The Complexity of Network Satisfaction Problems for Symmetric Relation Algebras with a Flexible Atom
Abstract
Robin Hirsch posed in 1996 the Really Big Complexity Problem: classify the computational complexity of the network satisfaction problem for all finite relation algebras . We provide a complete classification for the case that is symmetric and has a flexible atom; in this case, the problem is NP-complete or in P. The classification task can be reduced to the case where is integral. If a finite integral relation algebra has a flexible atom, then it has a normal representation . We can then study the computational complexity of the network satisfaction problem of using the universal-algebraic approach, via an analysis of the polymorphisms of . We also use a Ramsey-type result of Nešetřil and Rödl and a complexity dichotomy result of Bulatov for conservative finite-domain constraint satisfaction problems.
1 Introduction
One of the earliest approaches to formalise constraint satisfaction problems over infinite domains is based on relation algebras [LM94, Hir97]. We think about the elements of a relation algebra as binary relations; the algebra has operations for intersection, union, complement, converse, and composition of relations, and constants for the empty relation, the full relation, and equality, and is required to satisfy certain axioms. Important examples of relation algebras are the Point Algebra, the Left Linear Point Algebra, Allen’s Interval Algebra, RCC5, and RCC8, just to name a few.
The most important computational problems related to relation algebras that have been studied are called network satisfaction problems (NSPs). A finite network is given in which the nodes are linked via the elements of a fixed relation algebra. Then the computational task of the NSP for this relation algebra is to decide whether the network satisfies a strong notion of consistency with respect to the laws of the fixed relation algebra. Such NSPs for a finite relation algebra can be used to model many computational problems in temporal and spatial reasoning [Dün05, RN07, BJ17].
More than two decades ago, [Hir96] asked the Really Big Complexity Problem (RBCP): can we classify the computational complexity of the network satisfaction problem for every finite relation algebra? For example, the complexity of the network satisfaction problem for the Point Algebra and the Left Linear Point Algebra is in P [VKvB89, BK07], while it is NP-complete for all of the other examples mentioned above [All83, RN99]. There also exist relation algebras where the complexity of the network satisfaction problem is not in NP: Hirsch gave an example of a finite relation algebra with an undecidable network satisfaction problem [Hir99]. This result might be surprising at first sight; it is related to the fact that the representation of a finite relation algebra by concrete binary relations over some set can be quite complicated. We also mention that not every finite relation algebra has a representation [Lyn50]. There are even non-representable relation algebras that are symmetric [Mad06a]; a relation algebra is symmetric if every element is its own converse.
A simple condition that implies that a finite relation algebra has a representation is the existence of a so-called flexible atom [Com84, Mad82]. A flexible atom is an element of that is maximally unconstrained in its interaction with other elements of the relation algebra; the formal definitions can be found in Section 2.
Such relation algebras have been studied intensively, for example in the context of the so-called flexible atoms conjecture [Mad94, AMM08]. We will see that integral relation algebras with a flexible atom even have a normal representation, i.e., a representation which is fully universal, square, and homogeneous [Hir96]. The network satisfaction problem for a relation algebra with a normal representation can be seen as a constraint satisfaction problem for an infinite structure that is well-behaved from a model-theoretic point of view; in particular, we may choose to be homogeneous and finitely bounded.
Constraint satisfaction problems over finite domains have been studied intensively in the past two decades, and tremendous progress has been made concerning systematic findings about their computational complexity. As a highlighting result, [Bul17] and [Zhu17, Zhu20] proved the famous Feder-Vardi dichotomy conjecture which states that every finite-domain CSP is in P or NP-complete. Both proofs build on an important connection between the computational complexity of constraint satisfaction problems and universal algebra.
The universal-algebraic approach can also be applied to study the computational complexity of countably infinite homogeneous structures with finite relational signature [BN06]. For an introduction to the field, we refer to [Bod21]. If is finitely bounded, then CSP is contained in NP (see, e.g., [Bod12]). If is homogeneous and finitely bounded then a complexity dichotomy has been conjectured, along with algebraic criteria that distinguish NP-complete from polynomial-time solvable problems [BPP21]. The exact formulation of the conjecture from [BPP21] in full generality requires concepts that we do not need to prove our results. In Theorem 9.1 we verify these conjectures for all normal representations of finite integral symmetric relation algebras with a flexible atom, and thereby also solve Hirsch’s RBCP for symmetric relation algebras with a flexible atom. Phrased in the terminology of relation algebras, our result is the following.
Theorem 1.1.
Let be a finite symmetric representable relation algebra with a flexible atom, and let be the set of atoms of . Then
-
there exists an operation that preserves the allowed triples of , satisfies
and satisfies the Siggers identity
in this case the network satisfaction problem for is in P, or
-
the network satisfaction problem for is NP-complete.
Moreover, the satisfiability of the Siggers identity in Theorem 1.1 is a decidable criterion for that is a sufficient condition for the polynomial-time tractability of the network satisfaction problem of . We want to mention that there are several other equivalent criteria that could be used instead of the first item in the theorem, namely all characterizations of Taylor Algebras for finite conservative algebras (see, e.g., [Bul03]).
This article is a postprint version of an article that appeared in the Journal of Artificial Intelligence Research [BK22]; a conference version of the article appeared under the title “Network satisfaction for symmetric relation algebras with a flexible atom” in [BK21].
1.1 Proof Strategy
Every finite integral representable relation algebra with a flexible atom has a normal representation ; for completeness, and since we are not aware of a reference for this fact, we include a proof in Section 3. It follows that the classification question about the complexity of the network satisfaction problem of can be translated into a question about the complexity of the constraint satisfaction problem for the relational structure .
We then associate a finite relational structure to and show that can be reduced to in polynomial-time (Section 4). If the structure satisfies the condition of the first case in Theorem 1.1, then known results about finite-domain s imply that is in P [Bul03, Bul16, Bar11], and hence is in P, too. If the first case in Theorem 1.1 does not apply, then known results about finite-domain algebras imply that there are such that the canonical polymorphisms of act as a projection on [Bul03, Bul16, Bar11]. We first observe NP-hardness of if does not have a binary injective polymorphism (Section 6). If has a binary injective polymorphism, we use results from structural Ramsey theory to show that must even have a binary injective polymorphism which is canonical (Section 7). This implies that none of equals . We then prove that does not have a binary -symmetric polymorphism; also in this step, we apply Ramsey theory. In Section 8 we show that this in turn implies that all polymorphisms of must be canonical on . Finally, we show that cannot have a polymorphism which acts as a majority or as a minority on , and thus by Schaefer’s theorem all polymorphisms of act as a projection on . This is again implied by results from Section 7. Finally it follows that is NP-hard. This concludes the proof of Theorem 1.1.
1.2 Organisation of the Article
Section 2 introduces all the basic concepts and tools that are used in this article. In Section 3 we define flexible atoms and obtain first results about representable relation algebras with a flexible atom. Section 4 is dedicated to the atom structure and the polynomial-time tractability results. In Section 5 we provide an additional perspective on the class of computational problems under consideration; in this section we define those problems completely without the use of the relation algebra framework. The reader can get better intuition of the class of problems studied in this article, however our results and proofs do not rely on that section. The Sections 6-8 contain the main parts of the proof as outlined in the previous paragraph. In Section 9 we put everything together and prove the main theorem. We end with a conclusion and a small discussion of our result.
2 Preliminaries
We recall some basic definitions and results about universal algebra, representable relation algebras, constraint satisfaction, model theory, and structural Ramsey theory.
2.1 Algebras and Structures
Let denote the natural numbers starting with and define . A signature is a set of relation symbols and function symbols. Each symbol is associated with a natural number, called the arity of the symbol. Function symbols of arity are called constant symbols. A -structure is a tuple where is a set, called the domain of , such that for every :
-
if is a relation symbol of arity , then is a subset of ,
-
if is a function symbol of arity , then is an operation .
Note that by a subset of can be seen as a Boolean value and the operation can be interpreted as a constant. As long as there is no risk of confusion we will often use the function symbols for the corresponding functions, and the relation symbols for the corresponding relations, i.e., we use instead of .
Let and be two -structures. A homomorphism h from to is a function such that
-
for every relation symbol of arity and every tuple , we have that ;
-
for every function symbol of arity and every tuple , we have that .
In the case that contains only function symbols and is surjective, then is called homomorphic image of . In general, the homomorphism is called an embedding if is injective and satisfies
-
for every relation symbol of arity and every tuple , we have that .
A surjective embedding is called an isomorphism. An endomorphism of a -structure is a homomorphism from to and an automorphism of is an isomorphism from to . We denote by the group of all automorphisms of .
A -structure is a substructure of a -structure if
-
;
-
for every relation symbol of arity and every tuple , we have that if and only if ;
-
for every function symbol of arity and every tuple , we have that .
Note that for every subset of the domain of a -structure there exists a unique substructure of with smallest domain and . We call this the substructure of induced by .
Let and be structures. The direct product is the -structure where
-
is the domain of ;
-
for every relation symbol of arity and every tuple , we have that if and only if and ;
-
for every function symbol of arity and every tuple , we have that .
We denote the direct product by . The -fold product is defined analogously and denoted by .
Structures with a signature that only contains function symbols are called algebras and structure with purely relational signature are called relational structures. Since we do not deal with signatures of mixed type in this article, we will from now on use the term structure for relational structures only.
2.2 Representable Relation Algebras and Network Satisfaction Problems
Relation algebras that are not representable have a trivial network satisfaction problem, namely the class of yes-instances of their network satisfaction problem is empty (see Definition 2.5). We thus omit the definition of relation algebras and start immediately with the simpler definition of representable relation algebras; here we basically follow the textbook by [Mad06b].
Definition 2.1.
Let be a set and an equivalence relation on .
Let be an algebra
with the following operations:
-
1.
,
-
2.
,
-
3.
,
-
4.
,
-
5.
,
-
6.
,
-
7.
.
A subalgebra of is called a proper relation algebra.
