Stability and stable groups in continuous logic
Abstract.
We develop several aspects of local and global stability in continuous first order logic. In particular, we study type-definable groups and genericity.
Key words and phrases:
stable theory ; continuous logic ; definable group2000 Mathematics Subject Classification:
03C45 ; 03C60 ; 03C90Introduction
Continuous first order logic was introduced by A. Usvyatsov and the author in [BU], with the declared purpose of providing a setting in which classical local stability theory could be developed for metric structures. The actual development of stability theory there is fairly limited, mostly restricted to the definability of -types for a stable formula , the properties of -independence, and in case the theory is stable, properties of independence. Many fundamental results of classical stability theory, and specifically those related to stable groups, are missing there, and it is this gap that the present article proposes to fill.
We assume familiarity with [BU] and follow the notation used therein. Throughout denotes a continuous theory in a language . We do not assume that is complete, so various constants, such as of Fact 2.1, are uniform across all completions of (provided that is stable in , i.e., in every completion of separately).
By a model we always mean a model of . Whenever this is convenient, we shall assume that such a model is embedded elementarily in a large monster model , i.e., in a strongly -homogeneous and saturated model, where is much bigger than the size of any set of parameters under consideration. Notice that we may not simply choose a single monster model for , as this would consist of choosing one completion.
1. General reminders
We shall consider throughout a formula whose variables are split in two groups. We recall from [BU] that a definable -predicate is a definable predicate ), possibly with parameters, which is equivalent to an infinitary continuous combination of instances of :
Equivalently, is a -predicate if it can be approximated arbitrarily well by finite continuous combinations of instances of , possibly restricted to the use of the connectives , , alone.
Local types, i.e., -types for a fixed formula , are discussed in [BU, Section 6]. For a model and a tuple in some extension , the -type of over , denoted , is the partial type given by . The space of all -types over is denoted , and it is a compact Hausdorff quotient of . If is a -predicate over then determines , so we may identify with a mapping , sending . Every such mapping is continuous, and conversely, every continuous mapping from to is of this form.
For we define to be the quotient of where two types are identified if all -definable -predicates agree on them. This is again a compact Hausdorff space, a common quotient of and of (for the appropriate ), and the continuous mappings are precisely the -definable -predicates. In particular this does not depend on the choice of .
Lemma 1.1.
Let be a structure, a (metrically) compact set and let be a formula (or a definable predicate, which we may always name by a new predicate symbol without adding any structure). Then is a -predicate (with parameters in ) and for any tuple , the infimum is attained by some .
In particular, is definable in .
Proof.
Since is compact we can find a sequence such that for every there is such that . Then is arbitrarily well approximated by formulae of the form as . Finally, the infimum of a continuous function on a compact set is always attained. ∎
It will also be convenient to adopt the following somewhat non standard terminology:
Definition 1.2.
Let be a model, a subset. We say that is saturated over if it is strongly -homogeneous and saturated. (In fact, for all intents and purposes it will suffice to require to be strongly -homogeneous and saturated once every member of is named.)
We say that a partial type over is -invariant if is saturated over and is fixed by the action of .
An essential notion for the study of definability of types and canonical bases in a stable theory is that of imaginary elements and sorts. Let us give a brief reminder of their construction, as given in [BU]. Consider a definable predicate with parameters in some set , let us denote it by . Then we may assume that is countable, say , and express as a uniform limit of formulae . Furthermore, using a forced limit argument, we may assume that the sequence converges uniformly to some infinitary definable predicate , giving sense to for any sequence . If is any structure, we equip with the pseudo-metric and define the sort of canonical parameters for in , denoted , as the complete metric space associated to . In other words, we divide by the kernel , obtaining a true metric on the quotient, and pass to the completion. The predicate is uniformly continuous with respect to in the metric , and therefore passes first to the quotient and then to the completion, thus inducing a uniformly continuous predicate , where . We may now add as a sort to and equip the new structure with an additional predicate symbol for . This does not add structure to the original sorts of , elementary embeddings of structures commute with this construction (by which we mean, in particular, that an elementary embedding extends uniquely to ), elementary classes and model completeness thereof are respected by this construction, and so. The construction can be slightly simplified when only uses finitely many parameters, e.g., if it is an honest formula rather than a definable predicate, but we are going to need the general case.
If is the image of in then is indeed a canonical parameter for , in the sense that an automorphism of or of an elementary extension thereof (and such an automorphism extends uniquely to ) fixes if and only if it fixes the predicate . By construction we have , and by a convenient abuse of notation we shall permit ourselves to write instead of either one.
By an imaginary sort we mean any sort added in this fashion, and by imaginary elements we mean members of such a sort. We may repeat this construction for any other definable predicate , or for any family of predicates. A delicate point here is that even with a countable language one can construct continuum many definable predicates for whose canonical parameters imaginary sorts may be added. For the purposes of stability theory, however, no more imaginary sorts than the size of the language are truly required: we need the sort for each formula (see Fact 2.1 for the predicate ). Therefore, with some intentional ambiguity, by we shall usually mean “ along with all the imaginary sorts we are going to need”, e.g., all the sorts . By , , etc., we mean the respective operations in the structure .
Lemma 1.3.
Let be a set of parameters and let be a definable predicate with parameters possibly outside . Then is -definable if and only if it is -invariant, i.e., if and only if its canonical base belongs to .
Similarly, a set which is definable (type-definable) with some parameters is definable (type-definable) over if and only it is -invariant.
Proof.
First, let us consider an arbitrary surjective continuous mapping between two compact Hausdorff topological spaces. Then is also closed, so is closed if and only if is closed in . Since is surjective, is open if and only if is open, and a mapping is continuous if and only if is continuous.
The assertion now follows from applying the previous paragraph to the restriction mapping , where contains all the needed parameters, using the correspondence between type-definable sets and closed sets, and between definable predicates and continuous functions. For a definable set , just argue for the definable predicate . ∎
Fact 1.4.
[BU, Lemma 6.8] Let be any formula, a set, a saturated model over , and let . Then acts transitively on the set of extensions of in .
