemma 18.7 ([73]). Every satisfiable sentence in (FO+LFP), and hence every
satisfiable sentence in µGF, and µCGF, has a model of countable cardinality.
Proof. We only have to look at fixed point formulae. Therefore, let us consider
ψ(x̄) := LFPR,x̄ ( ϕ(R, x̄) )(x̄), with first-order formula ϕ such that A |= ψ(ā) for
some infinite A.
For any ordinal α, let Rα be the α-th stage of the least fixed point induction
of ϕ on A. Expand A by a monadic relation U , a binary relation <, and an
(m + 1)-ary relation S (where m is the arity of R) such that
S := {(u, b̄) | for some ordinal α ≤ γ, u is the α-th element of (U, <), b̄ ∈ Rα }.
In the expanded structure A∗ := (A, U, <, S), the stages of the operator ϕA are
defined by the sentence:
Here, ϕ[ Rȳ/∃z(z < u ∧ Sz ȳ) ](x̄) is the formula obtained from ϕ(R, x̄) by re-
placing all occurrences of subformula Rȳ by ∃z( z < u ∧ Sz ȳ ).
Fixed point logics have a close relationship to infinitary logics (with bounded
number of variables). In order to formulate general statements we will consider
the following logics.
Definition 18.8. GF∞ and CGF∞ are the canonical infinitary variants of the
guarded fragments GF and CGF, respectively. For instance, GF∞ extends GF
by the following rule
for building
new formulae: If Φ ⊆ GF∞ is any set of
formulae, then also Φ and Φ are formulae of GF∞ . The definition for CGF∞
is analogous.
In the following we explicitly talk about the clique-guarded case only, i.e.,
about µCGF and CGF∞ but all results apply to the guarded as well. The fol-
lowing simple observation relates µCGF and CGF∞ . Recall that the width of
a formula is the the maximal number of free variables in its subformulae.
18 Introduction to Guarded Logics 327
Lemma 18.9 ([73]). For each ϕ ∈ µCGF of width k and each cardinal γ, there
is a ϕ ∈ CGF∞ , also of width k, which is equivalent to ϕ on all structures of
cardinality up to γ.
Proof. Consider a typical fixed point formula LFPR,x̄ ( ϕ(R, x̄) )(x̄). For every
ordinal α there is a formula ϕα ∈ CGF∞ that defines the stage α of the fixed
point induction of ϕ. Indeed, let ϕ0 := ⊥, let ϕα+1 := ϕ[Rȳ/ϕα (ȳ)](x̄), that is,
the formula that one obtains from ϕ(R, x̄) if one replaces each atom Rȳ (for any
ȳ) by the formula ϕα (ȳ), and for limit ordinals λ, let ϕλ := α<λ ϕα (x̄).
But on structures of bounded cardinality, also the closure ordinal of any fixed
point formula is bounded. Hence, for every cardinal γ there is an ordinal α such
that LFPR,x̄ ( ϕ(R, x̄) )(x̄) is equivalent to ϕα (x̄) on structures of cardinality at
most γ. 2
One of the main tools for the analysis of the modal µ-calculus is the notion of
bisimulation. We are going to generalise this idea now in the context of guarded
logics.
For GF, the so-called guarded bisimulations play a fundamental role for char-
acterising the expressive power, in the same way as bisimulation is crucial for un-
derstanding modal logics. For instance, the characterisation theorem by van Ben-
them, saying that a property is definable in propositional modal logic if and only
if it is first-order definable and invariant under bisimulation, has a natural ana-
logue for the guarded fragment.
Lemma 18.10 ([4]). GF can define precisely the model classes that are first-
order definable and closed under guarded bisimulations.
Since A |= ϕ(ā), there exists a tuple ā in A such that A |= clique(ā, ā ) ∧
ψ(ā, ā ). Suppose—towards a contradiction—that some f ∈ I takes the tuples ā
to b̄. Since the set {a1 , . . . , an , a1 , . . . , am } is clique-guarded there exists a partial
isomorphism g ∈ I, taking ā to b̄ and ā to some tuple b̄ in B. But then the set
{b1 , . . . , bn , b1 , . . . , bm } must be clique-guarded as well and B |= ¬ψ(b̄, b̄ ), which
contradicts the induction hypothesis.
For the direction from (ii) to (i), let I be the set of all partial isomorphisms
ā → b̄, taking a clique-guarded tuple ā in A to a clique-guarded tuple b̄ in B
such that for all formulae ϕ(x̄) ∈ CGF∞ of width at most k, A |= ϕ(ā) if and
only if B |= ϕ(b̄). Since A and B cannot be distinguished by sentences of width
k in CGF∞ , I contains the empty map and is therefore non-empty. It remains
to show that I satisfies the ‘back and forth’ properties.
For the ‘forth’ property, take any partial isomorphism f : X → Y in I and any
clique-guarded set X in A of size at most k. Let X = {a1 , . . . , an , a1 , . . . , am }
where X ∩ X = {a1 , . . . , an }. We only have to show that there exists g ∈ I,
defined on X that coincides with f on X ∩ X .
18 Introduction to Guarded Logics 329
Suppose that we cannot find such g. For ā := (a1 , . . . , an ), ā := (a1 , . . . , am ),
and b̄ := f (ā), let T be the set of all tuples b̄ := (b1 , . . . , bm ) such that
{b1 , . . . , bn , b1 , . . . , bm } is clique-guarded in B. Since there is no appropriate
g ∈ I there exists for every tuple b̄ ∈ T a formula ψb̄ (x̄, ȳ) ∈ CGF∞ such
that A |= ψb̄ (b̄, b̄ ). But then we can construct the formula
ϕ(x̄) := ( ∃ȳ.clique(x̄, ȳ) ) {ψb̄ (x̄, ȳ) | b̄ ∈ T }.
We show next that we can find a similar characterisation for CGF and clique-
guarded bisimulation as we have already seen in Chapter 14 for the propositional
modal logic as bisimulation-invariant fragments of first-order logic. The proof is
a straightforward adaption of van Benthem’s proof for modal logic. Recall that
every structure has an ω-saturated elementary extension.
Proof. We have already established that CGF-sentences (in fact, even sentences
from CGF∞ ) are invariant under clique-guarded bisimulation. For the converse,
suppose that ψ is a satisfiable first-order sentence that is invariant under clique-
guarded bisimulation. Let Φ be the set of sentences ϕ ∈ CGF such that ψ |= ϕ.
It suffices to show that Φ |= ψ.
Indeed, then by the compactness theorem, already a finite conjunction of
sentences from Φ will then imply, and hence be equivalent to, ψ.
Since ψ was assumed to be satisfiable, so is Φ. Take any model A |= Φ. We
have to prove that A |= ψ. Let TCGF (A) be the CGF-theory of A, i.e., the set of
all CGF-sentences that hold in A.
Claim. TCGF (A) ∪ {ψ} is satisfiable.
Otherwise, there were sentences ϕ1 , . . . , ϕn ∈ TCGF (A) such that ψ |= ¬(ϕ1 ∧
· · · ∧ ϕn ). Hence ¬(ϕ1 ∧ · · · ∧ ϕn ) is a CGF-sentence implied by ψ and is there-
fore contained in Φ. But then A |= ¬(ϕ1 ∧ · · · ∧ ϕn ) which is impossible since
ϕ1 , . . . , ϕn ∈ TCGF (A). This proves the claim.
Now, take any model B |= TCGF (A)∪{ψ}, and let A+ and B+ be ω-saturated
elementary extensions of A and B, respectively.
Claim. A+ and B+ are clique-bisimilar.
In order to prove the claim, let I be the set of partial isomorphisms f : X → Y
from clique-guarded subsets of B+ such that, for all formulae ϕ(x̄) in CGF and
all tuples ā from X, we have that A+ |= ϕ(ā) if and only if B+ |= ϕ(f ā).
330 Thoralf Räsch
The fact that A+ and B+ are ω-saturated implies that the ‘back and forth’
conditions for the clique-guarded bisimulations are satisfied by I. Indeed, let
f ∈ I, and let X be any clique-guarded set in A+ , with X ∩ X = {a1 , . . . , ar }
and X \ X = {a1 , . . . , as }. Let Φ be the set of all formulae in CGF of the form
ϕ(f a1 , . . . , f ar , y1 , . . . , ys ) such that A+ |= ϕ(a1 , . . . , ar , a1 , . . . , as ).
For every formula ϕ(f ā, ȳ) ∈ Φ, we have A+ |= ( ∃ȳ.clique(ā, ȳ) )ϕ(ā, ȳ) and
therefore also B+ |= ( ∃ȳ.clique(f ā, ȳ) )ϕ(f ā, ȳ). Hence, Φ is a consistent type of
(B+ , f ā) which—by ω-saturation—is realised in B+ by some fixed tuple b̄ such
that (f ā, b̄) is clique-guarded. And so, the function g taking ā to f ā and ā to b̄
is a partial isomorphism with domain X that coincides with f on X ∩ X .
The ‘back’ property is proved in the same way, exploiting that A+ is ω-
saturated.
We can now complete the proof of the theorem: Since B |= ψ and B+ is an
elementary extension of B, we have that B+ |= ψ. By assumption, ψ is invariant
under clique-guarded bisimulation, so A+ |= ψ and therefore also A |= ψ. 2
An analogous result applies to clique-k-bisimulations and CGF-sentences of
width k, for any k ∈ N .
We will now define the notion of tree width which is an important tool in graph
theory as well. It measures how closely a structure resembles a tree. Informally,
a structure has tree width ≤ k if it can be covered by (possibly overlapping)
substructures of size at most k + 1 which are arranged in a tree-like manner. So,
forests will have tree width 1 and cycles tree width 2 (cf. Figure 18.1).
