This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Stability and stable groups in continuous logic

Itaï Ben Yaacov Itaï Ben Yaacov
Université Claude Bernard – Lyon 1
Institut Camille Jordan, CNRS UMR 5208
43 boulevard du 11 novembre 1918
69622 Villeurbanne Cedex
France
http://math.univ-lyon1.fr/~begnac/
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 group
2000 Mathematics Subject Classification:
03C45 ; 03C60 ; 03C90
Author supported by ANR chaire d’excellence junior THEMODMET (ANR-06-CEXC-007) and by Marie Curie research network ModNet
Revision of 1st September 2025

Introduction

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 φ\varphi-types for a stable formula φ\varphi, the properties of φ\varphi-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 TT denotes a continuous theory in a language \mathcal{L}. We do not assume that TT is complete, so various constants, such as k(φ,ε)k(\varphi,\varepsilon) of Fact 2.1, are uniform across all completions of TT (provided that φ\varphi is stable in TT, i.e., in every completion of TT separately).

By a model we always mean a model of TT. Whenever this is convenient, we shall assume that such a model \mathcal{M} is embedded elementarily in a large monster model 𝔐\mathfrak{M}, i.e., in a strongly κ\kappa-homogeneous and saturated model, where κ\kappa 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 TT, as this would consist of choosing one completion.

1. General reminders

We shall consider throughout a formula φ(x¯,y¯)\varphi(\bar{x},\bar{y}) whose variables are split in two groups. We recall from [BU] that a definable φ\varphi-predicate is a definable predicate ψ(x¯\psi(\bar{x}), possibly with parameters, which is equivalent to an infinitary continuous combination of instances of φ\varphi:

ψ(x¯)θ(φ(x¯,b¯n))n,θ:[0,1][0,1] continuous.\displaystyle\psi(\bar{x})\equiv\theta\bigl{(}\varphi(\bar{x},\bar{b}_{n})\bigr{)}_{n\in\mathbb{N}},\qquad\theta\colon[0,1]^{\mathbb{N}}\to[0,1]\text{ continuous}.

Equivalently, φ(x¯)\varphi(\bar{x}) is a φ\varphi-predicate if it can be approximated arbitrarily well by finite continuous combinations of instances of φ\varphi, possibly restricted to the use of the connectives ¬\neg, 12\frac{1}{2}, .\mathbin{\mathchoice{\kern 2.77774pt\hbox to0.0pt{\hss\hbox{$\displaystyle-$}\hss}\raise 2.58334pt\hbox to0.0pt{\hss$\displaystyle.$\hss}\kern 2.77774pt}{\kern 2.77774pt\hbox to0.0pt{\hss\hbox{$\textstyle-$}\hss}\raise 2.58334pt\hbox to0.0pt{\hss$\textstyle.$\hss}\kern 2.77774pt}{\kern 2.45831pt\hbox to0.0pt{\hss\hbox{$\scriptstyle-$}\hss}\raise 1.80835pt\hbox to0.0pt{\hss$\scriptstyle.$\hss}\kern 2.45831pt}{\kern 2.29166pt\hbox to0.0pt{\hss\hbox{$\scriptscriptstyle-$}\hss}\raise 1.29167pt\hbox to0.0pt{\hss$\scriptscriptstyle.$\hss}\kern 2.29166pt}} alone.

Local types, i.e., φ\varphi-types for a fixed formula φ\varphi, are discussed in [BU, Section 6]. For a model \mathcal{M} and a tuple a¯\bar{a} in some extension 𝒩\mathcal{N}\succeq\mathcal{M}, the φ\varphi-type of a¯\bar{a} over \mathcal{M}, denoted tpφ(a¯/M)\operatorname{tp}_{\varphi}(\bar{a}/M), is the partial type given by {φ(x¯,b¯)=φ(a¯,b¯)}b¯M\{\varphi(\bar{x},\bar{b})=\varphi(\bar{a},\bar{b})\}_{\bar{b}\in M}. The space of all φ\varphi-types over MM is denoted Sφ(M)\operatorname{S}_{\varphi}(M), and it is a compact Hausdorff quotient of Sn(M)\operatorname{S}_{n}(M). If ψ(x¯)\psi(\bar{x}) is a φ\varphi-predicate over MM then tpφ(a¯/M)\operatorname{tp}_{\varphi}(\bar{a}/M) determines ψ(a¯)\psi(\bar{a}), so we may identify ψ\psi with a mapping ψ^:Sφ(M)[0,1]\hat{\psi}\colon\operatorname{S}_{\varphi}(M)\to[0,1], sending pψpp\mapsto\psi^{p}. Every such mapping is continuous, and conversely, every continuous mapping from Sφ(M)\operatorname{S}_{\varphi}(M) to [0,1][0,1] is of this form.

For AMA\subseteq M we define Sφ(A)\operatorname{S}_{\varphi}(A) to be the quotient of Sφ(M)\operatorname{S}_{\varphi}(M) where two types are identified if all AA-definable φ\varphi-predicates agree on them. This is again a compact Hausdorff space, a common quotient of Sφ(M)\operatorname{S}_{\varphi}(M) and of Sk(A)\operatorname{S}_{k}(A) (for the appropriate kk), and the continuous mappings Sφ(A)[0,1]\operatorname{S}_{\varphi}(A)\to[0,1] are precisely the AA-definable φ\varphi-predicates. In particular this does not depend on the choice of \mathcal{M}.

Lemma 1.1.

Let \mathcal{M} be a structure, KMK\subseteq M^{\ell} a (metrically) compact set and let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a formula (or a definable predicate, which we may always name by a new predicate symbol without adding any structure). Then infy¯Kφ(x¯,y¯)\inf_{\bar{y}\in K}\varphi(\bar{x},\bar{y}) is a φ\varphi-predicate (with parameters in KK) and for any tuple x¯\bar{x}, the infimum is attained by some y¯K\bar{y}\in K.

In particular, KK is definable in \mathcal{M}.

Proof.