The class of representable relation algebras, denoted by , consists of all algebras with signature with corresponding arities and that are isomorphic to some proper relation algebra. We use bold letters (such as ) to denote algebras from and the corresponding roman letter (such as ) to denote the domain of the algebra. An algebra is called finite if its domain is finite. We call symmetric if all its elements are symmetric, i.e., for every .
According to the previous definition, an algebra is in if it has an isomorphism to a proper relation algebra. Such an isomorphism is usually called the representation of . To link the theory of relation algebras with model theory it will be convenient to view representations of algebras in as relational structures and to not use the classical notation here. However, it is easy to see that the existence of a representation is equivalent under both definitions.
Definition 2.2.
Let . Then a representation of is a relational structure such that
-
is an -structure, i.e., the elements of are binary relation symbols of ;
-
The map is an isomorphism between and the proper relation algebra induced by the relations of in .
Let be a representable relation algebra and let us define a new operation on the set . Then the algebra is by definition a Boolean algebra and induces therefore a partial order on , which is defined by . Note that for proper relation algebras, this ordering coincides with the set-inclusion order. The minimal elements of this order in are called atoms. The set of atoms of is denoted by . A tuple is called an allowed triple if . Otherwise, is called a forbidden triple and in this case . We say that a relational -structure induces a forbidden triple (from ) if there exists and such that and hold and is a forbidden triple. Note that a representation of does not induce a forbidden triple.
Definition 2.3.
Let . An -network is a finite set together with a partial function , where is the domain of . An -network is satisfiable in a representation of if there exists an assignment such that for all the following holds:
An -network is satisfiable if there exists a representation of such that is satisfiable in .
We give an example of an instance of the for a representable relation algebra. The numbering of the representable relation algebra is by [AM94].
Example 2.4 (An instance of of representable relation algebra ).
Let be the symmetric representable relation algebra with the set of atoms and the values for the composition operation on these atoms be given by Table 1. Note that this determines the composition operation on the whole domain of , which is the following set:
Let be a set. Consider the map given by
The tuple is an example of an instance of of .
We will in the following assume that for an -network it holds that . Otherwise, is not satisfiable. Note that every -network can be viewed as an -structure on the domain : for all and the relation holds if and only if .
Definition 2.5.
The (general) network satisfaction problem for a finite representable relation algebra , denoted by , is the problem of deciding whether a given -network is satisfiable.
2.3 Normal Representations and CSPs
In this section we consider a subclass of introduced by Hirsch in 1996. For representable relation algebras from this class, corresponds naturally to a constraint satisfaction problem (CSP). In the last two decades a rich and fruitful theory emerged to analyse the computational complexity of CSPs. We use this theory to obtain results about the computational complexity of NSPs.
In the following let be in . An -network is called closed (transitively closed in the work by [Hir97]) if it is total and for all it holds that , , and . It is called atomic if the range of only contains atoms from .
Definition 2.6 (from [Hir96]).
Let be a representation of . Then is called
-
fully universal, if every atomic closed -network is satisfiable in ;
-
square, if ;
-
homogeneous, if for every isomorphism between finite substructures of there exists an automorphism of that extends this isomorphism;
-
normal, if it is fully universal, square and homogeneous.
Definition 2.7.
Let be a relational signature. A first-order formula is called primitive positive (pp) if it has the form
where are atomic formulas, i.e., formulas of the form for and , of the form for , or of the form false.
As usual, formulas without free variables are called sentences. A relational structure is called connected if it is not the disjoint union of two structures. A connected component of a relational structure is a substructure of that is maximal with respect to domain inclusion and is connected. For every primitive positive -sentence with variable set the canonical database is defined as the -structure on where is in the relation for if and only if is a conjunct from . Conversely every relational -structure induces a primitive positive -sentence on the variable set , the so-called conjunctive query, simply by taking the conjunction all atomic formulas that hold in . We say that the primitve positive -sentences are the connected components of a primitive positive -sentence if they are the conjunctive queries of the connected components of the canonical database .
Definition 2.8.
Let be a finite relational signature and let be a -structure. Then the constraint satisfaction problem of is the computational problem of deciding whether a given primitive positive -sentence holds in .
Consider the following translation which associates to each -network a primitive positive -sentences as follows: the variables of are the elements of and contains for every in the domain of the conjunct if and only if holds. For the other direction let be an -sentence with variable set and consider the -network with the following definition: for every , if does not appear in any conjunct from we leave undefined, otherwise let all the conjuncts from that contain . We compute in the element and define .
The following theorem, which subsumes the connection between network satisfaction problems and constraint satisfaction problems is based on this natural 1-to-1 correspondence between -networks and -sentences.
Theorem 2.9 ([Bod12], see also [BJ17, Bod18]).
Let be finite. Then the following holds:
-
1.
has a representation such that and are the same problem up to the translation between -networks and -sentences.
-
2.
If has a normal representation the problems and are the same up to the translation between -networks and -sentences.
2.4 Model Theory
Let be a finite relational signature. The class of finite -structures that embed into a -structure is called the age of , denoted by . If is a class of finite -structures, then is the class of all finite -structures such that no structure from embeds into . A class of finite structures is called finitely bounded if there exists a finite set of finite -structures such that (see, e.g., [Mac11]). The structures from are called bounds or forbidden substructures. It is easy to see that a class of -structures is finitely bounded if and only if it is axiomatisable by a universal -sentence. A structure is called finitely bounded if is finitely bounded.
Proposition 2.10 ([Bod18]).
Let be a normal representation of a finite . Then the following holds:
-
is finitely bounded by bounds of size at most three.
-
The -reduct of is finitely bounded by bounds of size at most three.
Definition 2.11.
A class of finite -structures has the amalgamation property if for all structures with embeddings and there exist a structure and embeddings and such that . If additionally , then we say that has the strong amalgamation property.
Let be -structures. Then is the -structure on the domain such that for every . If Definition 2.11 holds with then we say that has the free amalgamation property; note that the free amalgamation property implies the strong amalgamation property.
Theorem 2.12 (Fraïssé; see, e.g., [Hod97]).
Let be a finite relational signature and let be a class of finite -structures that is closed under taking induced substructures and isomorphisms and has the amalgamation property. Then there exists an up to isomorphism unique countable homogeneous structure such that .
Definition 2.13.
Let be a relational structure. A set is called an -orbit of if is preserved by all and for all there exists such that .
For a structure the automophism group is called transitive if has only one orbit. We want to remark that if is a normal representation of a finite then the 2-orbits of are exactly the relations induced by the atoms of .
2.5 The Universal-Algebraic Approach
In this section we present basic notions for the so-called universal-algebraic approach to the study of s.
Definition 2.14.
Let be some set. We denote by the set of all -ary operations on and by the set of all operations on . A set is called an operation clone on if it contains all projections of all arities and if it is closed under composition, i.e., for all and it holds that , where is the -ary function defined as follows
An operation is called conservative if for all it holds that A clone is called conservative if all its operations are conservative. Note that an operation clone on can be considered as an algebra with domain and an infinite signature that consists of symbols for all operations in (cf. Section 2.1). In this sense a conservative operation clone on induces on every set a subalgebra. It is easy to see that this subalgebra corresponds to an operation clone on . We call this the restriction of to . We later need the following classical result for clones over a two-element set.
Theorem 2.15 ([Pos41]).
Let be a conservative operation clone on . Then either contains only projections, or at least one of the following operations:
-
1.
the binary function ,
-
2.
the binary function ,
-
3.
the minority function,
-
4.
the majority function.
Operation clones occur naturally as polymorphism clones of relational structures. If and , then we write for the -tuple obtained by applying component-wise to the tuples .
Definition 2.16.
Let a structure with a finite relational signature and let . An -ary operation preserves a relation if for all it holds that
If preserves all relations from then is called a polymorphism of .
In order to provide an additional view on polymorphisms we give the following definition.
Definition 2.17.
Let be a relational signature and let be a -structure. For the structure is the -structure on the domain defined as follows. Let be of arity . Then
It is easy to see that the -ary polymorphisms of are precicely the homomorphisms from to .
The set of all polymorphisms (of all arities) of a relational structure is an operation clone on , which is denoted by . A Siggers operation is an operation that satisfies the Siggers identity (see Theorem 1.1). The following result can be obtained by combining known results from the literature.
Theorem 2.18 ([Sig10, Bul03]; see also [Bar11, Bul16]).
Let be a finite structure with a finite relational signature such that is conservative. Then precisely one of the following holds:
-
1.
There exist distinct such that for every the restriction of to is a projection. In this case, is NP-complete.
-
2.
contains a Siggers operation; in this case, is in P.
We now discuss fundamental results about the universal-algebraic approach for constraint satisfaction problems of structures with an infinite domain.
Theorem 2.19 ([BN06]).
Let be a homogeneous structure with finite relational signature. Then a relation is preserved by if and only if it is primitively positively definable in .
The following definition is a preparation to formulate the next theorem which is a well-known condition that implies NP-hardness of for homogeneous structures with a finite relational signature.
Definition 2.20.
Let be a class of algebras. Then we have
-
is the class of homomorphic images of algebras from and
-
is the class of subalgebras of algebras from .
-
is the class of finite products of algebras from .
An operation clone on a set can also be seen as an algebra with domain whose signature consists of the operations of such that for all .
The following is a classical condition for NP-Hardness, see for example Theorem 10 in the survey by [Bod08].
Theorem 2.21.
Let be a homogeneous structure with finite relational signature. If contains a 2-element algebra where all operations are projections, then is NP-hard.
In the following let be finite and with normal representation .
Definition 2.22.
Let be atoms of . Then the -ary relation is defined as follows:
An operation is called edge-conservative if it satisfies for all and all
Note that for every the structure contains the relation . Therefore the next proposition follows immediately since polymorphisms of preserve all relations of .
Proposition 2.23.
All polymorphisms of are edge-conservative.
Definition 2.24.
Let . An operation is called -canonical (with respect to ) if there exists a function such that for all and , if for all then An operation is called canonical (with respect to ) if it is -canonical. In this case we say that the behaviour is total. If we call a partial behaviour. The function is called the behaviour of on . If then is just called the behaviour of .