(i) For every guarded set X in A there exists a node v of T with X ⊆ F (v).
(ii) For every element b of A, the set of nodes {v ∈ V : b ∈ F (v)} is connected
(and hence induces a subtree of T ).
Lemma 18.15 ([73]). Guarded and clique-guarded sets are contained in some
F (v) of a tree decomposition (T, (F(v) | v ∈ T )).
Proof. By definition for guarded sets this is true. We show that it also holds for
a general clique-guarded set X. For each a ∈ X, let Va be the set of nodes v such
that a ∈ F (v). By definition of a tree decomposition, each Vb induces a subtree
18 Introduction to Guarded Logics 331
Fig. 18.1. The upper part shows a graph (circle) with seven nodes divided into clusters
of size ‘2 + 1’ whereas the lower one gives us the arrangement of these clusters in a
tree-like manner. Note, each edge of the circle is contained in at least one cluster and
the arrangement is in the sense of (ii) of Definition 18.14.
The definition is very general since it only requires that the bound t(ϕ) on
the tree width of a model for ϕ must be computable from ϕ.
We can prove the tree model property for the logics we are considering.
Theorem 18.17 ([73]). Every satisfiable sentence in µCGF with width k has
a countable model of tree width at most k − 1. In particular, µCGF has the
generalised tree model property.
Proof. We can unravel any given structure to get a bisimilar tree-like structure.
The idea is to look at the local situation of the structure and paste copies of
small parts of it together, arranged as a tree.
332 Thoralf Räsch
ML FO GF FO
Lemma 18.19 ([74]). The following fragments of second-order logic are equally
expressive.
Proof. It obviously suffices to present translations from (i) and (iii) to (ii).
For the direction from (i) to (ii) consider a second-order variable X in a
formula according to (i) which is meant to range over arbitrary rather than
guarded relations. Consider first the case of sentences. For any atom X occurring
in the scope of a guarded quantification (Qȳ.α(ȳ z̄))ϕ where the occurrence of
x̄ is free in ϕ, the x̄ all occur in α. It follows that only truth values of X x̄ for
guarded tuples have an influence on the truth value of ϕ.
For formulae with free variables the quantifier-free part (w.r.t. first-order
quantification) may depend on truth-values for unguarded tuples. However, since
the number of free variables is fixed, there is only a fixed number of possibilities
for the second-order variables that can be explicitly enumerated.
So, if ∃Xϕ(x̄) is a subformula of the given formula type (i) that occurs outside
of any first-order quantifier we translate it into type (ii) as follows: Let H =
H(X, x̄) be the set of all {X}-structures with universe x̄. For this transformation
we assume that no variable in ϕ is reused in quantifications. Replace ∃Xϕ(x̄)
with ∃X A∈H ϕA (x̄) where ϕA is obtained from ϕ by substituting all atoms
X ȳ where ȳ ⊆ x̄ with ⊥ if A |= X ȳ, and with otherwise.
For the direction from (iii) to (ii) it suffices to show that unrestricted first-
order quantification can be simulated by guarded second-order quantification
over GF ranging over monadic variables: Each element variable x is replaced by
a set-variable X, and we use the following rules for translating formulae:
x=y → ∀x(Xx ↔ Y x)
Rx̄ → (∃x̄.Rx̄) Xi xi
i
Z x̄ → ( ∃x̄.guarded(x̄) )( Xi xi ∧ Z x̄)
i
∃xϕ(x, ȳ) → ∃X( ϕ(X, ȳ) ∧ singleton(X) )
334 Thoralf Räsch
t
guarded(x1 , . . . , xn ) := ∃ȳ(Ri ȳ ∧ xl = yj )
i=1 l j
Proof. Obviously, we have MSO ⊆ GSO ⊆ SO. We now show that both inclu-
sions are strict.
First we consider Hamiltonicity of graphs, i.e. the question whether a given
graph contains a closed walk that contains every vertex exactly once. This prop-
erty can be expressed by the following GSO-formula:
∃H ( ∀x∀y (Hxy → Exy) ∧ ∀x (∃=1 y Hxy ∧ ∃=1 yHyx) ∧
∀X[ (∃xXx ∧ ∀x∀y(Hxy ∧ Xx → Xy)) → ∀xXx ]
Evaluated on a graph G = (V, E) the formula says that there exists a H ⊆ E with
unique successors and predecessors such that (V, H) is connected. This means
that G has a Hamilton cycle. As Hamiltonicity is known not to be expressible
in MSO (see [47]), this shows that GSO is more expressive than MSO.
In order to prove the second part we show that GSO collapses to MSO over
words. Hence, GSO is not more expressive than MSO over words, i.e., able to
define exactly the regular languages. On the other hand, full second-order logic
is known to capture the polynomial-time hierarchy and, hence, much stronger
than MSO.
We represent words w = w1 · · · wn−1 ∈ A∗ by word structures ({0, . . . , n −
1}, S, (Pa | a ∈ A)) where S = {(i, i + 1) | i < n − 1} and Pa is the set of positions
in the word carrying the letter a, i.e., Pa = {i < n | wi = a}. The predicate of
maximal arity in a word structure is the successor relation, so a guarded set is
either a singleton or a set {i, i + 1}. As guarded n-ary relation therefore contains
only n-tuples (a1 , . . . , an ) such that {a1 , . . . , an } ⊆ {i, i + 1} for some i and
can therefore be encoded by a sequence of monadic relations. For instance a
guarded n-ary relation X can be represented by (Xu | u ∈ {0, 1}n ), where for
each u = (u0 , . . . , un−1 ), Xu := {i < n − 1 | (i + u0 , . . . , i + un−1 ) ∈ X}. This
was all we needed for our goal. 2
Similarly to GSO, we now can apply this idea to the clique-guarded case.
18 Introduction to Guarded Logics 335
Is is also clear that GSO is strictly more powerful than µGF: For instance as
GSO includes all of MSO, it is neither decidable nor invariant under guarded
bisimulation.
It remains to prove the claim now. Consider a least fixed point expression of
the form LFPX,x̄ ( ϕ(X, x̄) ). Inductively, we assume that all fixed points within
ϕ are in strictly guarded form. Looking at X-atoms in ϕ(X, x̄), we distinguish
the following cases:
336 Thoralf Räsch
for all occurrences of X apart from those of type (iii). As ( LFPX (ϕ) )g ⊆
LFPX,x̄ ( ϕ(X, x̄) ) and by monotonicity, we clearly have
Note that, the only remaining free occurrences of X in ϕ( ( LFPX (ϕ) )g , X, x̄)
are at the quantifier free level. It follows that the fixed point iteration in
is bounded in the sense that the fixed point is attained within an uniformly
bounded finite number of iterations, since there are only finitely many quantifier
free types over a vocabulary enriched by names for all the X-free constituents
of ϕ( ( LFPX (ϕ) )g , X, x̄) which are static for the fixed point process in question.
By unravelling this finite number of iterations within GF we can conclude that
LFPX,x̄ ( ϕ(X, x̄) ) is GF-definable from strictly guarded fixed points. 2
The proof is based on the following powerful decidability result which can be
found in [163, 185]:
Theorem 18.24 (Shelah, LeTourneau). The monadic second-order theory of
the class of all trees is decidable.
From Chapter 12 we know that SωS the MSO-theory of countable trees is
decidable. So, we are going to reduce the given logic L to the CliqueSO-theory
of trees with bounded tree width which we will further reduce to SωS where we
know that satisfiability is decidable.
Let Ck (τ ) be the class of all τ -structures of tree width at most k. We are
going to prove the statement by reducing, for every k ∈ N , the CliqueSO-theory
of Ck (τ ) to the monadic second-order theory of trees.
consistent(R̄) := ā∈K m ∀x∀y((Exy ∧ a∈ā (Qa x∧Qa y) → (Rā x ↔ Rā y)).
The conjunction over all these conditions forms a first-order axiom θ over the
vocabulary τ ∗ := {E} ∪ {Qa | a ∈ K} ∪ {Rā | ā ∈ K m }. Given a tree structure T
with underlying tree T = (V, E) and monadic predicates Qa and Rā satisfying
θ, we obtain a structure D such that T (D) = T as follows:
For every constant a ∈ K, we call two nodes u and w of T a-equivalent
if T |= Qa v for all nodes v on the link between u and w. Clearly, this is an
equivalence relation on QTa . We write [v]a for an a-equivalence class of the node
v. The universe of D is the set of all a-equivalence classes of T for a a ∈ K, i.e.,
D := { [v]a | v ∈ T, a ∈ K, T |= Qa v}. For every m-ary relation symbol R in τ ,
we then define
RD := { ( [v1 ]a1 , . . . , [vm ]am ) | T |= Ra1 ···am v for some
(and hence all) v ∈ [v1 ]a1 ∩ · · · ∩ [vm ]am }.
The Translation
For every formula ϕ(x1 , . . . , xm ) ∈ CliqueSO(τ ) and every tuple ā = (a1 , . . . , am )
over K, we now construct a monadic second-order formula ϕā (z̄) of vocabulary
τ ∗ , with one free variable. The formula ϕa (z) describes in the associated tree
structure T (D) the same properties of guarded tuples as ϕ(x̄) does in D. We will
make this statement more precise below. To define this translation we exploit
the fact that clique-guarded tuples in D are somehow local in T (D), i.e., they
coexist at some node of T (D)—cf. Lemma 18.15.
On a directed tree T = (V, E) we can express that U contains all nodes on
the link between x and y by the formula
connect(U, x, y) := U x ∧ U y ∧ ∃r(U r ∧ ∀z(Ezr → ¬U z)
∧∀w∀z( Ewz ∧ U z ∧ z = r → U w) ).