The following notion and fact also appear (and are used much more extensively) in [Bena, Section 1]:
Definition 1.5.
Let and be two type-definable sets. We say that is a logical neighbourhood of , in symbols , if there is a set of parameters over which both and are defined such that in .
Notice that the interior of does depend on (i.e., if then calculated in may be larger than the pullback of the interior of in ). We may nonetheless choose any parameter set we wish:
Lemma 1.6.
Assume that is type-definable with parameters in , type-definable possibly with additional parameters not in . Then:
-
(i)
If then in for any set over which both and are defined.
-
(ii)
If then there is an intermediate logical neighbourhood , which can moreover be taken to be the zero set of a formula with parameters in .
-
(iii)
If then there is a logical neighbourhood such that . Moreover, we may take to be a zero set defined over .
Proof.
Assume , where is type-definable over , and over . Let consist of all formulae over which are zero on . If then , and is defined by the partial type . By compactness in there is a condition in which already implies . Let be the zero set of the formula where .
Then in we have , i.e., , proving the first two items. The third item now follows from the fact that is a normal topological space. ∎
2. Definability and forking of local types
Having fixed a theory , we shall call here a formula stable if it is stable in , that is, if it does not have the order property in any model of . The order property was defined for continuous logic in [BU], but the reader may simply use Fact 2.1 below as the definition of a stable formula.
Let us introduce some convenient notation. If is any formula with two groups of variables, denotes the same formula with the groups of variables interchanged. More generally, let us define
where is the median value combination:
Thus in particular and every instance of is a -predicate.
Fact 2.1.
Let be a stable formula. Let be a model and let be a complete -type. Then
-
(i)
The type is definable over , i.e., there exists an -definable -predicate such that for all . Moreover, this definition is uniform, in the sense that there exists an infinitary definable predicate which only depends on such that is equal to an instance where .
-
(ii)
For every there exists a number (which depends on and on but not on ) and a tuple in (which does depend on ) such that
-
(iii)
Assume moreover that is saturated over some subset . Then in the previous item the tuples can be chosen so that each realises .
Proof.
The first two items are taken from [BU, Lemma 7.4 and Proposition 7.6]. The third item, while not explicitly stated there, is immediate from the proof. ∎
By abuse of notation we may sometimes write , where is the canonical parameter for the definition. This canonical parameter is called the canonical base of , denoted .
We recall that for , does not fork over if it admits an extension which is definable over . In this case itself does not fork over or . A type over a model clearly admits a unique non forking extension to any larger model (and therefore set), so this definition does not depend on the choice of .
We proved in [BU, Proposition 7.15] that every -type over a set admits a non forking extension to every model (and therefore every set) containing . A minor enhancement of that result will be quite useful.
Lemma 2.2 (Existence of non forking extensions).
Let be a stable formula, a set, a saturated model over . Let be a consistent -invariant partial type over . Then there exists compatible with which does not fork over .
Proof.
Corollary 2.3.
Let be a stable formula, a set, a saturated model over . Then does not fork over if and only if it is -invariant.
Proof.
Left to right follows from the definition, right to left from Lemma 2.2. ∎
Corollary 2.4.
Let be a set, a saturated model over and a consistent -invariant partial type over . Then there exists a complete type such that , and for every stable formula the restriction does not fork over .
Proof.
We may assume that . Index all stable formulae of the form by . We define an increasing sequence of consistent -invariant partial types over , starting with . Given , by Lemma 2.2 there is be non forking over and compatible with , so is consistent and -invariant. For limit we define . Finally, let be any completion of . Then will do. ∎
It follows that if the theory is stable then every complete type over a set admits non forking extensions. The same fact was proved in [BU] using a somewhat longer “gluing” argument.
Fact 2.5 (Symmetry [BU, Proposition 7.16]).
Let be a model, , . Then .
Proposition 2.6.
Let be a stable formula, a model, . For each let be the definition of a non forking extension of to .
-
(i)
Each is a definable -predicate over .
-
(ii)
A -type does not fork over if and only if for all .
-
(iii)
A -type over is stationary, i.e., admits a unique non forking extension to every larger set.
-
(iv)
Let . Then the partial type defines the set of -types which do not fork over :
-
(v)
For every , the set is closed.
Proof.
The first item is by Fact 2.1 and the definition of non forking.
For the second, fix , let and let be the non forking extension defined by . Assume does not fork over , so is a -predicate over . By Fact 2.5,
Conversely, assume that for all , and let be any non forking extension of . Then , proving also the third item. The fourth item is just a re-statement of the second.
For the last item we may assume that . The set is closed, and so is its projection to . This projection is precisely the set of types which do not fork over . ∎
Proposition 2.6.(iii) is the analogue of the finite equivalence relation theorem in continuous logic. It has already appeared as [BU, Proposition 7.17]. In case is stationary, the unique non forking extension to will be denoted . Similarly, we write for the definition of where is any model (and this does not depend on the choice of ). Thus, in hindsight, in the statement of Proposition 2.6, the definitions are uniquely determined, .
Corollary 2.7.
Let be a stable formula, a set, a saturated model over . Let . Then acts transitively on the set of non forking extensions of in . If is stable and then acts transitively on the set of non forking extensions of to .
Proof.
Corollary 2.8.
Let be a stable formula a model. A type is definable over if and only if it does not fork over and is stationary.
Proof.
Corollary 2.9.
Let be a stable formula, a set, a complete type over , and let . Then is compatible with every non forking extension of .
Proof.
We pass to forking of single conditions.
Definition 2.10.
Let be an instance of a stable formula, a set. We say that a condition does not fork over if there exists a -type non forking over such that .
Proposition 2.11.
Let be an instance of a stable formula, a set of parameters. Then the following are equivalent:
-
(i)
The condition does not fork over .
-
(ii)
Every family of -conjugates of is consistent.
-
(iii)
For every set there exists a complete type such that does not fork over for any stable formula (if is stable: does not fork over ) and .
Proof.
-
(i) (ii).
Let witness that does not fork over . Then any non forking extension of to a large model is -invariant.