Since KK is compact we can find a sequence {c¯n}nK\{\bar{c}_{n}\}_{n\in\mathbb{N}}\subseteq K such that for every ε>0\varepsilon>0 there is m=m(ε)m=m(\varepsilon) such that Kn<mB(c¯n,ε)K\subseteq\bigcup_{n<m}B(\bar{c}_{n},\varepsilon). Then infy¯Kφ(x¯,y¯)\operatorname{inf}_{\bar{y}\in K}\varphi(\bar{x},\bar{y}) is arbitrarily well approximated by formulae of the form n<mφ(x¯,c¯n)\bigwedge_{n<m}\varphi(\bar{x},\bar{c}_{n}) as mm\to\infty. 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 \mathcal{M} be a model, AMA\subseteq M a subset. We say that \mathcal{M} is saturated over AA if it is strongly (|A|+0)+(|A|+\aleph_{0})^{+}-homogeneous and saturated. (In fact, for all intents and purposes it will suffice to require \mathcal{M} to be strongly 1\aleph_{1}-homogeneous and saturated once every member of AA is named.)

We say that a partial type π(x¯)\pi(\bar{x}) over \mathcal{M} is AA-invariant if \mathcal{M} is saturated over AA and π\pi is fixed by the action of Aut(/A)\operatorname{Aut}(\mathcal{M}/A).

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 AA, let us denote it by φ(x¯,A)\varphi(\bar{x},A). Then we may assume that AA is countable, say A=(an)nA=(a_{n})_{n\in\mathbb{N}}, and express φ(x¯,A)\varphi(\bar{x},A) as a uniform limit of formulae φn(x¯,a<n)\varphi_{n}(\bar{x},a_{<n}). Furthermore, using a forced limit argument, we may assume that the sequence φn(x¯,y<n)\varphi_{n}(\bar{x},y_{<n}) converges uniformly to some infinitary definable predicate φ(x¯,Y)\varphi(\bar{x},Y), giving sense to φ(x¯,B)\varphi(\bar{x},B) for any sequence B=(bn)nB=(b_{n})_{n}. If \mathcal{M} is any structure, we equip MM^{\mathbb{N}} with the pseudo-metric dφ(B,C)=supx¯|φ(x¯,B)φ(x¯,C)|d_{\varphi}(B,C)=\sup_{\bar{x}}\,|\varphi(\bar{x},B)-\varphi(\bar{x},C)| and define the sort of canonical parameters for φ\varphi in \mathcal{M}, denoted SφS_{\varphi}^{\mathcal{M}}, as the complete metric space associated to (M,dφ)(M^{\mathbb{N}},d_{\varphi}). In other words, we divide MM^{\mathbb{N}} by the kernel dφ(Y,Z)=0d_{\varphi}(Y,Z)=0, obtaining a true metric on the quotient, and pass to the completion. The predicate φ(x¯,Y)\varphi(\bar{x},Y) is uniformly continuous with respect to YY in the metric dφd_{\varphi}, and therefore passes first to the quotient and then to the completion, thus inducing a uniformly continuous predicate Pφ(x¯,z)P_{\varphi}(\bar{x},z), where zSφz\in S_{\varphi}. We may now add (Sφ,dφ)(S_{\varphi},d_{\varphi}) as a sort to \mathcal{M} and equip the new structure with an additional predicate symbol for PφP_{\varphi}. This does not add structure to the original sorts of \mathcal{M}, elementary embeddings of structures commute with this construction (by which we mean, in particular, that an elementary embedding 𝒩\mathcal{M}\preceq\mathcal{N} extends uniquely to (,Sφ)(𝒩,Sφ𝒩)(\mathcal{M},S_{\varphi}^{\mathcal{M}})\preceq(\mathcal{N},S_{\varphi}^{\mathcal{N}})), elementary classes and model completeness thereof are respected by this construction, and so. The construction can be slightly simplified when φ\varphi 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 c=[A]c=[A] is the image of AA in SφS_{\varphi} then cc is indeed a canonical parameter for φ(x¯,A)\varphi(\bar{x},A), in the sense that an automorphism of \mathcal{M} or of an elementary extension thereof (and such an automorphism extends uniquely to SφS_{\varphi}) fixes cc if and only if it fixes the predicate φ(x¯,A)\varphi(\bar{x},A). By construction we have Pφ(x¯,c)=φ(x¯,A)P_{\varphi}(\bar{x},c)=\varphi(\bar{x},A), and by a convenient abuse of notation we shall permit ourselves to write φ(x¯,c)\varphi(\bar{x},c) 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 ψ(x¯,Y)\psi(\bar{x}^{\prime},Y^{\prime}), 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 SdφS_{d\varphi} for each formula φ\varphi (see Fact 2.1 for the predicate dφd\varphi). Therefore, with some intentional ambiguity, by eq\mathcal{M}^{eq} we shall usually mean “\mathcal{M} along with all the imaginary sorts we are going to need”, e.g., all the sorts SdφS_{d\varphi}. By dcleq\operatorname{dcl}^{eq}, acleq\operatorname{acl}^{eq}, etc., we mean the respective operations in the structure eq\mathcal{M}^{eq}.

Lemma 1.3.

Let AA be a set of parameters and let φ(x¯)\varphi(\bar{x}) be a definable predicate with parameters possibly outside AA. Then φ\varphi is AA-definable if and only if it is AA-invariant, i.e., if and only if its canonical base belongs to dcleq(A)\operatorname{dcl}^{eq}(A).

Similarly, a set which is definable (type-definable) with some parameters is definable (type-definable) over AA if and only it is AA-invariant.

Proof.

First, let us consider an arbitrary surjective continuous mapping π:XY\pi\colon X\to Y between two compact Hausdorff topological spaces. Then π\pi is also closed, so FYF\subseteq Y is closed if and only if π1(F)\pi^{-1}(F) is closed in XX. Since π\pi is surjective, UYU\subseteq Y is open if and only if π1(U)\pi^{-1}(U) is open, and a mapping f:Y[0,1]f\colon Y\to[0,1] is continuous if and only if fπf\circ\pi is continuous.

The assertion now follows from applying the previous paragraph to the restriction mapping Sn(B)Sn(A)\operatorname{S}_{n}(B)\to\operatorname{S}_{n}(A), where BAB\supseteq A 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 XX, just argue for the definable predicate d(x¯,X)d(\bar{x},X). ∎

Fact 1.4.