For any set ā ⊆ K we can then construct a monadic second-order formula
linkā (x, y) := ∃U ( connect(U, x, y) ∧ ∀z(U z → Qa z) )
a∈ā
saying that the tuple ā occurs at all nodes on the link between x and y.
Lemma 18.25. For every tuple ā ∈ K m , let
cliqueā (z) := ∃y( linka,a (y, z) ∧ Rb̄ y ).
a,a ∈ā R∈τ b̄:a,a ∈b̄
• The nodes a and a occur at all nodes on the link from v to w and hence
represent the same elements d and d at w as they do at v.
• T (D) |= Rb̄ w for some predicate R and some tuple b̄ that contains both a
and a . By induction hypothesis, this means that d and d are components
of some tuple d̄ such that D |= Rd̄ .
Hence T (D) |= cliqueā (v) if and only if the tuple d¯ induces a clique in the
Gaifman graph G(D). 2
A clique-guarded relation X ⊆ Dm consists only of local tuples. Therefore,
X can be represented in the same way as the basic relations of D by a tuple
X̄ = (Xā | ā ∈ K m ) of monadic predicates on T (D). So, we define
clique-guarded(X̄) := consistent(X̄) ∧ ∀y( Xā y → cliqueā (y) ).
ā∈K m
The Showdown
We are now well-prepared to bring it all together using techniques from [73] with
the following two theorems:
340 Thoralf Räsch
The only cases that need to be discussed here are first-order and second-
order quantifications. If ϕ(x̄) = (∃ȳ .clique(x̄, ȳ))η(x̄, ȳ) and D |= ϕ(d),¯ then
there exists a tuple d̄ such that D |= clique(d, d̄ ) ∧ η(d, d̄ ).
¯ ¯
Let U be the set of nodes on the link between v and w. Then, the tuple d¯
occurs in F(u) for all nodes u ∈ U . It follows that T (D) |= linkā (v, w). Hence,
T (D) |= ϕā (v).
Conversely, if T (D) |= ϕā (v), then there exists a node w such that the con-
stants ā occur at all nodes on the link between v and w (and hence correspond
to the same tuple d) ¯ and such that T (D) |= clique (w) ∧ η (w) for some tuple
āb̄ āb̄
b̄. By induction hypothesis this implies that D |= clique(d, ¯ d̄ ) ∧ η(d,
¯ d̄ ) for some
tuple d̄ , hence D |= ϕ(d).
¯
With ϕ(x̄) = ∃Y θ the claim follows immediate from the induction hypothesis
and from Lemma 18.26. 2
Theorem 18.28. For each k ∈ N , the CliqueSO-theory of Ck is decidable.
Here, θ is the first-order axiom expressing that the tree T expanded by the
relations Q̄ and R̄ is a tree structure T (D) associated to some τ -structure D.
By Theorem 18.27, T (D) |= ∀xϕ? (x) if and only if D |= ϕ. Hence ϕ is true on
all structures of tree width at most k if and only if ϕ∗ is true on all trees. 2
Theorem 18.23 now follows immediately: Given an arbitrary sentence we test
satisfiability via the tree model property by testing satisfiability of its translation
into CliqueSO on trees. In this way we obtain an equivalent satifiability problem,
which is decidable by Theorem 18.28.
19 Automata for Guarded Fixed Point Logics
19.1 Introduction
The guarded fixed point logics µGF and µCGF introduced in the previous chap-
ter extend the guarded fragments of first-order logic GF and CGF on the one
hand and the modal µ-calculus on the other hand. Thereby, the expressive power
of the underlying formalisms is increased considerably. On transition systems,
for instance, µGF already subsumes the µ-calculus with backwards modalities.
Hence, the question arises, whether these logics are still manageable algorithmi-
cally. In this chapter we will study the complexity of their satisfiability problems.
As a consequence of the general criterion stated in Theorem 18.23, it fol-
lows that the satisfiability problems for µGF and µCGF are decidable. Yet, the
argument does not allow us to derive precise complexity bounds for the deci-
sion problem. A lower bound can be obtained from the respective results for Lµ
and GF. For Lµ the satisfiability problem is Exptime-complete [54], whereas
for GF it is complete for 2Exptime [73]. However, if we consider formulae of
bounded width, i.e., with a bounded number of variables, it becomes Exptime-
complete as well.
Following Grädel and Walukiewicz [75, 73] we will prove that even for µCGF,
the strongest logic considered, the satisfiability problem is still in 2Exptime
in the general case, and in Exptime for formulae of bounded width. In other
words, the fixed point extensions of guarded logics are almost for free in terms
of complexity of the satisfiability problem.
Given the expressive power of these logics, this result is rather surprising. For
instance, in contrast to Lµ , already the weakest guarded fixed point logic µGF
lacks the finite model property. An example of a formula with only infinite
models was given in the previous chapter:
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 343-355, 2002.
Springer-Verlag Berlin Heidelberg 2002
344 Dietmar Berwanger and Achim Blumensath
19.2 Requisites
19.2.1 Clique Guarded Fixed Point Formulae
When speaking of formulae we always mean µCGF-formulae as introduced in
the previous chapter. To simplify our notation we will, however, omit the clique
guards, i.e., instead of (∃x̄.clique(x̄))η(x̄) we will write ∃x̄.η(x̄) and accordingly
for universal formulae.
Furthermore, we will assume that all formulae are well named and in negation
normal form, that is, fixed point variables are defined at most once and negation
applies to atomic formulae only. Clearly, every µCGF-formula can be rewritten
to meet these requirements.
A crucial parameter of a formula is its width which is defined as the greatest
number of free variables occurring in a subformula. Equivalently, a formula has
width k iff it can be transformed, by renaming of variables, so that it uses only
k variables. In the following we will always assume that every formula of width k
is written with the variables {x0 , x1 , . . . , xk−1 }.
a
a
b c
b c
d d
e g
e f g
f
Fig. 19.1. A structure with relations of arity 1, 2, and 3 and its Gaifman graph
refer to at most k structure elements at the same time, which, moreover, have
to be guarded. On account of this, we associate to a given structure A a tree
whose nodes are labelled by the substructures of A induced by at most k guarded
elements. In addition, the overlap of two adjacent nodes is stored in the label of
their common edge.
Let us fix some notation for the remainder of this chapter. The set of guarded
subsets of size at most k of a σ-structure A is denoted by
Γk (A) := { K ⊆ A | K is k-clique-guarded in A }.
The substructures induced by these sets are mapped onto the fixed universe
[k] = {0, . . . , k − 1} and then arranged to form a tree while keeping track of
overlaps along the edges. Thus, the nodes of the resulting trees are labelled by
the alphabet
while the edges are labelled by subsets of [k]. We call trees labelled by these
alphabets shortly k-type trees. When we speak about a D-edge, we mean an
edge labelled with D ⊆ [k], and a D-neighbour or D-successor of a node is a
neighbour respectively a successor along some D-edge.
(i) The root of T is labelled with the empty structure (∅, σ) and all outgoing
edges are labelled with ∅.
(ii) Every node v ∈ Γk (A)∗ K is labelled with an isomorphic copy C of A|K , the
restriction of A to K ∈ Γk (A).
(iii) If π : A|K → C and π : A|K → C are isomorphisms labelling, respectively,
a node v ∈ Γk (A)∗ K and its successor v = vK , then π and π agree on
K ∩ K and the edge (v, v ) is labelled with π(K ∩ K ).
346 Dietmar Berwanger and Achim Blumensath
Remark 19.2. It is easy to see that for every D-edge (v, v ) of an unravelling
tree T the following conditions hold:
(i) Consistency: the labels C of v and C of v agree on D, that is, C|D = C |D .
(ii) Completeness: for any H ⊆ [k] the H-successors of v and v agree on
D ∩ H, i.e., there is a one-to-one map assigning to each H-successor w of v
an H-successor w of v such that the labels of w and w agree on D ∩ H.
Definition 19.3. Given a consistent k-type tree T , consider the disjoint sum of
its node labels,
D := · { (C, v) | C is the label of v ∈ T }.
Let ∼ be the least equivalence relation on the universe of D with
(i, v) ∼ (i, v ) if v is a successor of v and i is in the label of (v, v ).
1
2
3
{1} {2, 3}
2 2
1 1 3
{1, 3} {1, 2}
{2}
3
2
1 2 3 2
1
1
{1}
{3}
1
1
2
3
3
{1}
{2, 3}
2 1
3 1 2
3 2
3 2 1 2 1
1 3
{3} {1, 3}
1 1
3 2 3
(iii) every node v is labelled with A|K for some K ∈ Γk (A) via an isomorphism π.
Moreover, for each K ∈ Γk (A) there is a node v labelled with A|K , such
that all edges on the path between v and v include π(K ∩ K ) in their
labels.
c b c b
...... ......
a d a
e f g
Fig. 19.3. The structure recovered from the decomposition tree in Fig. 19.2.2
The language L(A) accepted by a tree automaton A is the set of all trees T ,
such that Player 0 has a winning strategy in the game G(A, T ).
Usually, automata are defined as devices scanning their input only in one
direction. However, for our purpose it is convenient to allow them to move back-
wards and remain still as well.
Definition 19.12. An alternating two-way tree automaton is given in the
same way as a (one-way) alternating automaton,
A2 = (Q, Σ, ∆, δ, qI , Ω)
where acceptance is defined in a different way. The game G(A2 , T ) associated
to a two-way automaton A2 and a tree T is the parity game obtained as in
Definition 19.11, but replacing rule (2) with
(ii’) either v = v or v is a d-neighbour of v.
The language L(A2 ) accepted by a two-way tree automaton A2 is the set of all
trees T such that Player 0 has a winning strategy in the game G(A2 , T ).