-
(iii) (iv).
We may assume that is saturated over . Let consist of all the -conjugates of in . It is consistent by assumption and -invariant by construction so we may apply Corollary 2.4.
-
(v) (i).
Immediate. ∎
We may define the non forking degree of over to be
An easy compactness argument shows that the infimum is attained and the condition does not fork over . In addition, by the existence of non-forking types we have .
Definition 2.12.
A faithful continuous connective in variables is a continuous function satisfying .
If is a faithful continuous connective and a sequence of definable predicates, then the definable predicate is called a faithful combination of .
Since a continuous function to can only take into account countably many arguments, we may always assume that . Notice that any connective constructed using and alone is faithful (so in particular the median value connective is). Similarly, any uniform limit of faithful combinations is faithful.
Lemma 2.13.
Let be a stable formula. Let be a set of parameters, a tuple, . Let . Then is a faithful combination of -conjugates of .
Proof.
By the preceding discussion and the last item of Fact 2.1. ∎
Lemma 2.14.
Let be an instance of a stable formula, a set of parameters. Then there exists an -definable predicate such that for every tuple (not necessarily in ):
Moreover, can be taken to be a faithful combination of -conjugates of .
Proof.
Fix a model , saturated over . Let . Let be the unique non forking extension of . Let , where is the canonical parameter for the definition. By the previous Lemma, is a faithful combination of -conjugates of .
Let be the set of -conjugates of . Since is algebraic over , is (metrically) compact. By Lemma 1.1 is a continuous combination of instances with , i.e., of -conjugates of , and it is clearly a faithful combination. Thus is a faithful combination of -conjugates of , and it is clearly over .
We may assume that , and let be the unique non forking extension of . Then
Since and are the sets of non forking extensions of and of , respectively, to , we are done. ∎
Theorem 2.15 (Open Mapping Theorem).
Assume is stable, and let be any sets of parameters. Let be the set of types which do not fork over . Then is compact and the restriction mapping sending is an open continuous surjective mapping.
Proof.
We already know that is compact and that is continuous and surjective.
Notice that a similar proof yields that if is stable then the restriction mapping is open, where denotes the set of -types which do not fork over .
It follows from Lemma 2.14 that a -type (and therefore a -type) over an arbitrary set is definable over , but of course the same definition applied to a larger set need not give a consistent complete type. This yields the following (adaptation of a) classical result:
Theorem 2.16 (Separation of variables).
Let be an instance of a stable formula, and let be a type-definable set in the sort of , say with parameters in . Then there is a subset (at most countable) and a -definable predicate such that .
Moreover, can be taken to be a faithful combination of instances such that (or even where is an arbitrary small subset).
Proof.
Fix a model , saturated over , and let . Let be as in Lemma 2.14. Then is definable over and therefore over where is an appropriate countable subset. Then for all we have . Now let be the monster model and . By saturation of we can find there some . Then and , as desired.
The moreover part follows from the proof. ∎
It follows that if is an -type-definable set and is a type-definable subset, then is type-definable over for some countable . If is a definable set then it is definable over (by Lemma 1.3, since the definable predicate is -invariant).
Proposition 2.17.
Let be an instance of a stable formula, a set of parameters. Then the following are equivalent:
-
(i)
The condition does not fork over .
-
(ii)
There is an -definable predicate which is a faithful combination of -conjugates of such that is consistent.
Proof.
Fix a model saturated over .
-
(i) (ii).
Let be as in Lemma 2.14. Let also be non forking over such that . Then .
-
(iii) (i).
Let be definable over as in the assumption (so and is a faithful continuous connective).
By Lemma 2.2 there exists compatible with and non forking over , so in particular -invariant. Then by faithfulness, so for all there exists such that . Up to an automorphism fixing we may assume that , and by invariance for every .
We have thus shown that for every , any set of -conjugates of is consistent. By compactness the same holds for . ∎
3. Heirs and co-heirs
We turn to study co-heirs, and more generally, approximately realised partial types, in continuous logic. In the context of stability, approximate realisability serves as a criterion for non forking. For an earlier treatment of co-heirs in the context of metric structures see [Ben05, Section 3.2].
Definition 3.1.
Let be two sets of parameters. We say that a partial type over is approximately realised in if every logical neighbourhood (Definition 1.5) of over is realised in .
If is a model, , and is approximately realised in , we may say that is a co-heir of its restriction to .
Remark 3.2.
-
(i)
The classical logic analogue of an approximately realised type is a finitely realised one, but this terminology would be misleading in the continuous setting.
-
(ii)
A complete type over a model is always approximately realised there. (This is essentially the Tarski-Vaught Criterion.)
Fact 3.3.
Let and let be a partial type over .
-
(i)
Let consist of all types over which are realised in , the closed set defined by . Then is approximately realised in if and only if . In particular, is the set of all complete -types over which are approximately realised in .
-
(ii)
If then is approximately realised in as a partial type over if and only if it is approximately realised in as a partial type over .
-
(iii)
If is approximately realised in then it extends to a complete type which is approximately realised in .
-
(iv)
A type over a model admits extensions to arbitrary sets which are approximately realised in .
Proof.
We prove the first two items together. Clearly if is approximately realised in as a partial type over then it is approximately realised in as a partial type over , in which case every neighbourhood of in intersects and by a compactness argument intersects . Finally, assume and assume that . Let and let be its projection to . Then is compact, , so is a neighbourhood of . By assumption there exists such that . Then , i.e., , as desired.
For the third item, any will do. For the fourth, use the fact that a type over a model is approximately realised there. ∎
Fact 3.4.
Let be a model saturated over . If or is approximately realised in then it is -invariant.
Proof.
We only consider the case , since the case follows from it. Say , , and let be given. By assumption there is such that
As we assumed that we have in particular and thus , for every . We conclude that , as desired. ∎
Lemma 3.5.
Let , approximately realised in , and assume is stable. Then does not fork over .
Proof.
Let be saturated over and let extend , still approximately realised in . Then , and thus , are -invariant, so does not fork over and neither does . ∎
Proposition 3.6.