We denote by the set of all polymorphisms of that are canonical with respect to . It will always be clear from the context what the domain of a behaviour is. An operation is called symmetric if for all it holds that . An -canonical function is called -symmetric if the behaviour of on is symmetric.
2.6 Ramsey Theory and Canonisation
We avoid giving an introduction to Ramsey theory, since the only usage of the Ramsey property is via Theorem 2.26, and rather refer to the survey by [Bod15] for an introduction.
Let be a homogeneous -structure such that has the strong amalgamation property. Then the class of all -structures such that is a linear order and whose -reduct (i.e. the structure on the same domain, but only with the relations that are denoted by symbols from , see e.g. the book by [Hod97]) is from is a strong amalgamation class, too (see e.g. [Bod15]). By Theorem 2.12 there exists an up to isomorphism unique countable homogeneous structure of that age, which we denote by . It can be shown by a straightforward back-and-forth argument that is isomorphic to an expansion of , so we identify the domain of and of along this isomorphism, and call the expansion of by a generic linear order.
Theorem 2.25 ([NR89, HN16]).
Let be a relational -structure such that has the free amalgamation property. Then the expansion of by a generic linear order has the Ramsey property.
The following theorem gives a connection of the Ramsey property with the existence of canonical functions and plays a key role in our analysis.
Theorem 2.26 ([BP21]).
Let be a countable homogeneous structure with finite relational signature and the Ramsey property. Let be an operation and let
Then there exists a canonical operation such that for every finite there exists such that .
Remark 2.27.
Let and be homogeneous structures with finite relational signatures. If and have the same domain and the same automorphism group, then has the Ramsey property if and only if has it (see, e.g., [Bod15]).
3 Representable Relation Algebras with a Flexible Atom
In this section we define the concept of a flexible atom and show how to reduce the classification problem for the network satisfaction problem for a finite with a flexible atom to the situation where is additionally integral (Proposition 3.3). A finite representable relation algebra is called integral if the element is an atom of (cf. [Mad06b]).
Then we show that an integral with a flexible atom has a normal representation. Therefore, the universal-algebraic approach is applicable; in particular, we make heavy use of polymorphisms and their connection to primitive positive definability in later sections (cf. Theorem 2.19). Furthermore, we prove that every normal representation of a finite representable relation algebra with a flexible atom has a Ramsey expansion (Section 3.2). Therefore, the tools from Section 2.6 can be applied, too. Finally we give some examples of representable relation algebras with a flexible atom (Section 3.3). We start with the definition of a flexible atom.
Definition 3.1.
Let and let . An atom is called flexible if for all it holds that .
This definition can for example be found in the book by [HH02, Chapter 11, Exercise 1]. Note that this definition does not require the representable relation algebra to be integral. This is slightly more general than the definition by [Mad94, Mad06b]. As mentioned before, we show in the following section that it is sufficient for our result to classify the computational complexity of s for finite representable relation algebras with a flexible atom that are additionally integral. This means that readers who prefer this second definition by [Mad94, Mad06b] (assuming integrality) can perfectly skip the following section and read the article with this other definition in mind. In this case representable relation algebras with a flexible atom are always implicitly integral.
3.1 Integral Representable Relation Algebras
Let and let . The atoms in are called identity atoms. Therefore, is integral if and only if has exactly one identity atom. The first claim of the following lemma is a well-known fact about atoms of representable relation algebras. This fact is used to prove a second claim about representable relation algebras with a flexible atom. In simple terms, this claim states that a representation of a non-integral representable relation algebra with a flexible atom is only “square” on precisely those elements that are in one certain identity atom.
Lemma 3.2.
Let be finite. Then there exists for every atom a unique with such that . Furthermore, if is a flexible atom then for all with and we have that .
Proof.
Note that by definition and therefore for all with . Since is an atom either or . By and there exists at least one with . In the next step we prove uniqueness of such an element . Assume for contradiction that there exist distinct with and such that and . Note that since and are identity atoms. Therefore, we get
which is a contradiction since is an atom. This proves the first statement.
For the second statement assume for contradiction that there exists such that and . Let be an atom with . Since we get . Since is a flexible atom it holds that and therefore
which is a contradiction. ∎
Proposition 3.3.
Let be finite and with a flexible atom. Then there exists a finite integral with a flexible atom such that the following statements hold:
-
1.
There exists a polynomial-time many-one reduction from to .
-
2.
There exists a polynomial-time many-one reduction from to .
-
3.
The atom structure of has a polymorphism that satisfies the Siggers identity if and only if the atom structure of has such a polymorphism (see Definition 4.1).
Proof.
If is integral there is nothing to be shown. So assume that is not integral and let be a flexible atom. Let be a representation of such that and are the same problem up to the translation between -networks and -sentences. Such a representation exists by Theorem 2.9. Let and let be the unique element with and that exists by Lemma 3.2. The second statement of Lemma 3.2 implies and therefore we have that and . Let be the substructure of on the domain . The set of relations of clearly induces a proper relation algebra which is integral. We denote this representable relation algebra by . Note that we can also consider as a subset of . Let be the representation of such that and are the same problem up to the translation between -networks and -sentences. As before, such a representation exists by Theorem 2.9.
Proof of 1.: Note that if a connected instance of is satisfiable, then either all variables are mapped to an atom from the subset of that corresponds to or all variables are mapped to one element with and . This leads to the following polynomial-time many-one reduction from to , which proves together with Theorem 2.9 the claim that there exists a polynomial-time many-one reduction from to . Consider the following algorithm: For a given primitive positive -sentence it computes the connected components (see paragraph after Definition 2.7). Then it checks for every whether there exists such that for every conjunct of it holds that . Let be the set of indices for which this is not the case. Then the algorithm defines new -sentences from for every by replacing every conjunct with where . Let be the conjunction of all for .
It is easy to see that is computable from in polynomial time and that is a satisfiable instance of if and only if is a satisfiable instance of .
Proof of 2.: Consider now an -network . We claim that is satisfiable as an -network if and only if it is satisfiable as an -network. Suppose that is satisfiable in a representation of by an assignment . Let be fresh elements for every atom with . We build the disjoint union of with one-element -structures and then close the structure under union and intersection of binary relations. This results in a representation of that satisfies again by the assignment . For the other direction, if is satisfiable in a representation of we can again consider the substructure on the domain and get a representation of that satisfies .
Proof of 3.: Let be a polymorphism of the atom structure of that satisfies the Siggers identity. By assumption satisfies
and therefore the restriction of to is a polymorphism of the atom structure of .
For the other direction choose an arbitrary ordering of the atoms . If is a Siggers polymorphism of the atom structure of one can extend to an operation by defining
It is easy to see that this operation satisfies also the Siggers identity. Furthermore, since every atom from is only contained in allowed triples of the form it follows that preserves the allowed triples from (see after Definition 4.1).
∎
3.2 Normal Representations
Let be for the rest of the section finite, integral, and with a flexible atom . We consider the following subset of :
Let be an -network and let be the corresponding -structure (see paragraph before Definition 2.5). Let be the -structure on the same domain as such that for all and we have
We call the -free companion of an -network .
The next lemma follows directly from the definitions of flexible atoms and -free companions.
Lemma 3.4.
Let be the class of -free companions of atomic closed -networks. Then has the free amalgamation property.
Proof.
Let and be structures in with embeddings and . Since is a flexible atom and is a class of -free companions we get that the structure is in . Therefore, the natural embeddings and prove the free amalgamation property of . ∎
As a consequence of this lemma we obtain the following.
Proposition 3.5.
has a normal representation .
Proof.
Let be the class from Lemma 3.4. This class is closed under taking substructures and isomorphisms. By Lemma 3.4 it also has the amalgamation property and therefore we get by Theorem 2.12 a homogeneous structure with . Let be the expansion of by the following relation
Let be the (homogeneous) expansion of by all Boolean combinations of relations from . Then is a representation of . Since is the class of all atomic closed -networks, is fully universal. The definition of witnesses that is a square representation of : for all elements there exists an atom such that holds. ∎
The next theorem is another consequence of Lemma 3.4.
Theorem 3.6.
Let be a normal representation of . Let be the expansion of by a generic linear order. Then has the Ramsey property.
Proof.
Let be the -reduct of . The age of this structure has the free amalgamation property by Lemma 3.4. Therefore, Theorem 2.25 implies that the expansion of by a generic linear order has the Ramsey property. By Remark 2.27 the structure also has the Ramsey property since and have the same automorphism group. ∎
Remark 3.7.
The binary first-order definable relations of form a proper relation algebra since has quantifier-elimination (see [Hod97]). By the definition of the generic order the atoms of this proper relation algebra are of the following form
-
for , or
-
for , or
-
,
where the relation consists of all tuples such that holds.
3.3 Examples
We give two concrete examples of finite integral symmetric representable relation algebras with a flexible atom (Examples 3.8 and 3.9), and a systematic way of building such algebras from arbitrary representable relation algebras (Example 3.10). The numbering of the algebras in the examples is by [AM94].
Example 3.8 (Representable relation algebra ).
The representable relation algebra has three atoms, namely the identity atom and two symmetric atoms and . The multiplication table for the atoms is given in Fig. 2. In this representable relation algebra the atoms and are flexible. Consider the countable, homogeneous, undirected graph , whose age is the class of all finite undirected graphs (see, e.g., [Hod97]), also called the Random graph. The expansion of by all binary first-order definable relations is a normal representation of the algebra . In this representation the atoms and are interpreted as the relation and the relation , where is defined as .
Example 3.9 (Representable relation algebra ).
The representable relation algebra also consists of three symmetric atoms. The multiplication table in Fig. 2 shows that in this algebra the element is a flexible atom. To see that is not a flexible atom, note that . Let be the countable, homogeneous, undirected graph, whose age is the class of all finite undirected graphs that do not embed the complete graph on three vertices (see, e.g., [Hod97]). This structure is called a Henson graph. If we expand by all binary first-order definable relations we get a normal representation of the algebra . To see this note that we interpret as the relation . That is triangle free, i.e. triangles of are forbidden, matches with the fact that holds in the representable relation algebra.