[BU, Lemma 6.8] Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be any formula, AA a set, \mathcal{M} a saturated model over AA, and let pSφ(A)p\in\operatorname{S}_{\varphi}(A). Then Aut(/A)\operatorname{Aut}(\mathcal{M}/A) acts transitively on the set of extensions of pp in Sφ(acleq(A))\operatorname{S}_{\varphi}(\operatorname{acl}^{eq}(A)).

The following notion and fact also appear (and are used much more extensively) in [Bena, Section 1]:

Definition 1.5.

Let XX and YY be two type-definable sets. We say that YY is a logical neighbourhood of XX, in symbols X<YX<Y, if there is a set of parameters AA over which both XX and YY are defined such that [X][Y][X]\subseteq[Y]^{\circ} in Sn(A)\operatorname{S}_{n}(A).

Notice that the interior of [Y][Y] does depend on AA (i.e., if AAA^{\prime}\supseteq A then [Y][Y]^{\circ} calculated in Sn(A)\operatorname{S}_{n}(A^{\prime}) may be larger than the pullback of the interior of [Y][Y] in Sn(A)\operatorname{S}_{n}(A)). We may nonetheless choose any parameter set we wish:

Lemma 1.6.

Assume that XX is type-definable with parameters in BB, YY type-definable possibly with additional parameters not in BB. Then:

  1. (i)

    If X<YX<Y then [X][Y][X]\subseteq[Y]^{\circ} in Sn(A)\operatorname{S}_{n}(A) for any set AA over which both XX and YY are defined.

  2. (ii)

    If X<YX<Y then there is an intermediate logical neighbourhood X<Z<YX<Z<Y, which can moreover be taken to be the zero set of a formula with parameters in BB.

  3. (iii)

    If YX=Y\cap X=\varnothing then there is a logical neighbourhood Z>XZ>X such that ZY=Z\cap Y=\varnothing. Moreover, we may take ZZ to be a zero set defined over BB.

Proof.

Assume X<YX<Y, where XX is type-definable over BB, and YY over ABA\supseteq B. Let Φ\Phi consist of all formulae φ(x¯)\varphi(\bar{x}) over BB which are zero on XX. If φ,ψΦ\varphi,\psi\in\Phi then φψΦ\varphi\vee\psi\in\Phi, and XX is defined by the partial type p(x¯)={φ(x¯)r:φΦ,r>0}p(\bar{x})=\{\varphi(\bar{x})\leq r\colon\varphi\in\Phi,r>0\}. By compactness in Sn(A)\operatorname{S}_{n}(A) there is a condition φ(x¯)r\varphi(\bar{x})\leq r in p(x¯)p(\bar{x}) which already implies x¯Y\bar{x}\in Y. Let ZZ be the zero set of the formula φ(x¯).r\varphi(\bar{x})\mathbin{\mathchoice{\kern 2.77774pt\hbox to0.0pt{\hss\hbox{$\displaystyle-$}\hss}\raise 2.58334pt\hbox to0.0pt{\hss$\displaystyle.$\hss}\kern 2.77774pt}{\kern 2.77774pt\hbox to0.0pt{\hss\hbox{$\textstyle-$}\hss}\raise 2.58334pt\hbox to0.0pt{\hss$\textstyle.$\hss}\kern 2.77774pt}{\kern 2.45831pt\hbox to0.0pt{\hss\hbox{$\scriptstyle-$}\hss}\raise 1.80835pt\hbox to0.0pt{\hss$\scriptstyle.$\hss}\kern 2.45831pt}{\kern 2.29166pt\hbox to0.0pt{\hss\hbox{$\scriptscriptstyle-$}\hss}\raise 1.29167pt\hbox to0.0pt{\hss$\scriptscriptstyle.$\hss}\kern 2.29166pt}}r^{\prime} where 0<r=k2m<r0<r^{\prime}=\frac{k}{2^{-m}}<r.

Then in Sn(A)\operatorname{S}_{n}(A) we have [X][φ(x¯)<r][φ(x¯)r][φ(r¯)<r][Y][X]\subseteq[\varphi(\bar{x})<r^{\prime}]\subseteq[\varphi(\bar{x})\leq r^{\prime}]\subseteq[\varphi(\bar{r})<r]\subseteq[Y], i.e., [X][Z][Z][Y][X]\subseteq[Z]^{\circ}\subseteq[Z]\subseteq[Y]^{\circ}, proving the first two items. The third item now follows from the fact that Sn(A)\operatorname{S}_{n}(A) is a normal topological space. ∎

2. Definability and forking of local types

Having fixed a theory TT, we shall call here a formula φ(x¯,y¯)\varphi(\bar{x},\bar{y}) stable if it is stable in TT, that is, if it does not have the order property in any model of TT. 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 φ(x¯,y¯)\varphi(\bar{x},\bar{y}) is any formula with two groups of variables, φ~(y¯,x¯)\tilde{\varphi}(\bar{y},\bar{x}) denotes the same formula with the groups of variables interchanged. More generally, let us define

φ~n(y¯,x¯2n)=medn(φ(x¯i,y¯))i2n,\displaystyle\tilde{\varphi}^{n}(\bar{y},\bar{x}_{\leq 2n})=\operatorname{med}_{n}\bigl{(}\varphi(\bar{x}_{i},\bar{y})\bigr{)}_{i\leq 2n},

where medn:[0,1]2n+1[0,1]\operatorname{med}_{n}\colon[0,1]^{2n+1}\to[0,1] is the median value combination:

medn(t2n)=w[2n+1]n+1iwti=w[2n+1]n+1iwti.\displaystyle\operatorname{med}_{n}(t_{\leq 2n})=\bigwedge_{w\in[2n+1]^{n+1}}\,\bigvee_{i\in w}\,t_{i}=\bigvee_{w\in[2n+1]^{n+1}}\,\bigwedge_{i\in w}\,t_{i}.

Thus in particular φ~0=φ~\tilde{\varphi}^{0}=\tilde{\varphi} and every instance of φ~n\tilde{\varphi}^{n} is a φ~\tilde{\varphi}-predicate.