Let be a stable formula, a model, . Let also be a complete -type, and a complete type over such that . Then the following are equivalent:
-
(i)
is approximately realised in .
-
(ii)
is approximately realised in .
-
(iii)
does not fork over .
Proof.
Similarly,
Proposition 3.7.
Assume is stable. Let be a model of , , . Then the following are equivalent:
-
(i)
does not fork over .
-
(ii)
is approximately realised in .
If is saturated over then these are further equivalent to
-
(iii)
is -invariant.
Definition 3.8.
Let be a model, . A type is said to be an heir of its restriction to if for every formula with and , and for every , there are such that .
Clearly every type over a model is an heir of itself. Also, it is not difficult to check that if is a model and , are two tuples possibly outside then
Finally, a standard compactness argument yields that if and is an heir of then it admits an extension to which is an heir as well.
Lemma 3.9.
Let be a model, . Then is definable if and only if it has a unique heir to every superset .
Proof.
(We follow Poizat [Poi85, Théorème 11.07].) For left to right, assume is definable and let be an heir of , where . Let be a formula over and let be the -definition of , . Assume that , i.e., . Then there is such that , a contradiction.
Conversely, assume admits a unique heir to every structure. let be along with a new predicate symbol for each formula (here is fixed, may vary with ). We define an expansion of by interpreting . Assume now that and let
It is not difficult to see that is a complete, consistent type, and that it is moreover an heir of . Since, by assumption, the heir is unique, is the unique expansion of which is an elementary extension of . In other words, a model of admits at most one expansion to a model of . By Beth’s Theorem (see [Ben09]) for each formula there exists an -definable predicate such that . In particular, for every , and is definable. ∎
Notice that for a pair of models we could have defined a notion of a -type over a being an heir of its restriction to , in which case Lemma 3.9 holds, with the same proof, for local types.
Theorem 3.10.
The following are equivalent for a theory :
-
(i)
The theory is stable.
-
(ii)
Every type over a model has a unique co-heir to any superset.
-
(iii)
Every type over a model has a unique heir to any superset.
Proof.
-
(i) (ii).
Assume is stable, , and is a co-heir of . Let be saturated over and let extend , also a co-heir of . Then is -invariant and therefore the unique non forking extension of to . Thus is the unique non forking extension of to .
-
(iii) (iv).
Let be a model, . In order to show that has a unique heir to every it is enough to consider the case where is a finite tuple. So indeed, assume that realises an heir of to . Then is a co-heir of and by assumption it is uniquely determined by and by . It follows that is uniquely determined by and , as desired.
-
(v) (i).
The assumption and Lemma 3.9 yield that every type is definable, so is stable. ∎
4. Invariant types, indiscernible sequences and dividing
Fact 4.1.
Let be a model saturated over , and let be -invariant. Let be a sequence constructed inductively, choosing each to realise .
Then the sequence is -indiscernible, and its type over depends only on .
Proof.
Standard. ∎
The common type over of such sequences will be denoted by . For every finite or countable we may construct just as well. By a gluing argument, is a complete type of an -indiscernible sequence in , and is of course -invariant.
Lemma 4.2.
Let be a set, a stable formula, a stationary -type. Let be saturated over , and let , invariant over . Let be an -indiscernible sequence as constructed in Fact 4.1.
Then the sequence converges uniformly to the definition at a rate which only depends on .
Proof.
Since is -invariant, it does not fork over , so .
Fix . By Fact 2.1 there is and a sequence such that , and such that furthermore . By Fact 4.1 we have . In addition, is over , so .
Consider now . First of all, by exactly the same argument as above, for every we have . In addition, for any there exists a subset such that (from any set of reals one can choose a subset of size with the same median value). Thus for all , where depends only on and , as desired. ∎
Proposition 4.3.
Let be an instance of a stable formula, a set of parameters. Then the following are equivalent:
-
(i)
The condition does not fork over .
-
(ii)
If is an -indiscernible sequence, , then the set of conditions is consistent (i.e., the condition does not divide over ).
Proof.
-
(i) (ii).
If is an -indiscernible sequence and then each is an -conjugate of .
-
(iii) (i).
Fix models where is saturated over . Let . By Lemma 2.2 there exists extending such that does not fork over , i.e., such that . Let be an -invariant extension of . Finally, let . Then is an -indiscernible sequence, and a fortiori -indiscernible, in . Thus by assumption there exists such that for all . In addition, by Lemma 4.2 we have
Let be a non forking extension of . Then , witnessing that does not fork over , as desired. ∎
5. Canonical bases
Recall that the canonical base of a stationary type in a stable theory is , namely the set of all canonical parameters of -definitions of .
Proposition 5.1.
Assume is stable, and let be stationary. Then:
-
(i)
.
-
(ii)
does not fork over .
-
(iii)
is stationary.
-
(iv)
is minimal for the three previous properties, meaning that if and is a stationary non forking restriction then .
Proof.
The first two items are immediate, while the third is by Corollary 2.8. Under the assumptions of the fourth we have . ∎
The four properties listed in Proposition 5.1 determine the canonical base up to inter-definability. Indeed, if has all four then but also , whereby . In this case we say that is a canonical base for .
Proposition 5.2.
Assume is stable, and let be stationary. Let be the unique non forking extension of , where is saturated over . Then a (small) set is a canonical base for if and only if, for every : .
Proof.
Let . It follows directly from the definitions that an automorphism of fixes if and only if it fixes for every formula , if and only if it fixes every member of . A small set is another canonical base for if and only if which is further equivalent to and being fixed by the same automorphisms. ∎
We propose an alternative characterisation of canonical bases using Morley sequences. In the case of classical first order logic it is more or less folklore. Recall that a Morley sequence in a (stationary) type is a sequence of realisations of which is independent over , i.e., such that for all . It follows by standard independence calculus that for every two disjoint index sets . From stationarity of it follows that the sequence is indiscernible over , and its type over , which we may denote by , is uniquely determined by .