Example 3.10.
Consider an arbitrary finite, integral . Clearly does not have a flexible atom in general. Nevertheless we can expand the domain of to implement an “artificial” flexible atom.
Let be some symbol not contained in . Let us mention that every element in can uniquely be written as a union of atoms from . Let be the set of all subsets of . The set is the domain of our new algebra . Note that on there exists the subset-ordering and is closed under set-union and complement (in ) We define to be symmetric and therefore get the following unary function ∗ in as follows. For an element we define
The new function symbol in is defined on the atoms as follows:
One can check that is a finite integral representable relation algebra with a flexible atom . Note that the forbidden triples of are exactly those of together with triples which are permutations of for some .
4 Polynomial-time Tractability
In this section we introduce for every finite an associated finite structure, called the atom structure of . Note that it is closely related, but not the same, as the type structure introduced by [BM16]. In the context of relation algebras the atom structure has the advantage that its domain is the set of atoms of , rather than the set of 3-types, which would be the domain of the type structure of [BM16]; hence, our domain is smaller and has some advantages on which the main result of this section (Proposition 4.4) is based. Up to a some differences in the signature, our atom structure is the same as the atom structure introduced by [Lyn50] which was used there for different purposes (see also [Mad82, HH01, HJK19]).
Let be a normal representation of a finite . We will reduce to the of the atom structure of . This means that if the of the atom structure is in P, then so are and . For our main result we will show later that every network satisfaction problem for a finite integral symmetric representable relation algebra with a flexible atom that cannot be solved in polynomial time by this method is NP-complete.
Definition 4.1.
The atom structure of is the finite relational structure with domain and the following relations:
-
for every the unary relation ,
-
the binary relation ,
-
the ternary relation .
Note that the relation consists of the allowed triples of . We say that an operation preserves the allowed triples if it preserves the relation .
Proposition 4.2.
Let be a fully universal representation of a finite . There is a polynomial-time reduction from to .
Proof.
Let be an instance of with variable set . We construct an instance of as follows. The variable set of is given by . The constraints of are of the two kinds:
-
1.
Let be an element of and let be an atomic formula of . If , then we add the atomic (unary) formula to ; otherwise we add the atomic formula . If we additionally add .
-
2.
Let be such that . Then we add the atomic formula
to .
It remains to show that this reduction is correct. Let be a satisfying assignment for . This assignment maps every pair of variables and to a unique atom in and therefore induces a map . The map clearly satisfies all atomic formulas introduced by (1.). To see that it also satisfies all formulas introduced by (2.) note that maps the elements to a substructure of , which does not induces a forbidden triple.
For the other direction assume that is a satisfying assignment for . This induces an -structure on (maybe with some identification of variables) as follows: we add to the relation if and ; if otherwise and we add to the relation . It is clear that no forbidden triple from is induced by . Also note that satisfies by the choice of the (unary) constraints of the first kind. Since is a fully universal representation the structure is a substructure of . Hence, the instance is satisfiable in . ∎
The atom structure has another property which is fundamental for our proof of Theorem 1. Recall that every canonical polymorphism induces a behaviour . In the next proposition we show that then is a polymorphism of . Moreover the other direction also holds. Every is the behaviour of a canonical polymorphism of .
Proposition 4.3.
Let be a normal representation of a finite .
-
1.
Let be canonical and let be its behaviour. Then .
-
2.
Let . Then there exists a canonical whose behaviour equals .
Proof.
For (1): Let be canonical and let . Then by the definition of there exist tuples such that for all we have
We apply the canonical polymorphism and get . Then there exists an allowed triple such that
We have that and by the definition of the behaviour of a canonical function we get . The other relations in are preserved trivially and therefore .
For (2): Since is fully universal and homogeneous it follows by a compactness argument (see e.g., Lemma 2 by [BD13]) that every countable -structure which does not induce a forbidden triple and is square has a homomorphism to . It is therefore enough to show that every operation with behaviour does not induce a forbidden triple in the image. Let be such that the application of a canonical function with behaviour on would give a tuple with such that
Since preserves the triple is not forbidden. ∎
Recall from Proposition 2.23 that polymorphisms of are edge-conservative. Note that this implies that polymorphisms of are conservative. In fact, Theorem 2.18 and the Proposition 4.3 imply the following.
Proposition 4.4.
If contains a canonical polymorphism whose behaviour is a Siggers operation in then is in P.
We demonstrate how this result can be used to prove polynomial-time tractability of for a symmetric, integral with a flexible atom.
Example 4.5 (Polynomial-time tractability the NSP of representable relation algebra ).
The polynomial-time tractability of the NSP of the representable relation algebra (see Example 3.8) was first shown by [CH04] (see also Section 8.4 of [BP15]). Here we consider the following function .
Let be the normal representation of the algebra given in Example 3.8. Note that is the behaviour of an injective, canonical polymorphism of . The injectivity follows from the last line of the definition; if then . Therefore preserves all allowed triples, since in the algebra the only forbidden triples involve . One can check that is a Siggers operation and therefore we get by Proposition 4.4 that is in P.
Example 4.6.
Consider the construction of representable relation algebras with a flexible atom from Example 3.10. It is easy to see that for a finite integral has a polynomial-time reduction to where is the representable relation algebra with a flexible atom that is constructed in Example 3.10. We get as a consequence that if a normal representation of satisfies the condition of Proposition 4.4 then is in P.
Corollary 4.7.
If does not have a Siggers operation then there exist elements such that the restriction of every operation from to is a projection.
5 Network Consistency Problems
The purpose of this section is to give an additional perspective on the class of network satisfaction problems of finite symmetric integral representable relation algebras with a flexible atom. Even more, we define these computational problems in this section completely without the use of the relation algebra framework. Our classification result for these problems does not depend on the content of this section and the reader may skip it.
We introduce a class of computational decision problems which we call network consistency problems (NCPs). It is easy to see that NCPs are in a 1-to-1 correspondence with NSPs of finite, symmetric, integral with a flexible atom.
Definition 5.1.
Let be a finite set and . Then is called totally symmetric if for all bijections we have
We call an element identity element if for all the following holds:
A structure is called a stencil if is totally symmetric and it contains an identity element.
Definition 5.2.
Let be an undirected graph and let be a set. We call a map an edge -coloring of if for all with it holds that .
For each fixed stencil, we define an NCP as follows.
Definition 5.3.
Let be a stencil. The network completion problem of , denoted by , is the following problem. Given a finite undirected graph with an edge -coloring the task is to decide whether there exists an edge -coloring of such that
-
1.
for all with it holds that .
-
2.
for all with we have
The following proposition illustrates how NCPs correspond to a certain class of NSPs.
Proposition 5.4.
The class of NCPs and the class of NSPs for finite symmetric integral representable relation algebras with a flexible atom are in a natural 1-to-1 correspondence such that corresponding problems are polynomial-time equivalent.
Proof.
Let be a stencil with according to (2) in Definition 5.1 and let be new element with . We define a relational structure as follows. The domain of is the set . We assume that has every subset of its domain as a unary relation. Furthermore contains the binary relation and the ternary relation which is defined as follows:
One can find a finite symmetric integral algebra with domain and a flexible atom such that is the atom structure of (see, e.g., Theorem 2.2 and Theorem 2.6 in the article by [Mad82]).
Furthermore, given a finite symmetric integral algebra with a flexible atom let be the atom structure of . Let be the substructure of the -reduct of induced by . By the properties of we get that is a stencil.
We show that the instances of and are in a natural 1-to-1 correspondence that preserves the acceptance condition of the computational problems. Let be a finite undirected graph with an edge -coloring . We define an -network by defining
It is easy to see that is an accepted instance of if and only if is an accepted instance of . Since we can reverse this and find for every -network a finite undirected graph with an edge -coloring such that each of them is an accepted instance if and only if the other one is. These to reductions show that the computational decision problems and are polynomial-time equivalent.
∎
We end this section by providing a rich source of examples for s.
Example 5.5 (“Distance problems”).
Let be an arbitrary finite set that contains . We define the relation of all tuples which satisfy all instantiations of the triangle inequality, i.e.
where the addition is meant to be the usual one on rational numbers. The relation is by definition totally symmetric and the element is an identity element. Therefore, is a stencil.
Now consider a finite undirected graph with an edge -coloring . This can be seen as a labeling of each edge in the graph by a set of possible (or allowed) distances. The computational task of is to decide whether one can choose for each edge one of the possible distance such that in the end this choice satisfies on each triangle of edges the triangle inequalities of metric spaces.
By Proposition 5.4 there exists a finite symmetric integral algebra with a flexible atom such that and are polynomial-time equivalent. By the proof of Proposition 5.4 we have that the domain of is equal to . It is easy to observe that is the set of atoms from .
We define an operation as follows:
where the operation is the usual from in .
The allowed triples of are, up to triples that involve the flexible atom , those which arise from valid triangle inequalities. For this reason the operation preserves the allowed triples of . Moreover, one can check that satisfies the Siggers identity. This implies that all these “distance problems” satisfy the first condition in Theorem 1.1 and are therefore solvable by a polynomial-time algorithm. Furthermore, if the normal representation of induces a proper relation algebra on the set , then is a representable relation algebra with a normal representation. This follows from Proposition 2.7.4 in the thesis of [Con15] (see also [DLPS07]), since the composition operation of is associative. We get that if the representable relation algebra exists then the argument from Example 3.10 implies that is also in P.
6 Binary Injective Polymorphisms
We give in this section a proof of the following proposition.
Proposition 6.1.
Let be a normal representation of a finite, symmetric, integral with a flexible atom . If does not contain a 2-element algebra where all operations are projections, then has a binary injective polymorphism.
This statement is a consequence of well known results that can be found in the book by [Bod21] and results by [MP20] applied to the class of normal representations of finite, symmetric, integral with a flexible atom . An operation is called essentially unary if it depends on at most one of its variables and is called essential otherwise.
Following the terminology of [MP20], we now define free 2-orbits. The existence of a free 2-orbit appeared under the name ‘orbital extension property’ for example in the habilitation thesis of [Bod12].