Fact 2.1.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula. Let \mathcal{M} be a model and let pSφ()p\in\operatorname{S}_{\varphi}(\mathcal{M}) be a complete φ\varphi-type. Then

  1. (i)

    The type pp is definable over MM, i.e., there exists an MM-definable φ~\tilde{\varphi}-predicate dpφ(y¯)d_{p}\varphi(\bar{y}) such that φ(x,b¯)p=dpφ(b¯)\varphi(x,\bar{b})^{p}=d_{p}\varphi(\bar{b}) for all b¯M\bar{b}\in M. Moreover, this definition is uniform, in the sense that there exists an infinitary definable predicate dφ(y¯,X)d\varphi(\bar{y},X) which only depends on φ\varphi such that dpφ(y¯)d_{p}\varphi(\bar{y}) is equal to an instance dφ(y¯,C)d\varphi(\bar{y},C) where CMC\subseteq M.

  2. (ii)

    For every ε>0\varepsilon>0 there exists a number k=k(φ,ε)k=k(\varphi,\varepsilon)\in\mathbb{N} (which depends on φ\varphi and on ε\varepsilon but not on pp) and a tuple c¯2k=c¯2k(φ,ε)ε\bar{c}_{\leq 2k}=\bar{c}^{\varepsilon}_{\leq 2k(\varphi,\varepsilon)} in \mathcal{M} (which does depend on pp) such that

    |dpφ(y¯)φ~k(y¯,c¯2k)|<ε.\displaystyle|d_{p}\varphi(\bar{y})-\tilde{\varphi}^{k}(\bar{y},\bar{c}_{\leq 2k})|<\varepsilon.
  3. (iii)

    Assume moreover that \mathcal{M} is saturated over some subset AMA\subseteq M. Then in the previous item the tuples c¯2k\bar{c}_{\leq 2k} can be chosen so that each c¯n\bar{c}_{n} realises pAc¯<np{\restriction}_{A\bar{c}_{<n}}.

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 dpφ(y¯)=dφ(y¯,c)d_{p}\varphi(\bar{y})=d\varphi(\bar{y},c), where cMeqc\in M^{eq} is the canonical parameter for the definition. This canonical parameter is called the canonical base of pp, denoted Cb(p)\mathrm{Cb}(p).

We recall that for ABA\subseteq B\subseteq\mathcal{M}, pSφ(B)p\in\operatorname{S}_{\varphi}(B) does not fork over AA if it admits an extension p1Sφ(M)p_{1}\in\operatorname{S}_{\varphi}(M) which is definable over acleq(A)\operatorname{acl}^{eq}(A). In this case p1p_{1} itself does not fork over AA or BB. 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 \mathcal{M}.

We proved in [BU, Proposition 7.15] that every φ\varphi-type over a set AA admits a non forking extension to every model (and therefore every set) containing AA. A minor enhancement of that result will be quite useful.

Lemma 2.2 (Existence of non forking extensions).

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula, AA a set, A\mathcal{M}\supseteq A a saturated model over AA. Let π(x¯)\pi(\bar{x}) be a consistent AA-invariant partial type over MM. Then there exists pSφ(M)p\in\operatorname{S}_{\varphi}(M) compatible with π\pi which does not fork over AA.

Proof.

Let X={pSφ(M):pπ is consistent}X=\{p\in\operatorname{S}_{\varphi}(M)\colon p\cup\pi\text{ is consistent}\}. Then XX is non empty and AA-invariant. By [BU, Lemma 7.14], there is YXY\subseteq X which is AA-good, i.e., which is AA-invariant and metrically compact. By [BU, Lemma 7.13], any pYp\in Y would do. ∎

Corollary 2.3.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula, AA a set, A\mathcal{M}\supseteq A a saturated model over AA. Then pSφ(M)p\in\operatorname{S}_{\varphi}(M) does not fork over AA if and only if it is acleq(A)\operatorname{acl}^{eq}(A)-invariant.

Proof.

Left to right follows from the definition, right to left from Lemma 2.2. ∎

Corollary 2.4.

Let AA be a set, A\mathcal{M}\supseteq A a saturated model over AA and π(x¯)\pi(\bar{x}) a consistent AA-invariant partial type over MM. Then there exists a complete type pp such that πpSn(M)\pi\subseteq p\in\operatorname{S}_{n}(M), and for every stable formula φ(x¯,y¯)\varphi(\bar{x},\bar{y}) the restriction pφSφ(M)p{\restriction}_{\varphi}\in\operatorname{S}_{\varphi}(M) does not fork over AA.

Proof.

We may assume that A=acleq(A)A=\operatorname{acl}^{eq}(A). Index all stable formulae of the form φi(x¯,y¯i)\varphi_{i}(\bar{x},\bar{y}_{i}) by i<λi<\lambda. We define an increasing sequence of consistent AA-invariant partial types πi\pi_{i} over MM, starting with π0=π\pi_{0}=\pi. Given πi\pi_{i}, by Lemma 2.2 there is piSφi(M)p_{i}\in\operatorname{S}_{\varphi_{i}}(M) be non forking over AA and compatible with πi\pi_{i}, so πi+1=πipi\pi_{i+1}=\pi_{i}\cup p_{i} is consistent and AA-invariant. For limit ii we define πi=j<iπj\pi_{i}=\bigcup_{j<i}\pi_{j}. Finally, let pSn(M)p\in\operatorname{S}_{n}(M) be any completion of πλ\pi_{\lambda}. Then pp 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 \mathcal{M} be a model, p(x¯)Sφ(M)p(\bar{x})\in\operatorname{S}_{\varphi}(M), q(y¯)Sφ~(M)q(\bar{y})\in\operatorname{S}_{\tilde{\varphi}}(M). Then dpφ(y¯)q=dqφ~(x¯)pd_{p}\varphi(\bar{y})^{q}=d_{q}\tilde{\varphi}(\bar{x})^{p}.

Proposition 2.6.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula, \mathcal{M} a model, AMA\subseteq M. For each b¯M\bar{b}\in M let χb¯(x¯)\chi_{\bar{b}}(\bar{x}) be the definition of a non forking extension of tpφ~(b¯/acleq(A))\operatorname{tp}_{\tilde{\varphi}}(\bar{b}/\operatorname{acl}^{eq}(A)) to MM.

  1. (i)

    Each χb¯(x¯)\chi_{\bar{b}}(\bar{x}) is a definable φ\varphi-predicate over acleq(A)\operatorname{acl}^{eq}(A).