It is not difficult to check that if satisfies the assumptions of Fact 4.1 then the definition of which appears thereafter agrees with the one given here. In the general case, let be saturated over and let be the non forking extension of . Then by construction, , where the first is the type of a Morley sequence as defined here, and the second the type defined after Fact 4.1.
Definition 5.3.
Let be a sequence of tuples, or, for that matter, even of sets. Let denote the tail . We define the tail definable closure of as
It is not difficult to see that for an indiscernible sequence , consists precisely of all over which is indiscernible.
Lemma 5.4.
Let and be indiscernible sequences such that the concatenation is indiscernible as well. Then . Moreover, every automorphism which sends to necessarily fixes .
Proof.
For let be the sequence , namely the sequence obtained by replacing the first elements of with the corresponding elements from . Since is indiscernible so is for each , and there exists an automorphism sending . Now let . Since is definable over it is fixed by , so . This holds for all , whence .
Fix an automorphism which sends to (which must necessarily exist). Then , so . Thus fixes . Applying we obtain that , as desired. ∎
Theorem 5.5.
Let be a stationary type and let be a Morley sequence in . Then is a canonical base of .
Proof.
First of all, we have seen that is stationary, with the same canonical base as . It is also not difficult to check that a Morley sequence in is also a Morley sequence in . It is therefore enough to prove for , i.e., we may assume that .
So let be saturated over and let be the non forking extension of . As pointed above, . By Lemma 4.2 is definable over , so . Also, every tail of a Morley sequence is a Morley sequence, whence .
Conversely, let be an automorphism fixing . Then fixes and therefore sends to another Morley sequence in , say . Let be a third Morley sequence in , . Then both and can be verifies to be Morley sequences in (of length ), and in particular indiscernible. We can decompose where and . By the Lemma and this set is fixed by , and therefore by . Thus , and the proof is complete. ∎
It is also a fact, which we shall not prove here (but is proved as in classical logic), that in a stable theory every indiscernible sequence is a Morley sequence in some type, say . Let and , which does not depend on . By the Theorem, and is a Morley sequence in .
In the case of probability theory this is a well known fact. Indeed, in probability algebras or in spaces of random variables (say -valued, see [Benb]), the canonical base of a type (in the real sort) can be represented by a set of real elements, so there is no need to consider imaginaries. Then Theorem 5.5 tells us that if is sequence of random variables which is indiscernible (i.e., exchangeable) and is its tail algebra then the sequence is i.i.d. over , meaning that the random variables are independent over and have the same conditional distribution over .
Corollary 5.6.
Assume is stable, and let be stationary. Let be a Morley sequence in , . Then and .
Proof.
The first independence is immediate and implies . By Theorem 5.5 we have and the second independence follows. ∎
6. Stable type-definable groups and their actions
We turn to consider groups, and more generally, homogeneous spaces, which are definable or type-definable in a stable theory.
6.1. Generic elements and types in stable group actions
Let be a homogeneous space, type-definable in models of a stable theory . This is to say that is a type-definable group and a type-definable set, equipped with a type-definable (and therefore definable) transitive group action . For convenience let us assume that both are defined without parameters. We shall identify and with their sets of realisations in a monster model . We are particularly interested in the case where where acts on itself either on the left or on the right .
Given a partial type in the sort of we let denote the subset of defined by .
Definition 6.1.
-
(i)
A generic set in is a subset finitely many -translates of which cover .
-
(ii)
A generic partial type in is a partial type such that every logical neighbourhood of (as per Definition 1.5) defines in a generic set. Single conditions as well as complete types are generic if they are generic as partial types.
-
(iii)
We say that an element is generic over a set if is generic.
-
(iv)
A left-generic set in is a subset which is generic under the action of on itself on the left. We define partial types in the sort of to be left-generic accordingly. Similarly for right-generic.
Let be a partial type. Clearly, if is a generic set then is a generic partial type, but the converse is not always true. In classical logic, if consists of a single formula (i.e., if is a relatively definable subset of , and so is its complement), then is its own logical neighbourhood and the two notions coincide. Unfortunately, this will generally never happen in continuous logic (except for or ).
Lemma 6.2.
The following are equivalent for a partial type in the sort of , with parameters in a set :
-
(i)
The partial type is generic in .
-
(ii)
For every formula over , if the condition is a logical neighbourhood of then it is a generic condition.
Proof.
One direction is immediate, the other follow from Lemma 1.6. ∎
Let denote the set of all complete types over implying . Equipped with the induced topology from , it is a compact space, and the set of all generic complete types over is closed. Closed subsets of are in bijection with partial types over implying , i.e., with type-definable subsets of using parameters in . If are two such sets, say that is a logical neighbourhood of relative to , in symbols , if where the interior is calculated in . This is equivalent to saying that there exists a true logical neighbourhood such that . Thus a type-definable set is defined by a generic partial type in if and only if every relative logical neighbourhood of in defines a generic set.
For and , let . Somewhat superfluously, we may also define .
Lemma 6.3.
Let be a set of parameters, .
-
(i)
If is type-definable over , say by a partial type , then is also type-definable over by a partial type which will be denoted (or ). Moreover, is generic if and only if is.
-
(ii)
If is a complete type then , and is a homeomorphism, and restricts to a homeomorphism of the set of generic types with itself.
Proof.
We only prove the parts regarding genericity. Indeed, assume that is generic, and let . Then , so is a generic subset of . It follows immediately that so is . Thus is a generic partial type. For the converse replace with . ∎
Similarly, for and we define . For we define .
Lemma 6.4.
Let be a set of parameters, .
-
(i)
If is type-definable over , say by a partial type , then is also type-definable over by a partial type which will be denoted (or ). Moreover, if is left-generic then is generic.
-
(ii)
If is a complete type then , and is a continuous surjection, sending left-generic types to generic types.
Notice that we do not claim that every generic type in is the image under of a left-generic type in (this is true if is stable).
Proof.