Definition 6.2.
Let be a structure. A 2-orbit of is called free if for all elements there exists with and .
Note that if has a free 2-orbit then it is transitive. The following theorem generalises a fact that was first proved for first-order reducts of by [BK09].
Proposition 6.3 (Lemma 5.3.10 in [Bod12]).
Let be a structure such that has a free 2-orbit. If contains an essential operation then it contains a binary essential operation.
The following is essentially taken from the article by [MP20].
Definition 6.4.
Let be a structure. Then the canonical binary structure of is the structure with domain and a binary relation for each 2-orbit of such that implies .
Definition 6.5.
A -structure has finite duality if there exists a finite set of finite -structures such that a -structure has a homomorphism to if and only if no element of has a homomorphism to .
We establish finite duality for the class of structures which is important for our classification purposes.
Lemma 6.6.
Let be a normal representation of a finite, integral with a flexible atom . Then the canonical binary structure of has finite duality.
Proof.
Let . Note that since is integral and is homogeneous, the canonical binary structure of is precisely the -reduct of . Let be the set of all -structures with domain that do not have a homomorphism to . We show that the set witnesses the finite duality of the canonical binary structure of . Let be a -structure with a homomorphism to . If there exists with a homomorphism to , then also has a homomorphism to , contradicting the choice of . For the other direction assume that no element from has a homomorphism to . Let be the -expansion of the -reduct of where the relation is defined by
By the definition of the flexible atom it follows that no element from has a homomorphism to . This implies that for all distinct elements from the tuple is in at most one relation from . The definition of ensures that is in at least one relation from . Recall that Proposition 2.10 states that is finitely bounded by -structures of size at most three. Assume that one of those bounds embeds into . This implies by what we noted before that all elements of are in precisely one relation from . On the other hand is not in and therefore has a homomorphism to . Since all elements of are related by precisely one relation from , this homomorphism needs to be a embedding, contradicting our assumption on to be a bound. Therefore, none of the bounds of embeds into , which means that is a substructure of . Clearly, there exists a homomorphism from to which proves the lemma. ∎
The following proposition about the existence of injective operations is from [MP20], building on ideas of [BP15] and [BMPP19].
Proposition 6.7 ([MP20]).
Let be a homogeneous structure such that is transitive and such that the canonical binary structure of has finite duality. If contains a binary essential operation that preserves then it contains a binary injective operation.
We are now able to prove the main result of this section.
Proof of Proposition 6.1.
Note that since is integral and is homogeneous, the flexible atom is a free 2-orbit of . Furthermore, is transitive. Suppose that does not contain a 2-element algebra where all operations are projections. Since all operations of are edge conservative, it follows that contains an operation that does not behave as a projection on . This implies that contains an essential operation. By Proposition 6.3, must also contain a binary essential operation. Since the canonical binary structure of has finite duality by Lemma 6.6 we can apply Proposition 6.7 and get that contains a binary injective operation. ∎
The following shows how to use Proposition 6.1 to obtain a hardness result for the concrete from Example 3.9.
Example 6.8 (Hardness of representable relation algebra , see [BMPP19, BK20]).
Let be the normal representation of the algebra mentioned in Example 3.9. We claim that the structure does not have a binary injective polymorphism. To see this, consider a substructure of on elements such that , , and . Assume has a binary injective polymorphism . This means that holds. Then we get that , , and hold in , which is a contradiction, since in triangles of this form are forbidden. By the contraposition of Proposition 6.1 it follows that contains a 2-element algebra where all operations are projections. We conclude with Theorem 2.21 that is an NP-hard problem.
7 From Partial to Total Canonical Behaviour
In this section we prove that in many cases the existence of a polymorphism with a certain partial behaviour implies the existence of a canonical polymorphism with the same partial behaviour. Following this idea we start in Section 7.1 with the proof that the existence of an injective polymorphism implies the existence of a canonical injective polymorphism. In some cases the existence of an -canonical polymorphism implies the existence of a canonical polymorphism with the same behaviour on . We prove this separately for binary (Section 7.2) and ternary (Section 7.3) operations, making use of the binary injective polymorphism that exists by the results from Section 6 and Section 7.1.
Let us remark that most proofs of this section would fail if the representable relation algebra was not symmetric. Indeed, every representable relation algebra that contains a non-symmetric atom a normal representation would not satisfy Proposition 7.2 below, since the stated behaviour is not well-defined.
We assume for this section that is a normal representation of a finite, symmetric, integral with a flexible atom . Let furthermore be the expansion of by the generic linear order. The structure exists by the observations in Section 3.2.
Proposition 7.1.
Let be injective. Then there exists a polymorphism of and an injective endomorphism of such that
as mappings from to .
Proof.
Let and consider the substructure induced by on . There exists a linear ordering on , namely the lexicographic order given by the linear order of on each coordinate.
Let be the expansion of by the linear order that is induced by the lexicographic linear order of on the preimage. This is well defined since is injective. By the definition of and a compactness argument the structure embeds into . In this way we obtain a homomorphism from to . Again by a compactness argument also an endomorphism with the desired properties exists. ∎
7.1 Canonical Binary Injective Polymorphisms
We prove in this section that the existence of an injective polymorphism implies the existence of a canonical injective polymorphism. We say that a polymorphism of is canonical with respect to if satisfies Definition 2.24, where the underlying representable relation algebra is the proper relation algebra induced by the binary first-order definable relations (i.e., unions of 2-orbits) of . Note that in a normal representation the set of -orbits equals the set of the interpretations of the atoms of the representable relation algebra.
Proposition 7.2.
Let be a binary polymorphism of that is canonical with respect to . Let be the map such that for all
(cf. Remark 3.7). Then is well defined and there exists a canonical binary polymorphism of with behaviour .
Proof.
The function is well defined since all atoms are symmetric. We show that there exists a canonical polymorphism of that has as a behaviour. Consider the following structure on the domain . Let and let be atoms of with and . Then we define that holds if and only if .
We show in the following that has a homomorphism to . This is enough to prove the statement, because a homomorphism from to is a canonical polymorphism of . Since is homogeneous is suffices to show that every finite substructure of homomorphically maps to . Let be a finite substructure of and assume for contradiction that does not homomorphically map to . We can view as an atomic -network. Since is fully universal is not closed. There must exist elements of and atoms such that holds in and
This means that the substructure induced on the elements by contains a forbidden triple.
Now we consider the substructures that are induced on and by . Our goal is to order these elements such that for all
(1) |
If we achieve this we know that there exist elements in that induce isomorphic copies of the induced structures of the elements and with the additional ordering. Now the application of the polymorphism on these elements results in a structure whose -reduct is isomorphic to the substructure induced by and on by the definition of the canonical behaviour . This contradicts our assumption because a polymorphism can not have a forbidden substructure in its image.
It remains to show that we can choose orderings on the elements and such that (1) holds. Without loss of generality we can assume that holds. Now consider the following cases:
-
1.
and .
We can obviously choose linear orders on both sets such that (1) holds.
-
2.
and .
Assume that holds then the possible orders are
-
3.
and .
First consider the case that and hold. Then we choose as orders
In the second possible case we can assume without loss of generality that and hold. Note that otherwise we could change the role of two of the tuples and and get this case. The compatible order is then
-
4.
and .
In this case we choose the order -
5.
and .
Assume that holds and that we have -
6.
and .
For this case we trivially get
Note that up to the symmetry of the arguments for both coordinates these are all the possible cases. This completes the proof of the proposition. ∎
Corollary 7.3.
Suppose that has a binary injective polymorphism. Then also has a canonical binary injective polymorphism.
Proof.
By Proposition 7.1 we may assume that there exists also an injective polymorphism of . The structure has the Ramsey property by Theorem 3.6. Therefore, Theorem 2.26 implies that there also exists an injective canonical polymorphism of . According to Proposition 7.2 the restriction of the behaviour to the 2-orbits that satisfy induces the behaviour of a canonical polymorphism of which is also injective. ∎
7.2 Canonical -symmetric Polymorphisms
We will now use the results about binary injective polymorphism from Section 7.1 to show the existence of a canonical -symmetric polymorphism in case there exists an -symmetric polymorphism.
Lemma 7.4.
Let be atoms. Then every binary -symmetric polymorphism of is injective.
Proof.
Let be an -symmetric polymorphism. Without loss of generality . Assume for contradiction that is not injective. This means that there exist and with (for the notation see Definition 2.22) such that holds.
Case 1: . Since is a flexible atom we may choose such that and hold. By the choice of the polymorphism we get and which induces either the forbidden triple or the forbidden triple on , and .
Case 2: . We choose such that and . This is possible since is the flexible atom. We obtain and which again induces a forbidden triple on , and .
Case 3: . We choose such that and . This is possible since is the flexible atom. We obtain and which again induces a forbidden triple on , and .
Since we obtained in all cases a contradiction we conclude that is injective.∎
Proposition 7.5.
Let be atoms. If has a binary -symmetric polymorphism, then has also a binary canonical -symmetric polymorphism.
Proof.
Let be the binary -symmetric polymorphism. By Lemma 7.4 we know that is injective. By Proposition 7.1 it induces a polymorphism on . The structure has the Ramsey property by Theorem 3.6. Let be the canonization of that exists by Theorem 2.26. The restriction of the behaviour to the 2-orbits that satisfy induces by Proposition 7.2 the behaviour of a canonical polymorphism of . The way we obtained ensures that is -symmetric with the same behaviour on as . ∎
The following is an easy observation about -symmetric polymorphisms that we will use several times.
Observation 7.6.
Let be an atom and an -symmetric polymorphism of . Then .
Proof.
Suppose for contradiction that . Let be such that
and consider the substructure of that is induced by and . This structure induces a forbidden triple which contradicts the assumption ∎
In the remainder of the section we combine canonical -symmetric polymorphisms to obtain a single “maximal-symmetric” polymorphism.
Definition 7.7.
We call a subset an edge of and we call the elements in
the red edges of .
The terminology of the colored edges as well as the following lemma are from [Bul16].
Lemma 7.8.