  2. (ii)

    A φ\varphi-type pSφ(M)p\in\operatorname{S}_{\varphi}(M) does not fork over AA if and only if φ(x¯,b¯)p=χb¯(x¯)p\varphi(\bar{x},\bar{b})^{p}=\chi_{\bar{b}}(\bar{x})^{p} for all b¯M\bar{b}\in M.

  3. (iii)

    A φ\varphi-type over acleq(A)\operatorname{acl}^{eq}(A) is stationary, i.e., admits a unique non forking extension to every larger set.

  4. (iv)

    Let r(x¯)={|φ(x¯,b¯)χb¯(x¯)|=0}b¯Mr(\bar{x})=\big{\{}|\varphi(\bar{x},\bar{b})-\chi_{\bar{b}}(\bar{x})|=0\big{\}}_{\bar{b}\in M}. Then the partial type r(x¯)r(\bar{x}) defines the set of φ\varphi-types which do not fork over AA:

    a¯rtpφ(a¯/M) does not fork over A.\displaystyle\bar{a}\vDash r\quad\Longleftrightarrow\quad\operatorname{tp}_{\varphi}(\bar{a}/M)\text{ does not fork over }A.
  5. (v)

    For every BAB\supseteq A, the set {pSφ(B):p does not fork over A}\{p\in\operatorname{S}_{\varphi}(B)\colon p\text{ does not fork over }A\} is closed.

Proof.

The first item is by Fact 2.1 and the definition of non forking.

For the second, fix b¯M\bar{b}\in M, let q0=tpφ~(b¯/acleq(A))q_{0}=\operatorname{tp}_{\tilde{\varphi}}(\bar{b}/\operatorname{acl}^{eq}(A)) and let qSφ(M)q\in\operatorname{S}_{\varphi}(M) be the non forking extension defined by χb¯\chi_{\bar{b}}. Assume pSφ(M)p\in\operatorname{S}_{\varphi}(M) does not fork over MM, so dpφ(y¯)d_{p}\varphi(\bar{y}) is a φ~\tilde{\varphi}-predicate over acleq(A)\operatorname{acl}^{eq}(A). By Fact 2.5,

φ(x¯,b¯)p=dpφ(b¯)=dpφ(y¯)q0=dpφ(y¯)q=dqφ~(x¯)p=χb¯(x¯)p.\displaystyle\varphi(\bar{x},\bar{b})^{p}=d_{p}\varphi(\bar{b})=d_{p}\varphi(\bar{y})^{q_{0}}=d_{p}\varphi(\bar{y})^{q}=d_{q}\tilde{\varphi}(\bar{x})^{p}=\chi_{\bar{b}}(\bar{x})^{p}.

Conversely, assume that φ(x¯,b¯)p=χb¯(x¯)p\varphi(\bar{x},\bar{b})^{p}=\chi_{\bar{b}}(\bar{x})^{p} for all b¯M\bar{b}\in M, and let pSφ(M)p^{\prime}\in\operatorname{S}_{\varphi}(M) be any non forking extension of pacleq(A)p{\restriction}_{\operatorname{acl}^{eq}(A)}. Then p=pp=p^{\prime}, proving also the third item. The fourth item is just a re-statement of the second.

For the last item we may assume that BMB\subseteq M. The set [r]Sφ(M)[r]\subseteq\operatorname{S}_{\varphi}(M) is closed, and so is its projection to Sφ(B)\operatorname{S}_{\varphi}(B). This projection is precisely the set of types which do not fork over AA. ∎

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 pSφ(A)p\in\operatorname{S}_{\varphi}(A) is stationary, the unique non forking extension to BAB\supseteq A will be denoted pBp{\restriction}^{B}. Similarly, we write dpφd_{p}\varphi for the definition of pMp{\restriction}^{M} where A\mathcal{M}\supseteq A is any model (and this does not depend on the choice of \mathcal{M}). Thus, in hindsight, in the statement of Proposition 2.6, the definitions χb¯\chi_{\bar{b}} are uniquely determined, χb¯=db¯/acleq(A)φ~\chi_{\bar{b}}=d_{\bar{b}/\operatorname{acl}^{eq}(A)}\tilde{\varphi}.

Corollary 2.7.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula, AA a set, \mathcal{M} a saturated model over AA. Let pSφ(A)p\in\operatorname{S}_{\varphi}(A). Then Aut(/A)\operatorname{Aut}(\mathcal{M}/A) acts transitively on the set of non forking extensions of pp in Sφ(M)\operatorname{S}_{\varphi}(M). If TT is stable and pSn(A)p\in\operatorname{S}_{n}(A) then Aut(/A)\operatorname{Aut}(\mathcal{M}/A) acts transitively on the set of non forking extensions of pp to \mathcal{M}.

Proof.

The first assertion follows from Fact 1.4 and Proposition 2.6.(iii). For the second we need the even easier fact that Aut(/A)\operatorname{Aut}(\mathcal{M}/A) acts transitively on the extensions of a complete type pSn(A)p\in\operatorname{S}_{n}(A) to acleq(A)\operatorname{acl}^{eq}(A). ∎

Corollary 2.8.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula \mathcal{M} a model. A type pSφ(M)p\in\operatorname{S}_{\varphi}(M) is definable over AA if and only if it does not fork over AA and pAp{\restriction}_{A} is stationary.

Proof.

We may assume that \mathcal{M} saturated over AA. Let pSφ(M)p^{\prime}\in\operatorname{S}_{\varphi}(M) be any non forking extension of pAp{\restriction}_{A}. By Corollary 2.7 there is an automorphism fAut(/A)f\in\operatorname{Aut}(\mathcal{M}/A) sending pp to pp^{\prime}. If pp is definable over AA then p=f(p)=pp^{\prime}=f(p)=p. Conversely, if pp does not fork over AA and pAp{\restriction}_{A} is stationary then Aut(/A)\operatorname{Aut}(\mathcal{M}/A) fixes pp and therefore fixes dpφd_{p}\varphi. By Lemma 1.3, the latter is over AA. ∎

Corollary 2.9.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula, AA a set, q(x¯)Sn(A)q(\bar{x})\in\operatorname{S}_{n}(A) a complete type over AA, and let p0=qφSφ(A)p_{0}=q{\restriction}_{\varphi}\in\operatorname{S}_{\varphi}(A). Then qq is compatible with every non forking extension of p0p_{0}.