(iii) If ϕ = η ∧ ϑ or ϕ = η ∨ ϑ then
Finally, if the fixed point variables of ψ occur in the order Z1 , . . . , Zn the parity
condition is given by
2i ϕ = Zi x̄ and Zi is a GFP-variable,
2i + 1 ϕ = Zi x̄ and Zi is an LFP-variable,
Ω(ϕ, β) := 2n + 4 ϕ = ∀ȳ.η,
2n + 3 ϕ = ∃ȳ.η,
2n + 2 otherwise.
\ \
By the construction of G, we can easily verify that this mapping induces a
congruence relation ∼
among the positions of G,
(ϕ, β, v ) iff
(ϕ, β, v) ∼ (ϕ, β, v) = (ϕ, β, v ),
which is also a bisimulation on G.
Consider now the (strong homomorphic) image G of G under · . On the one
hand, G and G are bisimilar via · and, consequently, the same player has a win-
ning strategy in both plays. On the other hand, G is almost the model checking
game G = G(A, ψ) as defined in [11]. The only difference arises at positions
(ϕ, χ) where ϕ is an existential or universal formula, say ϕ = ∃ȳη(x̄, ȳ). Then,
the model checking game allows moves to (η, χ ) with χ such that
(i) χ and χ agree on the values of x̄ and
(ii) A, χ |= clique(x̄, ȳ),
whereas in G the legal moves go either to (ϕ, χ ) with χ as above, or to (η, χ).
352 Dietmar Berwanger and Achim Blumensath
Nevertheless, we will show that the same player wins both G and G. If
Player 0 has a winning strategy in the model checking game G , he can also
as long as no existential formula is met. Otherwise, at
play this strategy in G,
positions (ϕ, χ) as above, he can imitate the move to the position (η, χ ) he
would perform in G by taking two steps:
(i) move to (ϕ, χ ); this is possible since, for every χ agreeing with χ on the
free variables of ϕ, the position (ϕ, χ ) is reachable from (ϕ, χ) in one step.
(ii) At (ϕ, χ ) it’s still Player 0 turn: move to (η, χ ).
Towards a contradiction, let us assume that Player 1 wins this play. Then, after
any universal formula ϕ = ∀ȳη(x̄, ȳ) occurring in the play, there can follow
only finitely many positions with ϕ until Player 1 chooses some position (η, χ );
otherwise he would lose with the highest even priority. But then, Player 1 also
wins by choosing (ϕ, χ ) right from position (ϕ, χ) and proceeding with (η, χ ).
However, these two moves translate into one move in the corresponding play of
G which leads Player 1 to a win in G despite Player 0’s winning strategy, which
is not possible. This concludes our proof that a player has a winning strategy in
the model checking game iff he has one in the acceptance game.
The correctness of our construction relies on the fact that the input trees are
complete in the sense of Remark 19.2 (ii). That is, if the current node is labelled
by a k-clique of the represented structure, then every other k-clique appears in
the label of some successor node. Unfortunately, it is very hard to check whether
a given tree satisfies this property. By letting Aψ run as a two-way automaton
A2ψ , we can relax this requirement and claim instead that every k-clique shall be
reachable via a finite path from the current node.
Proposition 19.15. Given a formula ψ of width k and a structure A, let T be
a k-decomposition tree of A. Then the automaton A2ψ accepts T iff A |= ψ.
Proof. The idea is to show that A2ψ runs on T in a similar way as its one-way
variant does on the k-unravelling tree T of A. Towards this we will transform
the acceptance game G(A2ψ , T ) by introducing shortcuts into a game which is
bisimilar to the acceptance game G(A, T ) of the one-way automaton.
Let G ∗ be the least game extending G := G(A2ψ , T ) by new transitions in such
a way that, whenever there are two transitions
Player 1 used a shortcut to, say (ϕ, β, v ), for ϕ a universal formula. At that
point, Player 1 can move step by step via finitely many positions (ϕ, β, w) along
the path leading to the destination of the shortcut. From there, the play proceeds
like in π. Clearly, Player 1 wins this play in G in contradiction to our assumption
on Player 0’s winning strategy.
The mapping · which was defined in the proof of Proposition 19.14 can be
applied to the positions of G ∗ . It induces a congruence relation on G ∗ and, as
such, a bisimulation between G ∗ and its strong homomorphic image G∗ . This
image is precisely the game G( Aψ , T ) which is bisimilar to G(Aψ , T ).
Accordingly, the automaton A2ψ accepts the k-decomposition tree T iff Aψ
accepts the k-unravelling tree T .
19.4 Satisfiability
The model checking automata introduced above operate correctly on inputs
which represent structures. But in order to solve the satisfiability problem this
does not suffice. We need to make sure that all inputs which do not represent
structures are rejected.
At a node where the elements i and j are not guarded, Player 1 can challenge
his opponent to find a node where {i, j} appear guarded, along a path where
these elements persist in the edge labels.
δ((i, j), C) =
·{true} C |= clique(i, j),
{i, j}{(i, j)} otherwise.
Also, Player 1 may pick a currently valid atomic fact to check whether it is indeed
preserved along the edges that contain all involved elements in their label.
δ(Rā, C) =
·{true} C |= Rā,
·{false} otherwise.
If the player agree on a local test, the run is finite: δ(true) = δ(false) = ∅.
On an infinite run, the automaton assumes forever either the state check or
some state (i, j). Since in the first case Player 0 should win, we set Ω(check) = 0.
In the second case, instead, Player 0 should lose, because he does not provide a
witness to the guardedness of i and j after a finite number of steps. To enforce
that, we set Ω((i, j)) = 1 for all (i, j) ∈ [k]2 .
It is easy to see that the above checks ensure the consistency and the guard-
edness of the input tree.
Lemma 19.17. The automaton A2k recognises the set of k-decomposition trees
of all σ-structures.
19.5 Complexity
Since Vardi’s automata work on trees with unlabelled edges, we have to remove
the edge labels and place them into their target node. Then, our automaton has
to verify the validity of taken transitions, thus, requiring a blow-up of its state
set by the size of the edge alphabet. Taking into account this modification, we
can transfer the complexity of the emptiness test of Vardi’s automata to our
model.
Carsten Fritz
This chapter is intended to give the reader a brief overview of some basic no-
tations and theorems regarding fixed points of monotone functions on complete
lattices. The main results stated and proved here are the Knaster-Tarski Theo-
rem [175] (Theorem 20.4), and Theorem 20.12 on the characterization of simul-
taneous fixed points. That is, this chapter provides proofs and some additional
insights to propositions introduced in Chapter 10.
Therefore, our main interest is the µ-calculus dealing with fixed points of
monotone functions on the complete lattice of subsets of the states of a Kripke
structure. Consequently, power set lattices will be our main models of complete
lattices, but our approach will be somewhat more general.
20.1 Preliminaries
We fix a complete lattice L = (L, ≤, , ⊥), i.e.
(1) L is a non-empty set,
(2) ≤ is a partial order on L such that every subset M ⊆ L has a supremum
and an infimum,
(3) , ⊥ ∈ L are the greatest and least elements, respectively, of L, i.e., for every
x ∈ L, ⊥ ≤ x ≤ holds.
Note that inf ∅ = , sup ∅ = ⊥. Our main instance of a complete lattice
P
(A), ⊆, A, ∅) of
P
is the power set lattice ( an arbitrary set A. For a subset
M ⊆ (A), we have inf M = M , sup M = M .
Let On be the class of ordinals. For a cardinal c, let c+ be the least ordinal
such that |c+ | > c.
Definition 20.1. Let f : L → L be a function.
(1) x ∈ L is a fixed point of f iff f (x) = x.
(2) x is the least (greatest) fixed point of f iff x is a fixed point of f and
x ≤ y (y ≤ x) holds for all fixed points y of f .
(3) f is monotone iff for all x, y ∈ L, x ≤ y implies f (x) ≤ f (y).
(4) f is inflationary iff x ≤ f (x) holds for all x ∈ L.
(5) We inductively define a sequence (f α )α∈On of elements f α ∈ L by
f 0 := ⊥,
f α+1 := f (f α ),
f λ := sup f α for limit ordinals λ.
α<λ
E. Grädel et al. (Eds.): Automata, Logics, and Infinite Games, LNCS 2500, pp. 359-364, 2002.
Springer-Verlag Berlin Heidelberg 2002
360 Carsten Fritz
Proof. (i) First, let f be monotone. To show that f is inductive, we use induc-
tion on α:
(α = 0): Trivial.
(α → α + 1): Using the monotonicity of f and the induction hypothesis, we
have f α+1 = f (f α ) ≥ f (f β ) = f β+1 for all β < α. Thus f β ≤ f α+1 for all
β < α + 1.
(α a limit ordinal): Immediately by the definition of f α .
The proof for inflationary f is trivial.
(ii) Let f be inductive and define g : L → L, x
→ sup{x, f (x)}. Obviously, g is
inflationary. By induction, we show ∀α ∈ On(g α = f α ):
(α = 0): g 0 = ⊥ = f 0
(α → α + 1): The induction hypothesis yields g α+1 = g(g α ) = g(f α ) =
sup{f α , f (f α )}. Since f is inductive, sup{f α , f (f α )} = f α+1 .
(α a limit ordinal): By induction hypothesis, g α = sup g β = sup f β = f α .
β<α β<α
(iii) Assume that there is no such α. Then, for every α < β < |L|+ , f α = f β .
That is, the set { f α ∈ L | α < |L|+ } ⊆ L has cardinality ||L|+ | > |L|.
Contradiction. If L is the power set lattice of a set A, there is an xα ∈
f α+1 \ f α for every α < |A|+ . Thus X := { xα | α < |A|+ } is a subset of
A, but |X| = ||A|+ | > |A|. Contradiction.
Definition 20.3. The least α ∈ On such that f α+1 = f α is the closure ordi-
nal of f . Notation: cl(f ). For monotone f , we define f ! := f cl(f ) .