There exists a binary canonical polymorphism that is symmetric on all red edges and behaves on each non-red edge like a projection. We call this function maximal-symmetric.
Proof.
For each let be a canonical polymorphism such that its behaviour is symmetric on . We prove the lemma by an induction on the size of subsets of , i.e., we show that for every subset of size there exists a polymorphism that is symmetric on all edges from . For each subset of of size one, there exists by the definition of red edges a canonical polymorphism with a behaviour that is symmetric on . Let and suppose there exists a canonical polymorphism with symmetric behaviour on elements from . Let . We want to show that there exists a canonical polymorphism with a behaviour that is symmetric on all elements from . We may assume that this does not hold for , otherwise we are done. Therefore, and since is edge-conservative, we have
With this it is easy to see that is a polymorphism with a behaviour that is symmetric on all elements from .
This proves the first part of the statement. For the second part note that for a binary canonical edge-conservative polymorphism there are only 4 possibilities for the behaviour on a set . If is not a red edge then every binary canonical edge-conservative polymorphism behaves like a projection on . ∎
7.3 Canonical Ternary Polymorphisms
We obtain in this section a result that states that the existence of a ternary -canonical polymorphism implies the existence of a canonical polymorphism with the same behaviour on as (Corollary 7.11). This is done similarly as in Section 7.2.
Lemma 7.9.
Let be a binary, injective, maximal-symmetric polymorphism. Then the function where is defined by
is a homomorphism. Moreover, for all with it holds that
Note that this means that two distinct tuples in the image of have distinct entries in each coordinate.
Proof.
Let and suppose that holds for . By the definition of the product structure holds for all . Since is a polymorphism clearly holds by the definition of . Now we use again the definition of a product structure and get which shows that is a homomorphism.
For the second part of the statement let distinct. Suppose that , and hold for some , where at least one atom is different from . Since is injective we have
Again by the injectivity of we get
By the definition of this shows that holds. It is easy to see by analogous arguments that the same is true for the other coordinates. Therefore, the statement follows. ∎
Proposition 7.10.
Let and be atoms of such that . Let be ternary -canonical and be injective, maximal-symmetric. Then there exists a canonical with the same behaviour as on .
Proof.
By Lemma 7.8 we may assume that behaves on like the projection to the first coordinate since . Let be the function defined in Lemma 7.9 and consider the function which is defined by .
Claim 1: is injective. Let be two distinct elements. By Lemma 7.9 we know that holds. Since is a polymorphism of we directly get that holds, which proves the injectivity of .
Claim 2: is -canonical and behaves on like .
Let with such . Since behaves like the first projection on it follows that . Together with the -canonicity of this proves Claim 2.
Since is injective there exists by Proposition 7.1 a polymorphism of . Since is a Ramsey structure we can apply Theorem 2.26 to . Let be the resulting polymorphism that is canonical with respect to . Note that if we consider as a polymorphism of it behaves on like and therefore like . Now we consider the induced behaviour of on all 2-orbits that satisfy . Since all atoms of are symmetric and is conservative this induces a function .
Claim 3: The partial behaviour does not induce a forbidden triple.
We would like to note that we prove this claim with similar arguments as in the proof of Proposition 7.2. Assume for contradiction that there exist such that the application of an operation with behaviour would induce a forbidden triple.Without loss of generality we can order the elements of each coordinate of strictly with for . Note that if on some coordinate there would be the relation then we are out of the domain of the behaviour .
If we choose such an order we can find isomorphic copies of this structure (with the order) in . If we apply the polymorphism to this copy and forget the order of the structure we get a structure that is by definition isomorphic to the forbidden triple.This proves Claim 3.
To finish the proof of the lemma note that the composition of with the projection to the -th coordinate for is a canonical, injective polymorphism (for injectivity see Lemma 7.9) and therefore induces a behaviour . We define by . The composition is the behaviour of a canonical function of . If would induce a forbidden triple then also would induce a forbidden triple, which contradicts Claim 3. ∎
Corollary 7.11.
Let have a binary injective polymorphism. Let be such that no -symmetric polymorphism exists. Let be a ternary -canonical polymorphism. Then there exists a canonical polymorphism with the same behaviour on as .
Proof.
By assumption there is no canonical -symmetric polymorphism and therefore . By Corollary 7.3 has a canonical binary injective polymorphism. This polymorphism is a witness that for all . With Lemma 7.8 we get a maximal symmetric polymorphism . Since we get that is -symmetric for all . By Observation 7.6 it follows , which implies that is injective. Now Proposition 7.10 implies the statement. ∎
8 The Independence Lemma and How To Use It
The central result of this section is Proposition 8.2 which states that the absence of an -symmetric polymorphism implies that all polymorphism are canonical on . The main ingredients of our proof of this proposition are the fact that has a flexible atom and the following “Independence Lemma” (Lemma 8.1).
8.1 The Independence Lemma
The following lemma transfers the absence of a special partially canonical polymorphism to the existence of certain relations of arity that are primitively positively definable in . A lemma of a similar type appeared as Lemma 42 in an article by [BP14].
Lemma 8.1 (Independence Lemma).
Let be a homogeneous structure with finite relational signature. Let and be 2-orbits of such that , , and are primitively positively definable in . Then the following are equivalent:
-
1.
has an -canonical polymorphism that is -symmetric with .
-
2.
For every primitive positive formula such that and are satisfiable over , the formula is also satisfiable over .
-
3.
For every finite there exists a homomorphism from the substructure of induced by to that is -canonical with .
Proof.
The implication from (1) to (2) follows directly by applying the symmetric polymorphisms to tuples from the relation defined by .
For the implication from (2) to (3) let be a finite subset of . Let with be an enumeration of . To construct consider the formula with variables for that is the conjunction of all atomic formulas such that and hold in . Note that this formula states exactly which relations hold on in . Let be the set of pairs such that
and | |||
and | |||
and |
If we show that the formula
is satisfiable by an assignment , we get the desired homomorphism by setting . We prove the satisfiability of by induction over the size of subsets of . For the inductive beginning consider an element . Without loss of generality we have that holds. Therefore the assignment witnesses the satisfiability of the formula . For the inductive step let be of size and assume that the statement is true for subsets of size . Let and be two elements from . We define the following formula
Then by the inductive assumption the formulas and are satisfiable. The assumptions on the elements in give us that also
and
are satisfiable; since is a primitive positive definable relation we are done otherwise. But then we can apply the assumption of (2) and get that also is satisfiable, which proves the inductive step.
The direction from (3) to (1) is a standard application of König’s tree lemma. For a reference see for example Lemma 42 in the article by [BP14]. ∎
8.2 Absence of -symmetric Polymorphisms
We are now able to prove the main result of this section, which will be a corner stone in the proof of Theorem 1.1. Our proof of this proposition makes use of a -ary relation with the following first-order definition:
It is an easy observation that the -canonical polymorphisms of are precisely those that preserve the relation . By Theorem 2.19 we get that whenever is primitively positively definable in then all polymorphisms of preserve and are therefore -canonical. In the following proof we use the -ary relations that are provided by the second item of the Independence Lemma 8.1 to provide a primitive positive definition of .
Proposition 8.2.
Let be a normal representation of a finite integral symmetric relation algebra with a flexible atom . Suppose that has a binary injective polymorphism. Let and be two atoms such that has no -symmetric polymorphism. Then all polymorphisms are canonical on .
Proof.
By Corollary 7.3 there exists a canonical binary injective polymorphism of . Therefore, for every the edge is red and the maximal symmetric polymorphism that exists by Lemma 7.8 is symmetric on all these edges. Observation 7.6 implies that is injective. Note that behaves like a projection on since there exists no -symmetric polymorphism.
Let be the formula defined as follows:
We use to formulate and prove the following claim:
Claim 1: a) There exists a formula such that
b) There exists a formula that has the same property with and in exchanged roles.
Proof of Claim 1. There exists a formula that witnesses the negation of (2) in the Independence Lemma (Lemma 8.1) since does not have an -symmetric polymorphism with . Let be the formula that witnesses in the same way the non-existence of an -symmetric polymorphism of with . We define for the formula as follows:
(2) |
If and witness a) and b) in Claim 1, we are done. So suppose that they do not. Note that if we have and such that
(3) | |||
(4) |
then would satisfy the statement about in Claim 1, a) or in Claim 1, b). To see this note that we can apply the injective, maximal symmetric polymorphism that behaves like a projection on to the tuples and that witness (3) and (4). The first tuple satisfies , and since and . The second tuple satisfies . Then the injectivity of ensures that the tuples and witness Claim 1, a) or Claim 1, b). Figure 3 illustrates this situation.
By our assumption that Claim 1 is not satisfied by and we conclude that for at least one it holds that for
(5) | |||
(6) |
We distinguish the following different cases.
- 1.
- 2.
- 3.
Case 1: Consider the following formula with
We claim that satisfies b) in Claim 1. This proves Claim 1, because satisfies a) in Claim 1. To see that fulfills the two satisfiability statements in Claim 1, b) we can first amalgamate the structure induced (as a substructure of ) by the elements of a satisfying tuple for with the structure induced by the elements of a satisfying tuple for (see Figure 4 for an illustration). We amalgamate these two structures over their common substructure induced by the variables and , with the variable names from the definition of . In this amalgamation step all missing edges are filled with the flexible atom . In a second amalgamation step we amalgamate the resulting structure with another copy of the structure , but now with the common substructure on the variables and (again with refer to the names used in the definition of ). As before the missing edges are filled with the flexible atom . Figure 4 illustrates the situation. It follows from the choice of and and the definition of that is not satisfiable in .
Case 2: This case is analogous to Case 1.
Case 3: Consider the formula with
We show that is satisfiable in . Since (5) holds for and and since and are distinct, there exists such that
is satisfiable in . Similarly, since (5) holds for and , there exists such that
and |
Note that there are such that the following atomic formulas hold:
If we choose for the existentially quantified variable in the definition of the element then the tuple satisfies the formula
By an analogous argument also is satisfiable. It follows again from the choice of and and the definition of that is not satisfiable in . By an analogous definition we can find a formula that satisfies b) in Claim 1. Therefore, we are done with Case 3. Altogether this proves Claim 1.