Proof.

By Lemma 2.2, qq is compatible with at least one non forking extension of pp to the monster model. By Corollary 2.7 it is compatible with all of them. ∎

We pass to forking of single conditions.

Definition 2.10.

Let φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be an instance of a stable formula, AA a set. We say that a condition φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not fork over AA if there exists a φ\varphi-type pSφ(Ab¯)p\in\operatorname{S}_{\varphi}(A\bar{b}) non forking over AA such that φ(x¯,b¯)pr\varphi(\bar{x},\bar{b})^{p}\leq r.

Proposition 2.11.

Let φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be an instance of a stable formula, AA a set of parameters. Then the following are equivalent:

  1. (i)

    The condition φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not fork over AA.

  2. (ii)

    Every family of acleq(A)\operatorname{acl}^{eq}(A)-conjugates of φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r is consistent.

  3. (iii)

    For every set BA,b¯B\supseteq A,\bar{b} there exists a complete type pSn(B)p\in\operatorname{S}_{n}(B) such that pψp{\restriction}_{\psi} does not fork over AA for any stable formula ψ\psi (if TT is stable: pp does not fork over AA) and φ(x¯,b¯)pr\varphi(\bar{x},\bar{b})^{p}\leq r.

Proof.
  • (i) \Longrightarrow (ii).

    Let pp witness that φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not fork over AA. Then any non forking extension of pp to a large model is acleq(A)\operatorname{acl}^{eq}(A)-invariant.

  • (iii) \Longrightarrow (iv).

    We may assume that B=B=\mathcal{M} is saturated over AA. Let π\pi consist of all the acleq(A)\operatorname{acl}^{eq}(A)-conjugates of φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r in \mathcal{M}. It is consistent by assumption and acleq(A)\operatorname{acl}^{eq}(A)-invariant by construction so we may apply Corollary 2.4.

  • (v) \Longrightarrow (i).

    Immediate. ∎

We may define the non forking degree of φ(x¯,b¯)\varphi(\bar{x},\bar{b}) over AA to be

nf(φ(x¯,b¯)/A)=inf{r:φ(x¯,b¯)r does not fork over A}.\displaystyle\operatorname{nf}\bigl{(}\varphi(\bar{x},\bar{b})/A\bigr{)}=\inf\bigl{\{}r\colon\varphi(\bar{x},\bar{b})\leq r\text{ does not fork over }A\bigr{\}}.

An easy compactness argument shows that the infimum is attained and the condition φ(x¯,b¯)nf(φ(x¯,b¯)/A)\varphi(\bar{x},\bar{b})\leq\operatorname{nf}\bigl{(}\varphi(\bar{x},\bar{b})/A\bigr{)} does not fork over AA. In addition, by the existence of non-forking types we have nf(φ(x¯,b¯)/A)+nf(¬φ(x¯,b¯)/A)1\operatorname{nf}\bigl{(}\varphi(\bar{x},\bar{b})/A\bigr{)}+\operatorname{nf}\bigl{(}\neg\varphi(\bar{x},\bar{b})/A\bigr{)}\leq 1.

Definition 2.12.

A faithful continuous connective in α\alpha variables is a continuous function θ:[0,1]α[0,1]\theta\colon[0,1]^{\alpha}\to[0,1] satisfying infa¯θ(a¯)supa¯\inf\bar{a}\leq\theta(\bar{a})\leq\sup\bar{a}.

If θ:[0,1]α[0,1]\theta\colon[0,1]^{\alpha}\to[0,1] is a faithful continuous connective and (φi)i<α(\varphi_{i})_{i<\alpha} a sequence of definable predicates, then the definable predicate θ(φi)i<α\theta(\varphi_{i})_{i<\alpha} is called a faithful combination of (φi)i<α(\varphi_{i})_{i<\alpha}.

Since a continuous function to [0,1][0,1] can only take into account countably many arguments, we may always assume that αω\alpha\leq\omega. Notice that any connective constructed using \vee and \wedge alone is faithful (so in particular the median value connective medn:[0,1]2n+1[0,1]\operatorname{med}_{n}\colon[0,1]^{2n+1}\to[0,1] is). Similarly, any uniform limit of faithful combinations is faithful.

Lemma 2.13.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula. Let A=acleq(A)A=\operatorname{acl}^{eq}(A) be a set of parameters, a¯\bar{a} a tuple, |x¯|=|a¯||\bar{x}|=|\bar{a}|. Let p=tpφ(a¯/A)p=\operatorname{tp}_{\varphi}(\bar{a}/A). Then dpφ(x¯,y¯)d_{p}\varphi(\bar{x},\bar{y}) is a faithful combination of AA-conjugates of φ(a¯,y¯)\varphi(\bar{a},\bar{y}).

Proof.

By the preceding discussion and the last item of Fact 2.1. ∎

Lemma 2.14.

Let φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be an instance of a stable formula, AA a set of parameters. Then there exists an AA-definable predicate ψ(x¯)\psi(\bar{x}) such that for every tuple a¯\bar{a} (not necessarily in AA):

ψ(a¯)\displaystyle\psi(\bar{a}) =inf{φ(x¯,b¯)p:pSφ(Ab¯) is a non forking extension of tpφ(a¯/A)}\displaystyle=\inf\{\varphi(\bar{x},\bar{b})^{p}\colon p\in\operatorname{S}_{\varphi}(A\bar{b})\text{ is a non forking extension of }\operatorname{tp}_{\varphi}(\bar{a}/A)\}
=inf{φ(a¯,y¯)q:qSφ(Aa¯) is a non forking extension of tpφ~(b¯/A)}.\displaystyle=\inf\{\varphi(\bar{a},\bar{y})^{q}\colon q\in\operatorname{S}_{\varphi}(A\bar{a})\text{ is a non forking extension of }\operatorname{tp}_{\tilde{\varphi}}(\bar{b}/A)\}.

Moreover, ψ(x¯)\psi(\bar{x}) can be taken to be a faithful combination of AA-conjugates of φ(x¯,b¯)\varphi(\bar{x},\bar{b}).