The Knaster-Tarski Theorem [175] asserts the existence of a least and a greatest
fixed point of any monotone function on a complete lattice. More precisely, these
fixed points are the infimum and supremum, respectively, of certain subsets of
the complete lattice and can be generated inductively.
using induction on α.
(α = 0): f 0 = ⊥ ≤ x for all x ∈ L.
(α → α + 1): Let x ∈ Φ. By induction hypothesis, f α ≤ x. Thus we have
f α+1
= f (f α ) ≤ f (x) ≤ x, using the monotonicity of f .
(α a limit ordinal): By induction hypothesis, f β ≤ x holds for all β < α,
x ∈ Φ, which implies f α = sup f β ≤ x.
β<α
To generate the greatest fixed point in the same fashion, we introduce a dual
sequence (f ∗α )α∈On .
f ∗0 =
f ∗(α+1) = f (f ∗α )
f ∗λ = inf f ∗α for limit ordinals λ.
α<λ
P
Definition 20.8. For every function f : (A) → (A), the dual function P
P P
f : (A) → (A) of f is defined by f (X) := f (X) where X := A \ X.
Note that f = f .
P
Proof. (i) Let X, Y ∈ (A), X ⊆ Y . Thus Y ⊆ X, and f (Y ) ⊆ f (X) by the
monotonicity of f , which implies f (X) ⊆ f (Y ).
(ii) At first, we note that the first claim implies the second: If LFP(f ) = GFP(f )
then LFP(f ) = GFP(f ) = GFP(f ).
To prove the first claim, we show by induction that f α = f ∗α holds for all
α ∈ On.
(α = 0): f 0 = ∅ = A = f ∗0
(α → α + 1): We have
f α+1 = f (f α )
= f (f ∗α ) (Ind. Hyp.)
= f (f ∗α ) (by Def. 20.8)
= f ∗(α+1) (by Def. 20.6)
(α a limit ordinal): Here we have f α = fβ = fβ = f ∗β = f ∗α ,
β<α β<α β<α
using the induction hypothesis for the third equation.
f : L → L, (x0 , . . . , xn−1 )
→ (f0 (x0 , . . . , xn−1 ), . . . , fn−1 (x0 , . . . , xn−1 ))
20 Some Fixed Point Basics 363
is a monotone function as well and thus has a least and a greatest fixed point
(Theorem 20.4). These are called the simultaneous (least and greatest) fixed
points of f0 , . . . , fn−1 .
We now wish to compute the least and greatest fixed points of f by generating
nested fixed points of monotone functions defined on the lattices L0 , . . . , Ln−1 .
For the sake of brevity (and clarity), we restrict ourselves to the case n = 2 and
the computation of the least fixed point, but the generalization is straightfor-
ward.
Let g : L → L0 and h : L → L1 be monotone functions, and let f : L →
L, (x0 , x1 )
→ (g(x0 , x1 ), h(x0 , x1 )). Let LFP(f ) =: ((f ! )0 , (f ! )1 ) ∈ L0 ×L1 denote
the least fixed point of f , i.e., (f ! )i = pri (f ! ) (i = 0, 1). For α ∈ On, we define
fiα := pri (f α ) (i = 0, 1).
The following lemmas give us a computation recipe at hand:
For every x ∈ L0 , we define hx : L1 → L1 , y
→ h(x, y). The monotonicity of
h implies the monotonicity of hx , so we can generate, for every x ∈ L0 , the least
fixed point LFP(hx ) = h!x ∈ L1 (cf. Lemma 20.5).
since g is monotone.
Hence we show ∀α ∈ On(∀x, x ∈ L0 (x ≤0 x → hα x ≤1 hx )) by induction
α
on α:
(α = 0): Trivial.
(α → α + 1): Let x ≤0 x . We have hα+1
x = h(x, hα
x ) ≤1 h(x , hx ) = hx .
α α+1
β
(α a limit ordinal): Let x ≤0 x . hx = sup hx ≤1 sup hx = hx .
α β α
β<α β<α
Next, we show that (f ! )1 is a fixed point of h(f ! )0 ; this implies h!(f ! )0 ≤1 (f ! )1 :
h(f ! )0 ((f ! )1 ) = h((f ! )0 , (f ! )1 ) = pr1 (f ((f ! )0 , (f ! )1 )) = (f ! )1 , since ((f ! )0 , (f ! )1 )
is a fixed point of f .
Now we show e! ≤0 (f ! )0 . This implies h!e! ≤1 h!(f ! )0 , since x
→ h!x is mono-
tone.
(e! ≤0 (f ! )0 ): Using h!(f ! )0 ≤1 (f ! )1 and the monotonicity of g, we have
e((f ! )0 ) = g((f ! )0 , h!(f ! )0 ) ≤0 g((f ! )0 , (f ! )1 ) = pr0 (f ((f ! )0 , (f ! )1 )) = (f ! )0 , that
is, (f ! )0 ∈ { x ∈ L0 | e(x) ≤0 x }. Now since e! = inf{ x ∈ L0 | e(x) ≤ x }, we
have e! ≤0 (f ! )0 .
We now know that h!e! ≤1 h!(f ! )0 ≤1 (f ! )1 and e! ≤0 (f ! )0 .
To show that f ! = ((f ! )0 , (f ! )1 ) ≤ (e! , h!e! ) and hence (e! , h!e! ) = LFP(f ), it
suffices to establish ∀α ∈ On(f α ≤ (e! , h!e! )), as usual by induction on α.
(α = 0): Trivial.
364 Carsten Fritz
In other words,
1. Martı́n Abadi, Leslie Lamport, and Pierre Wolper, Realizable and unrealizable
specifications of reactive systems, Proceedings of the 16th International Collo-
quium on Automata, Languages and Programming, ICALP ’89, Lecture Notes in
Computer Science, vol. 372, Springer-Verlag, 1989, pp. 1–17. [40]
2. Luca de Alfaro and Thomas A. Henzinger, Concurrent omega-regular games, Pro-
ceedings of the 15th IEEE Symposium on Logic in Computer Science, LICS 2000,
IEEE Computer Society Press, 2000, pp. 141–154. [40]
3. Luca de Alfaro, Thomas A. Henzinger, and Freddy Y. C. Mang, The control of
synchronous systems, Proceedings of the 11th International Conference on Con-
currency Theory, CONCUR 2000, Lecture Notes in Computer Science, vol. 1877,
Springer-Verlag, 2000, pp. 458–473. [40]
4. Hajnal Andréka, István Németi, and Johan van Benthem, Modal logic and bounded
fragments of predicate logic, Journal of Philosophical Logic 27 (1998), no. 3, 217–
274. [321, 322, 327, 356]
5. André Arnold, The µ-calculus alternation-depth hierarchy is strict on binary trees,
Theoretical Informatics and Applications 33 (1999), no. 4–5, 329–340. [185, 195,
202]
6. Klaus Barthelmann, When can an equational simple graph be generated by hyper-
edge replacement?, Proceedings of the 23rd International Symposium on Mathe-
matical Foundations of Computer Science, MFCS ’98, Lecture Notes in Computer
Science, vol. 1450, Springer-Verlag, 1998, pp. 543–552. [263, 283, 318]
7. Johan van Benthem, Modal correspondence theory, Ph.D. thesis, Instituut voor
Logica en Grondslagenonderzoek van Exacte Wetenschappen, Universiteit van
Amsterdam, The Netherlands, 1976. [258]
8. , Dynamic bits and pieces, Tech. Report LP-97-01, Institute for Logic,
Language and Computation, University of Amsterdam, The Netherlands, 1997.
[356]
9. , Modal logic in two gestalts, Advances in Modal Logic, Volume II (Stan-
ford, California), CSLI Publications, 1998, pp. 73–100. [356]
10. Orna Bernholtz, Moshe Y. Vardi, and Pierre Wolper, An automata-theoretic ap-
proach to branching-time model checking, Proceedings of the 6th International
Conference on Computer Aided Verification, CAV ’94, Lecture Notes in Com-
puter Science, vol. 818, Springer-Verlag, 1994, pp. 142–155. [203]
11. Dietmar Berwanger and Erich Grädel, Games and model checking for guarded log-
ics, Proceedings of the 8th International Conference on Logic for Programming,
Artificial Intelligence and Reasoning, LPAR 2001, Lecture Notes in Artificial In-
telligence, vol. 2250, Springer-Verlag, 2001, pp. 70–84. [351, 356]
12. Achim Blumensath, Prefix-recognizable graphs and monadic second order logic,
Tech. Report AIB-06-2001, RWTH Aachen, Germany, 2001. [263, 264, 283, 318]
13. Julian C. Bradfield, The modal mu-calculus alternation hierarchy is strict, Pro-
ceedings of the 7th International Conference on Concurrency Theory, CON-
CUR ’96, Lecture Notes in Computer Science, vol. 1119, Springer-Verlag, 1996,
pp. 232–246. [185, 202]
14. , The modal µ-calculus alternation hierarchy is strict, Theoretical Com-
puter Science 195 (1998), no. 2, 133–153. [202]
366 Literature
53. E. Allen Emerson, Temporal and modal logic, Handbook of Theoretical Computer
Science (Jan van Leeuwen, ed.), vol. B: Formal Models and Sematics, Elsevier,
1990, pp. 995–1072. [203]
54. E. Allen Emerson and Charanjit S. Jutla, The complexity of tree automata and
logics of programs (exteded abstract), Proceedings of the 29th Annual Symposium
on Foundations of Computer Science, FoCS ’88, IEEE Computer Society Press,
1988, pp. 328–337. [130, 168, 203, 321, 343]
55. , Tree automata, mu-calculus and determinacy (extended abstract), Pro-
ceedings of the 32nd Annual Symposium on Foundations of Computer Science,
FoCS ’91, IEEE Computer Society Press, 1991, pp. 368–377. [39, 95, 103, 108,
130, 135, 138, 202, 354]
56. E. Allen Emerson, Charanjit S. Jutla, and A. Prasad Sistla, On model-checking
for fragments of µ-calculus, Proceedings of the 5th International Conference on
Computer Aided Verification, CAV ’93, Lecture Notes in Computer Science, vol.