Let and be the two formulas that exist by Claim 1. We define the following formulas
We also define a formula as follows (see also Figure 5):
Analogously to Case 1, an amalgam of the structures that are induced by tuples that satisfy and shows that the formulas and are satisfiable in . Note that this is possible since we ensured in Claim 1 that there exist tuples that additionally satisfy . It also holds that a tuple that satisfies also satisfies
(7) |
Assume that for a tuple that satisfies it holds that . Then there exist such that holds. But this is by the definition of only possible if and hold, in contradiction to . The same argument works for proving that holds.
We complete the proof with a primitive positive definition of . We have the following primitive positive formula
and define by
For the forward direction of this equivalence, let be a tuple from such that holds for . Let and let and be two elements from such that and for every and every holds. Such elements exists since in the substructure of that is induced by all appearing triangles are allowed by the definition of the flexible atom . The elements and witness that satisfies the formula .
For the other direction assume that a tuple satisfies . Then there exist and such that is satisfied. Since we observed that the tuples and both satisfy (7) we can assume that holds for . It follows also from (7) that holds for and then (7) implies that holds, which proves the backward direction of the stated equivalence.
∎
9 Proof of the Result
In this section we prove the main results of this article. We first obtain a dichotomy theorem for a class of CSPs (Theorem 9.1). This is used in combination with the observations in Section 3 to conclude the proof of Theorem 1.1.
Theorem 9.1.
Let be finite integral symmetric and with a flexible atom and let be the set of atoms of . Then either
-
there exists an operation that preserves the allowed triples of , satisfies
and satisfies the Siggers identity
in this case, is in P, or
-
contains a 2-element algebra where all operations are projections; in this case, is NP-complete.
Proof.
Let be a normal representation of that exists by Proposition 3.5. The finite boundedness of implies that is in NP. Let be the atom structure of . We can assume that has a binary injective polymorphism, because otherwise Proposition 6.1 would directly imply the second item. The existence of a binary injective polymorphism implies by Corollary 7.3 the existence of a canonical binary injective polymorphism .
If the first item of the theorem is satisfied then the operation is a Siggers polymorphism of and the statement follows by the Propositions 4.3 and 4.4.
Assume therefore that the first item in the theorem does not hold. By Corollary 4.7 there exist elements such that the subalgebra of on contains only projections. It holds that , since is a witness that can not be in the domain of a subalgebra that contains only projections. Since all operations from are projections on there exists no canonical polymorphism of that is -symmetric. By Proposition 7.5 there exists also no -symmetric polymorphism of . Since has a binary injective polymorphism we can apply Proposition 8.2 and get that all polymorphisms of are -canonical. The last step is to show that all polymorphisms of behave like projections on .
Assume for contradiction that there exists a ternary, -canonical polymorphism that behaves on like a majority or like a minority. By Corollary 7.11 there exists a canonical polymorphism that is also a majority or minority on (here we use again the existence of an injective polymorphism). This contradicts our assumption that is trivial on . We get that every polymorphism of does not behave on as an operation from Post’s theorem (Theorem 2.15) and therefore must behave as a projection on by Theorem 2.15. Thus, contains a -element algebra whose operations are projections and is NP-hard, according to Theorem 2.21. ∎
We can prove the main result.
Proof of Theorem 1.1.
Let be as in the assumptions of the theorem and let be the finite symmetric integral representable relation algebra that exists by Proposition 3.3. Suppose that satisfies the first condition of Theorem 1.1. By Item 3) in Proposition 3.3, we get that then satisfies the first condition in Theorem 9.1 and therefore for the normal representation of is in P. By Section 2.3 we know that and are polynomial-time equivalent. This, together with the Turing reduction from to by Item 1) in Proposition 3.3 implies that is in P. This proves the first part of the theorem.
Assume that does not satisfy the first condition of Theorem 1.1. Item 3) in Proposition 3.3 again implies that does not satisfy the first condition in Theorem 9.1 and therefore for the normal representation of is NP-complete. As before we get that is NP-complete and the many-one reduction from to by Item 2) in Proposition 3.3 implies the NP-hardness of . The containment in NP follows by Item 1) in Proposition 3.3. This concludes the proof of Theorem 1.1. ∎
Corollary 9.2.
For a given finite, symmetric with a flexible atom it is decidable which of the two items in Theorem 1.1 holds. In particular, it is decidable whether is solvable in polynomial time.
Proof.
Since is a finite set one can go through all possible operations that preserve all the allowed triples of and check whether the Siggers identity is satisfied. If , it follows from Theorem 1.1 that it is decidable whether is in P. In the case of this is a trivial task. ∎
The problem of deciding whether certain identities hold in the polymorphism clone of a structure is well known problem in the study of CSPs. The computational complexity is known to be in NP [CL17]. The precise complexity of deciding whether a polymorphism clone of an explicitly given structure has a conservative operation that satisfies the Siggers identity is open [CL17].
10 Connection to Smooth Approximations
We discuss the relationship of our results to the techniques developed by [MP20]. The main invention of [MP20] are smooth approximations which are equivalence relations on sets of -tuples. The purpose of these equivalence relations is to approximate the prominent orbit-equivalence relation. We have seen the importance of these relations in the present article: the polymorphisms which preserve the orbit-equivalence relation are precisely the canonical ones and they store the information about possible finite-domain algorithms that can be used to solve the infinite-domain CSP (cf. Section 4).
We start by rearranging the results of Section 8 such that we get the following theorem. In order to repeat the key steps of our main proof with a focus on the similarities to [MP20] the assumptions in this theorem are a natural starting point.
Theorem 10.1.
Let be a normal representation of a finite integral symmetric relation algebra with a flexible atom and let be the atom structure of . Suppose that contains a binary injective polymorphism and does not have a Siggers operation. Then there exists two atoms and such that one of the following holds:
-
1.
The orbit-equivalence relation is primitively positively definable in .
-
2.
For all and every primitive positive formula such that and are satisfiable over , the formula is also satisfiable over .
This theorem is similar to the “Loop lemma of approximations” (Theorem 10 in [MP20]), even though it does not make proper use of the approximation idea ( approximates itself). The first case in the theorem leads to the hardness condition that contains a -element algebra whose operations are projections and therefore is NP-complete.
However, the second case is “stronger” than the second case in Theorem 10 by [MP20]. The strength in the statement relies on the special class of problems in our article. To see what we mean by this, we continue with the strategy of [MP20]. As a next step we can restrict the “Independence Lemma” (Lemma 8.1) as follows:
Lemma 10.2.
Let be a homogeneous structure with finite relational signature. Let and be 2-orbits of such that , , and are primitively positively definable in .
Assume that for every primitive positive formula such that and are satisfiable over , the formula is also satisfiable over . Then has an -canonical polymorphism that is -symmetric with .
Note that the assumptions in this lemma are precisely what we get from case 2) in Theorem 10.1. A lemma of similar style can be also found as Lemma 13 in the article of [MP20]. They obtain in this lemma a weakly commutative operation. The -symmetric operations from our article are a special case of weakly commutative operations. The difference between these two properties seems crucial for the next step of our proof: while -symmetric operations can often be “lifted” to canonical -symmetric operations this seems not clear in general for weakly commutative operations. This last step is necessary in order to obtain a contradiction to our assumption that does not have a Siggers operation.
In order to apply the results of [MP20] directly to obtain the dichotomy results of our article one would have to find a way to canonize the weakly commutative operations in a suitable way. It seems that this can be done by a similar proof as for Lemma 56 of [MP20] were the authors used among other things arguments which are also incorporated in our proof of the loop lemma (Theorem 10.1).
11 Conclusion
We classified the computational complexity of the network satisfaction problem for finite symmetric with a flexible atom and obtained a P versus NP-complete dichotomy. We gave a decidable criterion for that is a sufficient condition for the membership of in P, which is also necessary unless P=NP. We want to mention that if we drop the assumptions on to be symmetric and to have a flexible atom then the statement of Theorem 1.1 is false. An example for this is the Point Algebra; even though the of this representable relation algebra is in P [VKvB89], the first condition of Theorem 1.1 does not apply. However, if we only drop the symmetry assumption we conjecture that Theorem 1.1 still holds; yet it is not clear how to obtain the results of Section 7 in this case. Similarly, if we only drop the flexible atom assumption we conjecture that the statement also remains true. For this generalization it would be necessary to obtain Ramsey-type results like Theorem 3.6 without the assumption of the existence of a flexible atom.
Acknowledgements
The first author has received funding from the European Research Council under the European Community’s Seventh Framework Programme (FP7/2007-2013 Grant Agreement no. 681988, CSP-Infinity). The second author is supported by DFG Graduiertenkolleg 1763 (QuantLA).
References
- [All83] James F. Allen. Maintaining knowledge about temporal intervals. Communications of the ACM, 26(11):832–843, 1983.
- [AM94] Hajnal Andréka and Roger D. Maddux. Representations for small relation algebras. Notre Dame Journal of Formal Logic, 35(4):550–562, 1994.
- [AMM08] Jeremy F. Alm, Roger D. Maddux, and Jacob Manske. Chromatic graphs, Ramsey numbers and the flexible atom conjecture. The Electronic Journal of Combinatorics, 15(1), March 2008.
- [Bar11] Libor Barto. The dichotomy for conservative constraint satisfaction problems revisited. In Proceedings of the Symposium on Logic in Computer Science (LICS), Toronto, Canada, 2011.
- [BD13] Manuel Bodirsky and Víctor Dalmau. Datalog and constraint satisfaction with infinite templates. Journal on Computer and System Sciences, 79:79–100, 2013. A preliminary version appeared in the proceedings of the Symposium on Theoretical Aspects of Computer Science (STACS’05).
- [BJ17] Manuel Bodirsky and Peter Jonsson. A model-theoretic view on qualitative constraint reasoning. Journal of Artificial Intelligence Research, 58:339–385, 2017.