Essentially identical to that of Lemma 6.3. ∎
Under the assumption that the theory is stable we shall show that generic types exist and study some of their properties. We follow a path similar to that followed in [Pil96]. Toward this end we construct an auxiliary multi-sorted structure in a language (in addition to sorts and , consists of additional sorts which we shall described later). We define the distance on the first two sorts by
This coincides with the original distance in if the latter is invariant under the action of (on the left). In any case, is a distance function, invariant under the action of , and satisfies . On the other hand, if in then in as well (if not, then by a compactness argument, for some there would exist such that , an absurd). It follows that is a complete metric space. The same observations hold for .
Let now consist of all -formulae of the form where is in the sort of . For each , there will be a sort , consisting of all canonical parameters of instances of in . The canonical parameter of will be denote , or if there is no ambiguity. We put on it the standard metric, namely
The only symbols in the language , in addition to the distance symbols of the various sorts, are a predicate symbol for each formula , interpreted by
Since is uniformly continuous in all its variables, so is . These definitions make a continuous -structure.
If is definable then is interpretable in and is stable (assuming is). In the general case, all we know is that is saturated for quantifier-free types in which only appear. It follows from stability in that each formula , with any partition of the variables, is stable.
For define a mapping by sending to , to , and fixing all the auxiliary sorts. This is easily verified to be an automorphism of . Since the action of on is assumed to be transitive, if then all elements of have the same type over in , and similarly all elements of .
Lemma 6.5.
Assume that is stable. Then the following are equivalent for an instance :
-
(i)
The condition is generic in .
-
(ii)
The condition does not fork in over .
-
(iii)
The condition does not fork in over .
Proof.
Recall that the -formula with this (or any other) partition of the variables is stable in . For let . By Lemma 6.2, the condition is generic if and only if is a generic set for all .
-
(i) (ii).
Assume first that is generic in , i.e., that the set is generic for every . Find such that , and find such that does not fork over (in symbols ). Since we may assume that , so . Thus does not fork over . Applying we see that does not fork over either. It follows that does not fork over .
-
(iii) (iv).
Immediate.
-
(v) (i).
Assume now that does not fork over . By Proposition 2.17 there are for and a faithful combination which is definable over and such that is consistent. Since is saturated for quantifier-free types involving only , there is such that . Since all elements of have the same type over in , we see that for all . Assume (toward a contradiction) that there exists such that is not generic. By compactness we can find such that for all , i.e., . Since the combination above was faithful we get , a contradiction. ∎
Lemma 6.6.
Assume that is stable and that is a generic condition in . Then it does not fork over .
Proof.
By Proposition 4.3 it will be enough to show that does not divide over . For this purpose let be any indiscernible sequence with . Since , the sequence is indiscernible as well, and thus the sequence is indiscernible in . On the other hand, since the condition is generic, by Lemma 6.5 the condition does not fork over , so is consistent. Since is saturated for such formulae, there is such that , i.e., , for all , as desired. ∎
From now on we assume that is stable.
Proposition 6.7.
Let be a partial type over . Then is generic if and only if it extends to a complete generic type over , i.e., if and only if contains a generic type. In particular, generic types exist over every set.
Proof.
Right to left is clear, so let us prove left to right. Assume therefore that is a generic partial type. Since the set of complete generic types is closed it will be enough to show that every logical neighbourhood of contains a generic type, and we may further restrict our attention to logical neighbourhoods defined by a single condition . Since is generic in so is . By Lemma 6.5 does not fork over in . By Corollary 2.4 there exists a type such that and in addition does not fork over for every formula . Let
This type is approximately finitely realised in (since is in ) and therefore consistent. By Lemma 6.5 every condition in is generic (since does not fork over ), and by Lemma 6.2, is generic, and so is . We have thus found a generic type and the proof is complete. ∎
Proposition 6.8.
Assume . Then a type is generic if and only if it does not fork over and is generic. In particular, a generic type does not fork over .
Proof.
First of all, the last assertion follows from Lemma 6.6 and the fact that the set of non forking types is closed.
We now prove the main assertion. For left to right, if is generic then clearly so is , and by the previous paragraph does not fork over . For the converse, assume that does not fork over and is generic. Replacing with a non forking extension we may assume that . By Proposition 6.7 there is extending which is generic, and by what we have just shown it is also non forking over . Since there is sending to , and therefore to . Thus is generic as well. ∎
We can also complement Lemma 6.3:
Proposition 6.9.
The action of on the set of generic types in is transitive.
Proof.
Let be two generic types. Define
and define similarly. Let and let , . Since is saturated for formulae of this form we may realise and in , and by transitivity there exists such that is realised. Since is an automorphism of we must have . In addition, neither of , or forks over , whereby , i.e., . ∎
Theorem 6.10.
Let be a type-definable group in a stable theory , acting type-definably and transitively on a type-definable set .
-
(i)
If (where , ) and is left-generic over then is generic over and .
-
(ii)
An element is generic if and only if implies for every . Moreover, in this case is generic over as well.
-
(iii)
An element is left-generic over if and only if is.
-
(iv)
An element is left-generic if and only if it is right-generic (over ). From now on we shall only speak of generic elements and types in .
-
(v)
An element is generic over if and only if it is generic over and .
Proof.
We use Proposition 6.8 repeatedly.
For the first item, let , , and assume that . If is left-generic over then it is left-generic over . By Lemma 6.4 is generic over . It follows that is generic over and that , as desired.
For the second item, left to right, as well as the moreover part, are proved as in the previous argument, using Lemma 6.3. For right to left, assume that implies for all . We may choose which is left-generic over such that . Then by assumption, is generic over by the first item, and is generic over by the moreover part.
For the third item, let be left-generic over . Choose left-generic over such that . By the first item is generic over and . This can be re-written as . By the first item again, is left-generic over . Notice that is left-generic if and only if is right-generic, yielding the fourth item as well.
The last item is just Proposition 6.8. ∎
6.2. Stabilisers
We have already observed in Lemma 6.3 that for any set of parameters , a group element induces a homeomorphism on . It is also not difficult to check that , whence a group action of on . In addition, we have seen that it restricts to an action by homeomorphism of on the set of generic types in .
Specifically, we obtain an action of on . The stabiliser of a type under this action is . For a stationary type we define .
Proposition 6.11.