697, Springer-Verlag, 1993, pp. 385–396. [95, 104, 108]
57. , On model checking for the µ-calculus and its fragments, Theoretical Com-
puter Science 258 (2001), no. 1–2, 491–522. [202]
58. E. Allen Emerson and Chin-Laung Lei, Efficient model checking in fragments of
the propositional mu-calculus (extended abstract), Proceedings of the Symposium
on Logic in Computer Science, LICS ’86, IEEE Computer Society Press, 1986,
pp. 267–278. [202, 203]
59. E. Allen Emerson and A. Prasad Sistla, Deciding full branching time logic, Infor-
mation and Control 61 (1984), no. 3, 175–201. [92]
60. Joost Engelfriet and Hendrik Jan Hoogeboom, X-automata on ω-words, Theoret-
ical Computer Science 110 (1993), 1–51. [39]
61. Javier Esparza, David Hansel, Peter Rossmanith, and Stefan Schwoon, Efficient
algorithms for model checking pushdown systems, Proceedings of the 12th Inter-
national Conference on Computer Aided Verification, CAV 2000, Lecture Notes
in Computer Science, vol. 1855, Springer-Verlag, 2000, pp. 232–247. [311]
62. Shimon Even, Graph algorithms, Pitman Publishing, London, 1979. [55]
63. Jeanne Ferrante and Charles W. Rackoff, The computational complexity of logical
theories, Lecture Notes in Mathematics, vol. 718, Springer-Verlag, 1979. [258]
64. Markus Frick and Martin Grohe, The complexity of first-order and monadic
second-order logic revisited, Proceedings of the 17th IEEE Symposium on Logic in
Computer Science, LICS ’02, IEEE Computer Society Press, 2002, pp. 215–224.
[258]
65. Harald Ganziger, Christoph Meyer, and Margus Veanes, The two-variable guarded
fragment with transitive relations, Proceedings of the 14th IEEE Symposium on
Logic in Computer Science, LICS ’99, IEEE Computer Society Press, 1999, pp. 24–
34. [322, 356]
66. Kurt Gödel, Über formal unentscheidbare Sätze der Principia Mathematica und
verwandter Systeme I, Monatshefte für Mathematik und Physik 38 (1931), 173–
198. [207]
67. Elisabeth Gonçalves and Erich Grädel, Decidability issues for action guarded
logics, Proceedings of the 2000 International Workshop on Description Logics,
DL 2000, 2000, pp. 123–132. [322, 336, 356]
68. Georg Gottlob, Erich Grädel, and Helmut Veith, Datalog LITE: A deductive query
language with linear time model checking, ACM Transactions on Computional
Logic 3 (2002), no. 1, 42–79. [322, 356]
Literature 369
69. Erich Grädel, The decidability of guarded fixed point logic, JFAK. Essay Decicated
to Johan van Benthem on the occasion of his 50th Birthday, CD-ROM (Jelle
Gerbrandy, Maarten Marx, Maarten de Rijke, and Yde Venema, eds.), Amsterdam
University, 1999. [322, 356]
70. , Decision procedures for guarded logics, Proceedings of 16th International
Conference on Automated Deduction, CADE ’99, Lecture Notes in Artificial In-
telligence, vol. 1632, Springer-Verlag, 1999, pp. 31–51. [356]
71. , On the restrainning power of guards, Journal of Symbolic Logic 64 (1999),
1719–1742. [322, 343, 356]
72. , Why are modal logics so robustly decidable?, Current Trends in Theo-
retical Computer Science, Entering the 21st Century (Gheorghe Paun, Grzegorz
Rozenberg, and Arto Salomaa, eds.), World Scientific, 2001, pp. 393–498. [322,
356]
73. , Guarded fixed point logics and the monadic theory of countable trees,
Theoretical Computer Science 288 (2002), 129 – 152. [322, 326, 327, 328, 329,
330, 331, 339, 343, 346, 356]
74. Erich Grädel, Colin Hirsch, and Martin Otto, Back and forth between guarded and
modal logics, ACM Transactions on Computional Logic 3 (2002), no. 3, 418–463.
[322, 333, 334, 335, 356]
75. Erich Grädel and Igor Walukiewicz, Guarded fixed point logic, Proceedings of the
4th Annual IEEE Symposium on Logic in Computer Science, LICS ’99, IEEE
Computer Society Press, 1999, pp. 45–54. [322, 343, 356]
76. Yuri Gurevich, Monadic second-order theories, Model-Theoretical Logics (Jon
Barwise and Solomon Feferman, eds.), Springer-Verlag, 1985, pp. 479–506. [258]
77. Yuri Gurevich and Leo Harrington, Trees, automata and games, Proceedings of
the 14th Annual ACM Symposium on Theory of Computing, STOC ’82, ACM
Press, 1982, pp. 60–65. [39, 95, 108, 135, 141]
78. Yuri Gurevich, Menachem Magidor, and Saharon Shelah, The monadic theory of
ω2 , Jounal of Symbolic Logic 48 (1983), 387–398. [258]
79. Stephan Heilbrunner, An algorithm for the solution of fixed-point equations for
infinite words, R.A.I.R.O. Informatique théorique/Theoretical Informatics 14
(1980), no. 2, 131–141. [228]
80. B. Herwig, Zur Modelltheorie von Lµ , Ph.D. thesis, Universität Freiburg, Ger-
many, 1989. [130]
81. David Hilbert, Mathematische Probleme. Vortrag, gehalten auf dem interna-
tionalen Mathematiker-Kongress zu Paris 1900, Nachrichten von der Königl. Ge-
sellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse
(1900), 253–297. [207]
82. Colin Hirsch, Guarded logics: Algorithms and bisimulation, Ph.D. thesis, RWTH
Aachen, Germany, 2002. [356]
83. Wilfrid Hodges, Model theory, Encyclopedia of Mathematics and its Applications,
Cambridge University Press, 1993. [226]
84. Ian Hodkinson, Loosely guarded fragment has finite model property, Studia Logica
70 (2002), 205–240. [356]
85. A. Hoffmann and Richard M. Karp, On nonterminating stochastic games, Man-
agement Science 12 (1966), 359–370. [108, 125]
86. Eva Hoogland, Maarten Marx, and Martin Otto, Beth definability for the guarded
fragment, Proceedings of the 6th International Conference on Logic for Program-
ming and Automated Reasoning, LPAR ’99), Lecture Notes in Computer Science,
vol. 1705, Springer-Verlag, 1999, pp. 273–285. [322, 356]
370 Literature
87. John E. Hopcroft and Jeffrey D. Ullman, Introduction to automata theory, lan-
guages, and computation, Addison-Wesley, 1979. [44, 63, 236]
88. Robert Hossley and Charles W. Rackoff, The emptiness problem for automata
on infinite trees, Proceedings of the 13th Annual Symposium on Switching and
Automata Theory, IEEE Computer Society Press, 1972, pp. 121–124. [168]
89. David Janin, Propriérés logiques du non-déterminisme et µ-calcul modal, Ph.D.
thesis, LaBRI – Université de Bordeaux I, France, 1995. [249]
90. David Janin and Igor Walukiewicz, Automata for the modal mu-calculus and re-
lated results, Proceedings of the 20th International Symposium on Mathematical
Foundations of Computer Science, MFCS ’95, Lecture Notes in Computer Science,
vol. 969, Springer-Verlag, 1995, pp. 552–562. [249, 259]
91. , On the expressive completeness of the propositional mu-calculus with re-
spect to monadic second order logic, Proceedings of the 7th International Confer-
ence on Concurrency Theory, CONCUR ’96, Lecture Notes in Computer Science,
vol. 1119, Springer-Verlag, 1996, pp. 263–277. [239, 259]
92. Marcin Jurdziński, Deciding the winner in parity games is in UP ∩ co-UP, Infor-
mation Processing Letters 68 (1998), no. 3, 119–124. [96, 103, 104, 108, 110, 112,
115, 130]
93. , Small progress measures for solving parity games, Proceedings of the 17th
Annual Symposium on Theoretical Aspects of Computer Science, STACS 2000,
Lecture Notes in Computer Science, vol. 1770, Springer-Verlag, 2000, pp. 290–301.
[96, 103, 106, 108, 109, 117, 119, 124, 130, 151, 153, 202]
94. Charanjit S. Jutla, Determinization and memoryless winning strategies, Informa-
tion and Computation 133 (1997), no. 2, 117–134. [80]
95. Alexander S. Kechris, Classical descriptive set theory, Graduate Texts in Mathe-
matics, Springer-Verlag, 1995. [30]
96. Bakhadyr Khoussainov and Anil Nerode, Automata theory and its applications,
Progress in Computer Science and Applied Logic, vol. 21, Birkhäuser, 2001. [39]
97. Nils Klarlund, Progress measures for complementation of omega-automata with
applications to temporal logic, Proceedings of the 32nd Annual Symposium on
Foundations of Computer Science, FoCS ’91, IEEE Computer Society Press, 1991,
pp. 358–367. [61, 62, 63]
98. , Progress measures, immediate determinacy, and a subset construction for
tree automata, Annals of Pure and Applied Logic 69 (1994), no. 2–3, 243–268.