Proof.

Fix a model A,b¯\mathcal{M}\supseteq A,\bar{b}, saturated over AA. Let G=Aut(/A)G=\operatorname{Aut}(\mathcal{M}/A). Let qSφ~(M)q\in\operatorname{S}_{\tilde{\varphi}}(M) be the unique non forking extension of tpφ~(b¯/acleq(A))\operatorname{tp}_{\tilde{\varphi}}(\bar{b}/\operatorname{acl}^{eq}(A)). Let χ(x¯,c)=dqφ~(x¯)\chi(\bar{x},c)=d_{q}\tilde{\varphi}(\bar{x}), where cacleq(A)c\in\operatorname{acl}^{eq}(A) is the canonical parameter for the definition. By the previous Lemma, χ(x¯,c)\chi(\bar{x},c) is a faithful combination of acleq(A)\operatorname{acl}^{eq}(A)-conjugates of φ(x¯,b¯)\varphi(\bar{x},\bar{b}).

Let CC be the set of AA-conjugates of cc. Since cc is algebraic over AA, CC is (metrically) compact. By Lemma 1.1 ψ(x¯)=infcCχ(x¯,c)\psi(\bar{x})=\inf_{c^{\prime}\in C}\chi(\bar{x},c^{\prime}) is a continuous combination of instances χ(x¯,c)\chi(\bar{x},c^{\prime}) with cCc^{\prime}\in C, i.e., of AA-conjugates of χ(x¯,c)\chi(\bar{x},c), and it is clearly a faithful combination. Thus ψ(x¯)\psi(\bar{x}) is a faithful combination of AA-conjugates of φ(x¯,b¯)\varphi(\bar{x},\bar{b}), and it is clearly over AA.

We may assume that a¯M\bar{a}\in M, and let pSφ(M)p\in\operatorname{S}_{\varphi}(M) be the unique non forking extension of tpφ(a¯/acleq(A))\operatorname{tp}_{\varphi}(\bar{a}/\operatorname{acl}^{eq}(A)). Then

ψ(a¯)\displaystyle\psi(\bar{a}) =infgGχ(a¯,gc)=infgGdgqφ~(a¯)=\displaystyle=\inf_{g\in G}\chi(\bar{a},gc)=\inf_{g\in G}d_{gq}\tilde{\varphi}(\bar{a})=\ldots
=infgGdg1pφ(y¯)q=infgGφ(x¯,b¯)gp,\displaystyle\ldots=\inf_{g\in G}d_{g^{-1}p}\varphi(\bar{y})^{q}=\inf_{g\in G}\varphi(\bar{x},\bar{b})^{gp},
=infgGφ(a¯,y¯)gq.\displaystyle\ldots=\inf_{g\in G}\varphi(\bar{a},\bar{y})^{gq}.

Since {gp}gG\{gp\}_{g\in G} and {gq}gG\{gq\}_{g\in G} are the sets of non forking extensions of tpφ(a¯/A)\operatorname{tp}_{\varphi}(\bar{a}/A) and of tpφ~(b¯/A)\operatorname{tp}_{\tilde{\varphi}}(\bar{b}/A), respectively, to MM, we are done. ∎

Theorem 2.15 (Open Mapping Theorem).

Assume TT is stable, and let ABA\subseteq B be any sets of parameters. Let XSn(B)X\subseteq\operatorname{S}_{n}(B) be the set of types which do not fork over AA. Then XX is compact and the restriction mapping ρA:XSn(A)\rho_{A}\colon X\to\operatorname{S}_{n}(A) sending ppAp\mapsto p{\restriction}_{A} is an open continuous surjective mapping.

Proof.

We already know that XX is compact and that ρA\rho_{A} is continuous and surjective.

Consider a basic open subset UXU\subseteq X, of the form U=X[φ(x¯,b¯)<1]U=X\cap[\varphi(\bar{x},\bar{b})<1]. Let ψ(x¯)\psi(\bar{x}) be as in Lemma 2.14 and let V=[ψ(x¯)<1]Sn(A)V=[\psi(\bar{x})<1]\subseteq\operatorname{S}_{n}(A). By Corollary 2.4 every φ\varphi-type over BB which does not fork over AA extends to a complete type over BB which does not fork over AA, whence V=ρA(U)V=\rho_{A}(U). ∎

Notice that a similar proof yields that if φ(x¯,y¯)\varphi(\bar{x},\bar{y}) is stable then the restriction mapping ρA,φ:XφSφ(A)\rho_{A,\varphi}\colon X_{\varphi}\to\operatorname{S}_{\varphi}(A) is open, where XφSφ(B)X_{\varphi}\subseteq\operatorname{S}_{\varphi}(B) denotes the set of φ\varphi-types which do not fork over AA.

It follows from Lemma 2.14 that a φ~\tilde{\varphi}-type (and therefore a φ\varphi-type) over an arbitrary set AA is definable over AA, 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 φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be an instance of a stable formula, and let XX be a type-definable set in the sort of x¯\bar{x}, say with parameters in AA. Then there is a subset (at most countable) BXB\subseteq X and a BB-definable predicate ψ(x¯)\psi(\bar{x}) such that ψ(x¯)X=φ(x¯,b¯)X\psi(\bar{x}){\restriction}_{X}=\varphi(\bar{x},\bar{b}){\restriction}_{X}.

Moreover, ψ(x¯)\psi(\bar{x}) can be taken to be a faithful combination of instances φ(x¯,b¯)\varphi(\bar{x},\bar{b}^{\prime}) such that b¯Bb¯\bar{b}^{\prime}\equiv_{B}\bar{b} (or even b¯Bb¯\bar{b}^{\prime}\equiv_{B^{\prime}}\bar{b} where BXB^{\prime}\subseteq X is an arbitrary small subset).

Proof.