Let be stationary. Then stabiliser is a sub-group of type-definable over .
Moreover, assume that , and . Then if and only if .
Proof.
We may assume that .
Let be a formula, in the sort of . Let be a variable in the sort of . Then is a definable predicate on , i.e., a continuous function . By Tietze’s Extension Theorem this extends to a continuous function . For clarity we shall use to denote the corresponding definable predicate.
Once this technical preliminary is taken care of we see that is defined by the following axiom scheme:
The moreover part easily follows. ∎
Lemma 6.12.
Let be a type-definable subgroup of bounded index, say with parameters in , and let . Then is generic over in if and only if it is generic over in .
Proof.
Naming in the language we may assume that . Since has bounded index we may enumerate its cosets . Let be generic over . Then for some , and is generic in . Now let be generic in . Without loss of generality we may assume that . Then is generic both in and in and . Thus is generic in as well. We have thus shown that every generic of is a generic of . A similar argument shows that every generic of in is generic in . ∎
Proposition 6.13.
A type is generic if and only if has bounded index in .
Proof.
There are only boundedly many generic types over , since they do not fork over and therefore determined by their restriction to . In addition, the action of on restrict to an action of on the space of generic types, so the stabiliser of a generic type must be of bounded index.
Conversely, assume has bounded index, and let . Then there exists which is generic in over , and we may further assume that . Then is generic over , i.e., is generic. ∎
Since acts transitively on the generic types over , the stabilisers of generic types are all conjugate. It is also not difficult to check that if is generic, is a generic type of (and therefore of ), and , then . If is any other generic of then (since acts transitively on its own generic types, on the left as well as on the right) there exists such that and . Thus the right action of on send each and every generic type of onto the generic types of , complementing Lemma 6.4.
Theorem 6.14.
Let be a type-definable group in a stable theory, say over . Then admits a smallest type-definable group of bounded index (over any set of parameters), called the connected component of , and denoted . It has the following additional properties.
-
(i)
The connected component is a normal subgroup of , type-definable over .
-
(ii)
The stabiliser of every generic type is equal to .
-
(iii)
Each coset contains a unique generic type over .
-
(iv)
The generic type of is definable over .
-
(v)
If is any stationary generic type over a small set then .
Proof.
We start by constructing and proving the second item. Since left generic and right generic are the same, the action of on the generic types is transitive on either side. In particular, if are generic then there exists such that , and thus . Let this unique stabiliser of generic types be denoted . Then is type-definable, and since -invariant, so is , and we may conclude that is type-definable over as well. We also already know that has bounded index in . Assume now that is another type-definable subgroup of bounded index, say over (otherwise name the parameters in the language). Then there exists generic in , so , whereby . Thus is indeed the smallest type-definable subgroup of of bounded index. Notice that is also type-definable of bounded index for every , so is normal in . This concludes the proof of the first two items.
Let be generic in . Since acts transitively on its generic types, is the unique generic type in . It follows that a coset contains a unique generic type . The uniqueness of the generic type of implies that it is -invariant, and therefore definable over .
Finally, let be any stationary generic type over a small set. Then is the unique generic type in some coset . It follows that is -invariant, so . Thus . Conversely, let , and let , . Since must also be the right-stabiliser of we have as well, and , as desired. ∎
It follows that is connected (i.e., ) if and only if it has a unique generic type.
6.3. Global group ranks
We have seen that a type of a member of is generic if and only if the corresponding type in is a non forking extension of the unique type over , i.e., if its -type has the same Cantor-Bendixson ranks as all of for every . Thus the various --Cantor-Bendixson ranks play the role of stratified local ranks characterising genericity. In a superstable (and even more so in an -stable) theory one would expect a similar characterisation via global Lascar and/or Morley ranks. We shall consider here the case of superstability and Lascar ranks. Morley ranks are studied in a subsequent paper [Bena], and similar results are proved.
Significant work regarding superstability has been carried out in the context of metric Hausdorff cats, and in many cases definitions and proofs transfer verbatim to continuous logic. The objects of study in this context are partial types constructed from complete types and conditions of the form via conjunction and existential quantification. It is a general fact that if is a partial type then the property , where the existential quantifier is interpreted in a sufficiently saturated model, is definable by a partial type as well. This mostly happens in the following form. Let be a partial type and a real number. We define to be the partial type expressing that , i.e., that is realised in the -neighbourhood of . For a tuple we shall use as a notational representation for the somewhat vague concept of “ known up to distance ”, so in particular is just another representation for . Accordingly, if then we define , so in particular .
Definition 6.15.
To an arbitrary theory we define to be the least infinite cardinal, if such exists, such that for every complete type over a set , and for every , there is a subset such that does not divide over .
We say that is simple if , and that it is supersimple if .
It follows from our earlier results that every stable theory is simple. It is true (but we shall not require it) that if is not simple then .
Definition 6.16.
We say that is -stable if for every set , , the metric density character of is at most . We define to be the least (infinite) cardinal of stability for . We say that is superstable if it is -stable for all big enough.
In the context of Henson’s logic for Banach space, this definition dates back to Iovino [Iov99]. It was shown in [BU] that is stable if and only if it is -stable for all . In particular, is stable if and only if , in which case . The following is an example for a result whose statement and proof translate word-for-word to the continuous logic setting.
Fact 6.17 ([Ben05, Theorem 4.13]).
A theory is -stable if and only if .
Corollary 6.18.
A theory is superstable if and only if it is stable and supersimple.
Many of the arguments that follow are valid both for stable and for simple theories, and are stated as such, even though no development of simplicity theory for continuous logic exists in the literature. The reader may either refer to the development of simplicity in the context of cats [Ben03a, Ben03b], which encompasses that of continuous logic, or simply restrict his or her attention to the stable case.
In light of Corollary 6.18 and of the definition of we wish to define global forking ranks for such that if and only if “-depends” on over . We have quite a bit of liberty in choosing what “-depends” should mean (i.e., different choices can give rise to ranks which have all the properties we seek). For example, we could say that this happens if forks over , or if forks over . For consistency with earlier work we shall opt for a slightly more complex definition which was given in [Ben06] and which has some advantages over other definitions for the purposes of that paper.