[39]
99. Nils Klarlund, Madhavan Mukund, and Milind A. Sohoni, Determinizing Büchi
asynchronous automata, Proceedings of the 15th Conference on Foundations of
Software Technology and Theoretical Computer Science, FSTTCS ’95, Lecture
Notes in Computer Science, no. 1026, Springer-Verlag, 1995, pp. 456–470. [44]
100. Dexter Kozen, Results on the propositional mu-calculus, Theoretical Computer
Science 27 (1983), 333–354. [95, 108, 171, 202, 203]
101. Orna Kupferman, P. Madhusudan, P. S. Thiagarajan, and Moshe Y. Vardi, Open
systems in reactive environments: Control and synthesis, Proceedings of the 11th
International Conference on Concurrency Theory, CONCUR 2000, Lecture Notes
in Computer Science, vol. 1877, Springer-Verlag, 2000, pp. 92–107. [40]
102. Orna Kupferman and Moshe Y. Vardi, Weak alternating automata are not that
weak, Proceedings of the Fifth Israel Symposium on Theory of Computing and
Systems, ISTCS ’97, IEEE Computer Society Press, 1997, pp. 147–158. [61, 77]
103. , Weak alternating automata and tree automata emptiness, Proceedings of
the 30th Annual ACM Symposium on Theory of Computing, STOC ’98, ACM
Press, 1998, pp. 224–233. [168]
Literature 371
104. , Chruch’s problem revisted, The Bulletin of Symbolic Logic 5 (1999), no. 2,
245–263. [40]
105. , The weakness of self-complementation, Proceedings of the 16th Annual
Symposium on Theoretical Aspects of Computer Science, STACS ’99, Lecture
Notes in Computer Science, vol. 1563, Springer-Verlag, 1999, pp. 455–466. [168]
106. , An automata-theoretic approach to reasoning about infinite-state systems,
Proceedings of the 12th International Conference on Computer Aided Verification,
CAV 2000), Lecture Notes in Computer Science, vol. 1855, Springer-Verlag, 2000.
[303, 312, 313, 316, 317, 318]
107. , Weak alternating automata are not that weak, ACM Transactions on
Computional Logic 2 (2001), no. 3, 408–429. [61, 77]
108. Ralf Küsters and Thomas Wilke, Determinizing Büchi asynchronous automata,
Proceedings of the 22th Conference on Foundations of Software Technology and
Theoretical Computer Science, FSTTCS ’02, Lecture Notes in Computer Science,
Springer-Verlag, 2002, To appear. [202]
109. Richard E. Ladner, The computational complexity of provability in systems of
propositinal modal logic, SIAM Journal on Computing 6 (1977), no. 3, 467–480.
[321]
110. Lawrence H. Landweber, Decision problems for ω-automata, Mathematical Sys-
tems Theory 3 (1969), no. 4, 376–384. [20, 39]
111. H. Läuchli and J. Leonard, On the elementary theory of linear order, Fundamenta
Mathematicae 59 (1966), 109–116. [228]
112. Giacomo Lenzi, A hierarchy theorem for the mu-calculus, Proceedings of the
23rd International Colloquium on Automata, Languages and Programming,
ICALP ’96, Lecture Notes in Computer Science, vol. 1099, Springer-Verlag, 1996,
pp. 87–97. [185, 202]
113. Matti Linna, On ω-sets associated with context-free languages, Information and
Control 31 (1976), no. 3, 272–293. [4]
114. Christof Löding, Optimal bounds for the transformation of omega-automata, Pro-
ceedings of the 19th Conference on Foundations of Software Technology and The-
oretical Computer Science, FSTTCS ’99, Lecture Notes in Computer Science, vol.
1738, Springer-Verlag, 1999, pp. 97–109. [16, 17, 18, 19, 39, 79, 87, 88, 89]
115. Christof Löding and Wolfgang Thomas, Alternating automata and logics over
infinite words, Proceedings of the IFIP International Conference on Theoretical
Computer Science, IFIP TCS 2000, Lecture Notes in Computer Science, vol. 1872,
Springer-Verlag, 2000. [61, 68]
116. David E. Long, Anca Browne, Edmund M. Clarke, Somesh Jha, and Wilfredo R.
Marrero, An improved algorithm for the evaluation of fixpoint expressions, Pro-
ceedings of the 6th International Conference on Computer Aided Verification,
CAV ’94, Lecture Notes in Computer Science, vol. 818, Springer-Verlag, 1994,
pp. 338–350. [202]
117. Walter Ludwig, A subexponential randomized algorithm for the simple stochastic
game problem, Information and Computation 117 (1995), no. 1, 151–155. [111,
125, 130]
118. P. Madhusudan and P.S. Thiagarajan, Distributed controller synthesis for local
specifications, Proceedings of the 28th International Colloquium on Automata,
Languages and Programming, ICALP ’01, Lecture Notes in Computer Science,
vol. 2076, Springer-Verlag, 2001, pp. 396–407. [40]
119. Donald A. Martin, Borel determinacy, Annals of Mathematics 102 (1975), 363–
371. [30, 95, 130]
372 Literature
120. Maarten Marx, Tolerance logic, Tech. Report IR-469, Faculteit der Exacte Weten-
schappen, Vrije Universiteit Amsterdam, The Netherlands, 1999. [356]
121. Yuri Matiyasevich, Diophantine nature of enumerable sets (Russian), Doklady
Akademija Nauk SSSR 191 (1970), no. 2, 279–282. [207]
122. Oliver Matz, Dot-depth and monadic quantifier alternation over pictures, Ph.D.
thesis, RWTH Aachen, Germany, 1999, Aachener Informatik Berichte 99-08. [231,
233, 258]
123. , Dot-depth, monadic quantifier alternation, and first-order closure over
grids and pictures, Theoretical Computer Science 270 (2002), no. 1–2, 1–70. [231,
233, 258]
124. Robert McNaughton, Finite-state infinite games, Tech. report, Project MAC,
Massachusetts Institute of Technology, USA, 1965. [39]
125. , Testing and generating infinite sequences by a finite automaton, Infor-
mation and Control 9 (1966), no. 5, 521–530. [39, 43, 61, 92]
126. , Infinite games played on finite graphs, Annals of Pure and Applies Logic
65 (1993), no. 2, 149–184. [39, 96, 108, 130]
127. A. R. Meyer, Weak monadic second order theory of successor is not elementary-
recursive, Proceedings of the Boston University Logic Colloquium, Springer-
Verlag, 1975, pp. 132–154. [258]
128. Max Michel, Complementation is more difficult with automata on infinite words,
Manuscript, CNET, Paris, 1988. [16, 39]
129. Satoru Miyano and Takeshi Hayashi, Alternating finite automata on ω-words,
Theoretical Computer Science 32 (1984), 321–330. [76]
130. Faron Moller and Alexander Rabinovich, On the expressive power of CTL*, Pro-
ceedings of the 14th IEEE Symposium on Logic in Computer Science, LICS ’99,
IEEE Computer Society Press, 1999, pp. 360–369. [259]
131. Andrzej Wlodzimierz Mostowski, Regular expressions for infinite trees and a stan-
dard form of automata, Computation Theory, Lecture Notes in Computer Science,
vol. 208, Springer-Verlag, 1984, pp. 157–168. [39, 130]
132. , Games with forbidden positions, Tech. Report 78, Instytut Matematyki,
Uniwersytet Gdański, Poland, 1991. [39, 95, 130, 354]
133. David E. Muller, Infinite sequences and finite machines, Proceedings of the 4th
IEEE Symposioum on Switching Circuit Theory and Logical Design, 1963, pp. 3–
16. [3, 39, 43, 92, 168]
134. David E. Muller, Ahmed Saoudi, and Paul E. Schupp, Alternating automata, the
weak monadic theory of the tree, and its complexity, Proceedings of the 13th In-
ternational Colloquium on Automata, Languages and Programming, ICALP ’86,
Lecture Notes in Computer Science, vol. 226, Springer-Verlag, 1986, pp. 275–283.
[168]
135. David E. Muller and Paul E. Schupp, The theory of ends, pushdown automata,
and second-order logic, Theoretical Computer Science 37 (1985), 51–75. [263,
318]
136. , Alternating automata on infinite trees, Theoretical Computer Science 54
(1987), 267–276. [68, 168]
137. , Simulating alternating tree automata by nondeterministic automata: New
results and new proofs of the theorems of Rabin, McNaughton and Safra, Theo-
retical Computer Science 141 (1995), no. 1–2, 69–107. [43, 92, 154, 162]
138. Damian Niwiński, On fixed-point clones (extended abstract), Proceedings of the
13th International Colloquium on Automata, Languages and Programming,
ICALP ’86, Lecture Notes in Computer Science, vol. 226, Springer-Verlag, 1986,
pp. 464–473. [176]
Literature 373
158. Shmuel Safra, On the complexity of omega-automata, Proceedings of the 29th An-
nual Symposium on Foundations of Computer Science, FoCS ’88, IEEE Computer
Society Press, 1988, pp. 319–327. [16, 43, 61, 92]
159. , Exponential determinization for omega-automata with strong-fairness ac-
ceptance condition (extended abstract), Proceedings of the 24th Annual ACM
Symposium on the Theory of Computing, STOC ’92, ACM Press, 1992, pp. 275–
282. [18, 79, 80, 86, 92]
160. Detlef Seese, The structure of the models of decidable monadic theories of graphs,
Annals of Pure and Applied Logic 53 (1991), no. 2, 169–195. [209]
161. Helmut Seidl, Fast and simple nested fixpoints, Information Processing Letters 59
(1996), no. 6, 303–308. [117, 130, 202]
162. Alexei L. Semenov, Decidability of monadic theories, Proceedings of the 11th
International Symposium on Mathematical Foundations of Computer Science,
MFCS ’84, Lecture Notes in Computer Science, vol. 176, Springer-Verlag, 1984,
pp. 162–175. [285]
163. Saharon Shelah, The monadic second order theory of order, Annals of Mathemat-
ics 102 (1975), 379–419. [258, 285, 337]
164. Ludwig Staiger, Research in the theory of ω-languages, Journal of Information
Processing Cybernetics EIK 23 (1987), 415–439. [39]
165. , ω-languages, Handbook of Formal Language Theory, vol. III, Springer-
Verlag, 1997, pp. 339–387. [4, 39]
166. Ludwig Staiger and Klaus W. Wagner, Automatentheoretische und Automaten-
freie Charakterisierungen Topologischer Klassen Regulärer Folgenmengen, Elek-
tronische Informationsverarbeitung und Kybernetik EIK 10 (1974), 379–392. [20]
167. Colin Stirling, Local model checking games, Proceedings of the 6th International
Conference on Concurrency Theory, CONCUR ’95, Lecture Notes in Computer
Science, vol. 962, Springer-Verlag, 1995, pp. 1–11. [130]
168. , Decidability of bisimulation equivalence for pushdown processes, Tech.