- [BK07] Manuel Bodirsky and Martin Kutz. Determining the consistency of partial tree descriptions. Artificial Intelligence, 171:185–196, 2007.
- [BK09] Manuel Bodirsky and Jan Kára. The complexity of temporal constraint satisfaction problems. Journal of the ACM, 57(2):1–41, 2009. An extended abstract appeared in the Proceedings of the Symposium on Theory of Computing (STOC).
- [BK20] Manuel Bodirsky and Simon Knäuer. Hardness of network satisfaction for relation algebras with normal representations. In Relational and Algebraic Methods in Computer Science, pages 31–46. Springer International Publishing, 2020.
- [BK21] Manuel Bodirsky and Simon Knäuer. Network satisfaction for symmetric relation algebras with a flexible atom. In Proceedings of AAAI, 2021. Preprint https://arxiv.org/abs/2008.11943.
- [BK22] Manuel Bodirsky and Simon Knäuer. The complexity of network satisfaction problems for symmetric relation algebras with a flexible atom. Journal of Artificial Intelligence Research, 75, 2022.
- [BM16] Manuel Bodirsky and Antoine Mottet. Reducts of finitely bounded homogeneous structures, and lifting tractability from finite-domain constraint satisfaction. In Proceedings of the 31th Annual IEEE Symposium on Logic in Computer Science (LICS), pages 623–632, 2016. Preprint available at ArXiv:1601.04520.
- [BMPP19] Manuel Bodirsky, Barnaby Martin, Michael Pinsker, and András Pongrácz. Constraint satisfaction problems for reducts of homogeneous graphs. SIAM Journal on Computing, 48(4):1224–1264, 2019. A conference version appeared in the Proceedings of the 43rd International Colloquium on Automata, Languages, and Programming, ICALP 2016, pages 119:1-119:14.
- [BN06] Manuel Bodirsky and Jaroslav Nešetřil. Constraint satisfaction with countable homogeneous templates. Journal of Logic and Computation, 16(3):359–373, 2006.
- [Bod08] Manuel Bodirsky. Constraint satisfaction problems with infinite templates. In Heribert Vollmer, editor, Complexity of Constraints (a collection of survey articles), volume 5250 of Lecture Notes in Computer Science, pages 196–228. Springer, 2008.
- [Bod12] Manuel Bodirsky. Complexity classification in infinite-domain constraint satisfaction. Mémoire d’habilitation à diriger des recherches, Université Diderot – Paris 7. Available at arXiv:1201.0856v8, 2012.
- [Bod15] Manuel Bodirsky. Ramsey classes: Examples and constructions. In Surveys in Combinatorics. London Mathematical Society Lecture Note Series 424. Cambridge University Press, 2015. Invited survey article for the British Combinatorial Conference; ArXiv:1502.05146.
- [Bod18] Manuel Bodirsky. Finite relation algebras with normal representations. In Relational and Algebraic Methods in Computer Science - 17th International Conference, RAMiCS 2018, Groningen, The Netherlands, October 29 - November 1, 2018, Proceedings, pages 3–17, 2018.
- [Bod21] Manuel Bodirsky. Complexity of Infinite-Domain Constraint Satisfaction. Lecture Notes in Logic (52). Cambridge University Press, 2021.
- [BP14] Manuel Bodirsky and Michael Pinsker. Minimal functions on the random graph. Israel Journal of Mathematics, 200(1):251–296, 2014.
- [BP15] Manuel Bodirsky and Michael Pinsker. Schaefer’s theorem for graphs. Journal of the ACM, 62(3):52 pages (article number 19), 2015. A conference version appeared in the Proceedings of STOC 2011, pages 655-664.
- [BP21] Manuel Bodirsky and Michael Pinsker. Canonical functions: a proof via topological dynamics. Homogeneous Structures, A Workshop in Honour of Norbert Sauer’s 70th Birthday, Contributions to Discrete Mathematics, 16(2):36–45, 2021.
- [BPP21] Manuel Bodirsky, Michael Pinsker, and András Pongrácz. Projective clone homomorphisms. Journal of Symbolic Logic, 86(1):148–161, 2021.
- [Bul03] Andrei A. Bulatov. Tractable conservative constraint satisfaction problems. In Proceedings of the Symposium on Logic in Computer Science (LICS), pages 321–330, Ottawa, Canada, 2003.
- [Bul16] Andrei A. Bulatov. Conservative constraint satisfaction re-revisited. Journal Computer and System Sciences, 82(2):347–356, 2016. ArXiv:1408.3690.
- [Bul17] Andrei A. Bulatov. A dichotomy theorem for nonuniform CSPs. In 58th IEEE Annual Symposium on Foundations of Computer Science, FOCS 2017, Berkeley, CA, USA, October 15-17, pages 319–330, 2017.
- [CH04] Matteo Cristiani and Robin Hirsch. The complexity of the constraint satisfaction problem for small relation algebras. Artificial Intelligence Journal, 156:177–196, 2004.
- [CL17] Hubie Chen and Benoît Larose. Asking the metaquestions in constraint tractability. TOCT, 9(3):11:1–11:27, 2017.
- [Com84] Stephen D. Comer. Combinatorial aspects of relations. Algebra Universalis, 18(1):77–94, February 1984.
- [Con15] Gabriel J. Conant. Model Theory and Combinatorics of Homogeneous Metric Spaces. PhD thesis, University of Illinois at Chicago, 10 2015.
- [DLPS07] Christian Delhommé, Claude Laflamme, Maurice Pouzet, and Norbert Sauer. Divisibility of countable metric spaces. European Journal of Combinatorics, 28(6):1746–1769, 2007.
- [Dün05] Ivo Düntsch. Relation algebras and their application in temporal and spatial reasoning. Artificial Intelligence Review, 23:315–357, 2005.
- [HH01] R. Hirsch and I. Hodkinson. Strongly representable atom structures of relation algebras. Transactions of the American Mathematical Society, 130(6):1819–1831), 2001.
- [HH02] Robin Hirsch and Ian Hodkinson. Relation Algebras by Games. North Holland, 2002.
- [Hir96] Robin Hirsch. Relation algebras of intervals. Artificial Intelligence Journal, 83:1–29, 1996.
- [Hir97] Robin Hirsch. Expressive power and complexity in algebraic logic. Journal of Logic and Computation, 7(3):309 – 351, 1997.
- [Hir99] Robin Hirsch. A finite relation algebra with undecidable network satisfaction problem. Logic Journal of the IGPL, 7(4):547–554, 1999.
- [HJK19] Robin Hirsch, Marcel Jackson, and Tomasz Kowalski. Algebraic foundations for qualitative calculi and networks. Theor. Comput. Sci., 768:99–116, 2019.
- [HN16] J. Hubička and J. Nešetřil. All those Ramsey classes (Ramsey classes with closures and forbidden homomorphisms). arXiv:1606.07979v2, 2016.
- [Hod97] Wilfrid Hodges. A shorter model theory. Cambridge University Press, Cambridge, 1997.
- [LM94] Peter B. Ladkin and Roger D. Maddux. On binary constraint problems. Journal of the Association for Computing Machinery, 41(3):435–469, 1994.
- [Lyn50] R. Lyndon. The representation of relational algebras. Annals of Mathematics, 51(3):707–729, 1950.
- [Mac11] Dugald Macpherson. A survey of homogeneous structures. Discrete Mathematics, 311(15):1599–1634, 2011.
- [Mad82] Roger Maddux. Some varieties containing relation algebras. Transactions of the American Mathematical Society, 272(2):501–501, February 1982.
- [Mad94] Roger D. Maddux. A perspective on the theory of relation algebras. Algebra Universalis, 31(3):456–465, September 1994.
- [Mad06a] Roger D. Maddux. Finite symmetric integral relation algebras with no 3-cycles. In Renate A. Schmidt, editor, Relations and Kleene Algebra in Computer Science, 9th International Conference on Relational Methods in Computer Science and 4th International Workshop on Applications of Kleene Algebra, RelMiCS/AKA 2006, Manchester, UK, August 29-September 2, 2006, Proceedings, volume 4136 of Lecture Notes in Computer Science, pages 2–29. Springer, 2006.
- [Mad06b] Roger Duncan Maddux. Relation Algebras: Volume 150. Studies in logic and the foundations of mathematics. Elsevier Science, London, England, May 2006.
- [MP20] Antoine Mottet and Michael Pinsker. Smooth approximations and CSPs over finitely bounded homogeneous structures, 2020. Preprint arXiv:2011.03978.
- [NR89] Jaroslav Nešetřil and Vojtěch Rödl. The partite construction and Ramsey set systems. Discrete Mathematics, 75(1-3):327–334, 1989.
- [Pos41] Emil L. Post. The two-valued iterative systems of mathematical logic, volume 5. Princeton University Press, Princeton, 1941.
- [RN99] Jochen Renz and Bernhard Nebel. On the complexity of qualitative spatial reasoning: A maximal tractable fragment of the region connection calculus. Artificial Intelligence, 108(1-2):69–123, 1999.
- [RN07] Jochen Renz and Bernhard Nebel. Qualitative spatial reasoning using constraint calculi. In M. Aiello, I. Pratt-Hartmann, and J. van Benthem, editors, Handbook of Spatial Logics, pages 161–215. Springer Verlag, Berlin, 2007.
- [Sig10] Mark H. Siggers. A strong Mal’cev condition for varieties omitting the unary type. Algebra Universalis, 64(1):15–20, 2010.
- [VKvB89] Marc Vilain, Henry Kautz, and Peter van Beek. Constraint propagation algorithms for temporal reasoning: A revised report. Reading in Qualitative Reasoning about Physical Systems, pages 373–381, 1989.
- [Zhu17] Dmitriy N. Zhuk. A proof of CSP dichotomy conjecture. In 58th IEEE Annual Symposium on Foundations of Computer Science, FOCS 2017, Berkeley, CA, USA, October 15-17, pages 331–342, 2017. https://arxiv.org/abs/1704.01914.
- [Zhu20] Dmitriy Zhuk. A proof of the CSP dichotomy conjecture. J. ACM, 67(5):30:1–30:78, 2020.