Definition 6.19.
Let and be tuples, a set and . We keep in mind that represents “ known up to distance ”.
-
(i)
We say that an indiscernible sequence could be in if there are and a sequence such that is -indiscernible, and .
-
(ii)
We say that if every indiscernible sequence in could be in .
-
(iii)
We define as may be expected: if and only if there is such that and . (In [Ben06] the notation was used.)
A few justifications for these definitions may be in place. Indeed, for it is easy to see that a sequence could be in if and only if it admits a conjugate which is in in the ordinary sense. By a compactness argument, this is further equivalent to the property that could be in for every .
Next, we wish to justify the notation . The result in [Ben06] on which this justification relies does not immediately make sense in continuous logic, since at some point it uses a local ranks argument which is specific to the formalism of compact abstract theories. The core of that argument resides in the following technical result, which indeed can be proved very quickly in many different contexts (simple or stable theories, classical logic, continuous logic, cats) using the appropriate local ranks. Since the notion of local forking ranks varies drastically between contexts we shall provide here a more combinatorial and therefore more universal argument.
Lemma 6.20.
Let , and be three tuples, possibly infinite, in a stable or even simple theory (classical, continuous, or any other setting in which basic simplicity or stability hold). Assume moreover that . Then if and only if .
Proof.
Assume not, say but . We construct by induction a sequence such that and for all . We start with , . Then for each we can choose such that and . Our induction hypothesis tells us that and . By standard independence calculus we obtain , whence , as desired.
Since , there exists a -indiscernible sequence , starting with , such that there exists no satisfying for all . For each we may choose a copy such that , so in particular this copy starts with and is indiscernible over . Since , we may choose to be indiscernible over .
Let us now define for all and consider the sequence . Applying compactness, we can find for arbitrarily big a sequence in , such that for each there exists a sequence which is indiscernible over such that , so in particular . Each of these sequences witnesses that , contradicting the local character. ∎
We can now prove:
Fact 6.21 ([Ben06, Lemma 1.13]).
Assume that is simple (or stable). For , and , the following imply one another from top to bottom:
-
(i)
.
-
(ii)
There is a Morley sequence for over which could be in .
-
(iii)
does not fork over .
-
(iv)
.
Proof.
We only need to provide a proof for (ii) (iii), for which the original proof used local ranks.
Indeed let and witness that a Morley sequence could be in . We may assume that is indiscernible over , and we may further assume that for some , the pair continues the sequence indiscernibly over . Standard arguments regarding indiscernibility provide that . Since we obtain . By Lemma 6.20 we have and thus . Let be an automorphism sending to . Then and . Thus does not fork over , as desired. ∎
Thus in particular if and only if does not fork over , i.e., if and only if , justifying our notation. By earlier observations, this is further equivalent to holding for all . It is further shown in [Ben06] that is supersimple if and only if is ordinal for every finite tuple and . Moreover, in a supersimple theory , ranks characterise independence: if and only if for all .
Since our global ranks depend (inevitably) on a metric resolution parameter we may only hope to characterise genericity in case the metric is invariant under the group action, i.e., if the action of each on is an isometry.
We have seen that if is generic over then is generic over . We now prove a converse:
Lemma 6.22.
Assume is a type-definable transitive group action in a stable theory , generic over a set , satisfying . Then there is , such that . Moreover, can be chosen generic over (i.e., over ).
Proof.
We may assume . First choose generic, . Then is generic over , so By standard independence calculus we obtain . Since the action is transitive we can find such that , and we may take it so that . Then is generic over , and so is , and in particular . Then is generic over as required. ∎
Theorem 6.23.
Assume is a type-definable transitive group action with an invariant metric in a superstable continuous theory , . Then is generic if and only if for all . In particular, types of maximal -rank exist.
Proof.
We may assume that . We shall use the fact that if , and is -definable and isometric then for all . The proof of this fact is left as an exercise to the reader.
Let , and assume first that is generic. Let realise an arbitrary type over . We may nonetheless assume that . By the Lemma there exists such that . Since multiplication by is isometric we obtain .
Conversely, let and assume that for all and all . Let , . Then . Thus equality holds all the way for all , whereby , so is generic. ∎
References
- [Bena] Itaï Ben Yaacov, Definability of groups in -stable metric structures, Journal of Symbolic Logic, to appear, arXiv:0802.4286.
- [Benb] by same author, On theories of random variables, submitted, arXiv:0901.1584.
- [Ben03a] by same author, Simplicity in compact abstract theories, Journal of Mathematical Logic 3 (2003), no. 2, 163–191.
- [Ben03b] by same author, Thickness, and a categoric view of type-space functors, Fundamenta Mathematicæ 179 (2003), 199–224.
- [Ben05] by same author, Uncountable dense categoricity in cats, Journal of Symbolic Logic 70 (2005), no. 3, 829–860.
- [Ben06] by same author, On supersimplicity and lovely pairs of cats, Journal of Symbolic Logic 71 (2006), no. 3, 763–776, arXiv:0902.0118.
- [Ben09] by same author, Modular functionals and perturbations of Nakano spaces, Journal of Logic and Analysis 1:1 (2009), 1–42, arXiv:0802.4285.
- [BU] Itaï Ben Yaacov and Alexander Usvyatsov, Continuous first order logic and local stability, Transactions of the American Mathematical Society, to appear, arXiv:0801.4303.
- [Iov99] José Iovino, Stable Banach spaces and Banach space structures, I and II, Models, algebras, and proofs (Bogotá, 1995), Lecture Notes in Pure and Appl. Math., vol. 203, Dekker, New York, 1999, pp. 77–117.
- [Pil96] Anand Pillay, Geometric stability theory, Oxford Logic Guides, vol. 32, The Clarendon Press Oxford University Press, New York, 1996, Oxford Science Publications.
- [Poi85] Bruno Poizat, Cours de théorie des modèles, Nur al-Mantiq wal-Ma’rifah, 1985.