Report EDI-INF-RR-0005, School of Informatics, University of Edinburgh, Scot-
tland, 2000. [282, 283, 318]
169. , Modal and temporal properties of processes, Texts in Computer Science,
Springer-Verlag, 2001. [202]
170. Larry J. Stockmeyer, The complexity of decision problems in automata theory
and logic, Ph.D. thesis, Deptartment of Electrical Engineering, MIT, Boston,
Massachusetts, 1974. [231, 233, 258]
171. Robert S. Streett, Propositional dynamic logic of looping and converse is elemen-
tary decidable, Information and Control 54 (1982), no. 1–2, 121–141. [39, 79]
172. Robert S. Streett and E. Allen Emerson, An automata theoretic decision procedure
for the propositional mu-calculus, Information and Computation 81 (1989), no. 3,
249–264. [202]
173. Jonathan Stupp, The lattice-model is recursive in the original model., Tech. re-
port, Institute of Mathematics, The Hebrew University, Jerusalem, Israel, 1975.
[285, 318]
174. Alfred Tarski, A decision method for elementary algebra and geometry, Tech.
report, Rand Corporation, Santa Monica, California, 1948. [208, 226]
175. , A lattice-theoretical fixpoint theorem and its applications, Pacific Journal
of Mathematics 5 (1955), 285–309. [122, 359, 360]
176. James W. Thatcher and Jesse B. Wright, Generalized finite automata theory with
an application to a decision problem of second-order logic, Mathematical Systems
Theory 2 (1968), no. 1, 57–81. [219, 221, 258]
Literature 375
1 5
Σ∗ finite words over Σ 3 even(α) symbols occurring infinitely
Σω infinite words over Σ 3 often in even positions
|α|a number of a in α 3 90
Occ(α) symbols occurring in α 3 odd (α) symbols occurring infinitely
Inf(α) symbols occurring infinitely often in odd positions 90
often 4
REG class of regular languages 4
6
L(A) language accepted by A 5
G[U ] subgame of G induced by U
96
2
vE E-successors of v 23 7
σ player 23
σ the other player 23 || · ||∞ maximum norm 116
π play 24 <i restricted lexicographical
|w|a number of a in w 28 ordering 117
Attrσ (G, X) σ-attractor of X in G 35 [i] {0, . . . , i − 1} 117
MG range of progress measures
119
3 MG extended range of progress
measures 120
b (α) number of b in word α 44 prog progress function 120
||ρ|| winning vertices 120
ordering on progress
4
measures 122
α(i) i-th letter of α 62 Lift lift operator 122
Occ(α) letters occurring in α 62 wπ vertex with highest priority
Inf(α) letters occurring infinitely occurring infinitely often
often in α 62 126
R(v) vertices reachable from v Pπ vertices before wπ 126
63 l(x) length of path to wπ 126
set of marked vertices 63 φ progress ordering 126
C
||G|| width of a sliced graph 64 ≺ value ordering 127
U (G) unmarked boundary 64
µ progress measure 64 8
B+ (X) positive Boolean formulas
68 Tω infinite binary tree 136
Mod(θ) minimal models 68 P re< (π) prefixes of π 136
θ dual of θ 69 TΣω , TΣ Σ-labeled trees 136
A dual of A 70 pi projection 136
GA,α weak min-parity game 70 T (A) tree language recognized
Attrσ (X) attractor 72 by A 137
378 Symbol Index
9 ϕA formula equivalent to A
214
P propositional variables 154
Aϕ automaton equivalent to ϕ
sR successors of s 154
214
Rs predecessors of s 154
O class of orderings 228
TCQ transition conditions 155
Q2 universal states 155
Q3 existential states 155 13
−
→q successor of q 155 ϕA
n defines 0∗ 10f (n)−1 10∗ 231
[v] the last letter of v 159 Σk S
{$k , 0k , 1k } 233
Θ, Θp set of tiles 164
S
k−1
Σ<k i=1 Σi 233
n
Σ>k i=k+1 Σi 233
10 ϕk cyclical counter 234
G ↓v subgame 171
[p
→ S ] substitution 172 14
G(A) transition graph of A 172 sr source of a transition system
ind(A) index of A 172 239
Fµ set of µ-formulas 173 sccM
r (s) r-successors of s 240
Fν set of ν-formulas 173 ∼ bisimulation 240
Fη Fµ ∪ Fν 173 c
M ω-unravelling of M 241
ψ≤ϕ ψ is a subformula of ϕ 174 extension relation 241
G(ϕ) graph of ϕ 176 C MSO (ϕ) class defined by a MSO-
SCCϕ (ψ) strongly connected compo- sentence 244
nent of G(ϕ) containing ψ C Lµ (ϕ) class defined by a Lµ -
176 sentence 245
α(ϕ) alternation depth of ϕ 176 L(A) language recognized by A
ψ state for ψ 177 248
11 15
Σnµ , Πnµ fixed point hierarchy of Lµ REG(Σ ∗ ) regular subset of Σ ∗ 264
188 TN tree over N 264
ϕ models of ϕ 188 VG nodes of G 265
edge from s to t in G 265
a
, Πnµ TR semantical hierarchy 188
Σnµ TR s−→t
Σn , Π n automata hierarchy 189
G
A class accepted by A 189
s=⇒t
a
G path from s to t in G 265
ΣnTR , ΠnTR semantical automata
L path in L from s to t 265
s−→t
hierarchy 189 L(G, s, t) language of G wrt. s and t
TR class of pointed transition 265
systems 189 L(G) language of G 265
TΣn Σn -test automaton 195 W (U→V ) sets of edges 265
TΠn Πn -test automaton 195 MTh(G) monadic theory of G 266
GA,q transformed transition ≤MSO MSO-interpretation 266
system 196 I (B) interpretation 266
Σ inverse alphabet 272
Σ̂ extended alphabet 272
12 u↓ normal form of u 272
|=W weak satisfaction relation h(b) extended substitution 272
210 Σ̂# marked extended alphabet
≺ prefix order 211 274
Symbol Index 379
17 20
∞ finite or infinite words over class of ordinals 359
T On
Σ 304 cl(f ) closure ordinal 360
T, l a Σ-labeled W -tree 304 f! f cl(f ) 360
ext(W ) alphabet W ∪· {3, ↑} 304 LFP least fixed point 360
↑ “go to parent node” 304 GFP greatest fixed point 360
Index
complementation, 44, 68, 74, 88, 89, 141, – – Jurdziński’s algorithm, 123
144, 146, 162 – – simple algorithm, 114
complemented pair condition, 9 – – strategy improvement algorithm, 125
concatenable, 164 – – subexponential algorithm, 111
configuration, 311 – – UP-algorithm, 117
conform, 27 – positional, 32
coordinate map, 266 – Rabin chain, 25
counter, 232 – reachability, 34
CTL∗ , 257 – regular, 25
– simple stochastic, 111
dead end, 24 – solitaire, 117
decomposition tree, 346 – subgame, 96
dense shuffling, 228 – weak parity, 70
determinacy, 30, 33, 99 game graph, 142
discounted payoff game, 110 game position, 70, 142
domain, 264 game progress measure, 120
dual automaton, 70 GF, see guarded fragment
dual positive Boolean formula, 69 GF∞ , 326
graph, 63
even cycle, 117 – context-free, 318
extended infinite binary tree, 211 – finitely marked, 64
extended parity progress measure, 120 – left concatenation, 277
extended substitution, 272 – prefix-recognizable, 263, 267
– pushdown, 263
fairness condition, 9 – regular, 263
finite memory determinacy, 33 – right concatenation, 277
finite model property, 229, 343 – Σ-graph, 264
finitely marked graph, 64 – sliced, 64
first-order logic, 226, 233, 236, 238 greatest fixed point, 173, 359
fixed point, 359 guarded bisimulation, 327
– greatest, 173, 359 guarded fragment, 323
– least, 173, 359 guarded second-order logic, 209, 332
– simultaneous, 363 guarded set, 323
fixed point formulas, 173
fixed point operators, 173 hit position, 12
FO, see first-order logic
forgetful determinacy, 33 IAR, see index appearance record
index, 172, 186
Gaifman graph, 324 index appearance records, 86
game, 24 inductive function, 359
– 1-game, 25, 36 infinite binary tree, 136, 211
– Büchi, 25, 37 infinity set, 79, 136
– discounted payoff, 110 inflated, 159
– initialized, 25, 310 inflationary function, 359
– µ-game, 248 initialized game, 25
– mean payoff, 109, 110 initially closed, 212
– memoryless, 32 instance, 155
– Muller, 25, 28 interpretation, 299
– parity, 25, 28, 141, 143, 310 inverse alphabet, 272
– – finite, 103, 106 inverse substitution, 272
Index 383