Fix a model A,b¯\mathcal{M}\supseteq A,\bar{b}, saturated over AA, and let C=X()C=X(\mathcal{M}). Let ψ(x¯)\psi(\bar{x}) be as in Lemma 2.14. Then ψ(x¯)\psi(\bar{x}) is definable over CC and therefore over BB where BCB\subseteq C is an appropriate countable subset. Then for all a¯C\bar{a}\in C we have ψ(a¯)=φ(a¯,y¯)tpφ~(b¯/C)=φ(a¯,b¯)\psi(\bar{a})=\varphi(\bar{a},\bar{y})^{\operatorname{tp}_{\tilde{\varphi}}(\bar{b}/C)}=\varphi(\bar{a},\bar{b}). Now let 𝔐\mathfrak{M} be the monster model and a¯X=X(𝔐)\bar{a}\in X=X(\mathfrak{M}). By saturation of \mathcal{M} we can find there some a¯ABb¯a¯\bar{a}^{\prime}\equiv_{AB\bar{b}}\bar{a}. Then a¯C\bar{a}^{\prime}\in C and φ(a¯,b¯)=φ(a¯,b¯)=ψ(a¯)=ψ(a¯)\varphi(\bar{a},\bar{b})=\varphi(\bar{a}^{\prime},\bar{b})=\psi(\bar{a}^{\prime})=\psi(\bar{a}), as desired.

The moreover part follows from the proof. ∎

It follows that if XX is an AA-type-definable set and YXY\subseteq X is a type-definable subset, then YY is type-definable over ABAB for some countable BXB\subseteq X. If YY is a definable set then it is definable over ABAB (by Lemma 1.3, since the definable predicate d(x¯,Y)d(\bar{x},Y) is ABAB-invariant).

Proposition 2.17.

Let φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be an instance of a stable formula, AA a set of parameters. Then the following are equivalent:

  1. (i)

    The condition φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not fork over AA.

  2. (ii)

    There is an AA-definable predicate ψ(x¯)\psi(\bar{x}) which is a faithful combination of AA-conjugates of φ(x¯,b¯)\varphi(\bar{x},\bar{b}) such that ψ(x¯)r\psi(\bar{x})\leq r is consistent.

Proof.

Fix a model A,b¯\mathcal{M}\supseteq A,\bar{b} saturated over AA.

  • (i) \Longrightarrow (ii).

    Let ψ(x¯)\psi(\bar{x}) be as in Lemma 2.14. Let also pSφ(Ab¯)p\in\operatorname{S}_{\varphi}(A\bar{b}) be non forking over AA such that φ(x¯,b¯)pr\varphi(\bar{x},\bar{b})^{p}\leq r. Then ψ(x¯)pφ(x¯,b¯)pr\psi(\bar{x})^{p}\leq\varphi(\bar{x},\bar{b})^{p}\leq r.

  • (iii) \Longrightarrow (i).

    Let ψ(x¯)=θ(φ(x¯,b¯n))n\psi(\bar{x})=\theta\bigl{(}\varphi(\bar{x},\bar{b}_{n})\bigr{)}_{n\in\mathbb{N}} be definable over AA as in the assumption (so b¯nAb¯\bar{b}_{n}\equiv_{A}\bar{b} and θ\theta is a faithful continuous connective).

    By Lemma 2.2 there exists pSφ(M)p\in\operatorname{S}_{\varphi}(M) compatible with ψ(x¯)r\psi(\bar{x})\leq r and non forking over AA, so in particular acleq(A)\operatorname{acl}^{eq}(A)-invariant. Then infnφ(x¯,b¯n)pr\inf_{n}\varphi(\bar{x},\bar{b}_{n})^{p}\leq r by faithfulness, so for all r>rr^{\prime}>r there exists nn such that φ(x¯,b¯n)p<r\varphi(\bar{x},\bar{b}_{n})^{p}<r^{\prime}. Up to an automorphism fixing AA we may assume that φ(x¯,b¯)p<r\varphi(\bar{x},\bar{b})^{p}<r^{\prime}, and by invariance φ(x¯,b¯)p<r\varphi(\bar{x},\bar{b}^{\prime})^{p}<r^{\prime} for every b¯acleq(A)b¯\bar{b}^{\prime}\equiv_{\operatorname{acl}^{eq}(A)}\bar{b}.

    We have thus shown that for every r>rr^{\prime}>r, any set of acleq(A)\operatorname{acl}^{eq}(A)-conjugates of φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r^{\prime} is consistent. By compactness the same holds for φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r. ∎

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 ABA\subseteq B be two sets of parameters. We say that a partial type π\pi over BB is approximately realised in AA if every logical neighbourhood (Definition 1.5) of π\pi over BB is realised in AA.

If \mathcal{M} is a model, BMB\supseteq M, and pSn(B)p\in\operatorname{S}_{n}(B) is approximately realised in MM, we may say that pp is a co-heir of its restriction to \mathcal{M}.

Remark 3.2.
  1. (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.

  2. (ii)

    A complete type over a model \mathcal{M} is always approximately realised there. (This is essentially the Tarski-Vaught Criterion.)

Fact 3.3.

Let ABA\subseteq B and let π(x¯)\pi(\bar{x}) be a partial type over BB.

  1. (i)

    Let XSn(B)X\subseteq\operatorname{S}_{n}(B) consist of all types over BB which are realised in AA, [π]Sn(B)[\pi]\subseteq\operatorname{S}_{n}(B) the closed set defined by π\pi. Then π\pi is approximately realised in AA if and only if [π]X¯[\pi]\cap\overline{X}\neq\varnothing. In particular, X¯\overline{X} is the set of all complete nn-types over BB which are approximately realised in AA.

  2. (ii)

    If CBC\supseteq B then π\pi is approximately realised in AA as a partial type over BB if and only if it is approximately realised in AA as a partial type over CC.

  3. (iii)

    If π\pi is approximately realised in AA then it extends to a complete type πpSn(B)\pi\subseteq p\in\operatorname{S}_{n}(B) which is approximately realised in AA.

  4. (iv)

    A type over a model \mathcal{M} admits extensions to arbitrary sets which are approximately realised in MM.

Proof.

We prove the first two items together. Clearly if π\pi is approximately realised in AA as a partial type over CC then it is approximately realised in AA as a partial type over BB, in which case every neighbourhood of [π][\pi] in Sn(B)\operatorname{S}_{n}(B) intersects XX and by a compactness argument [π][\pi] intersects X¯\overline{X}. Finally, assume [π]X¯[\pi]\cap\overline{X}\neq\varnothing and assume that πφ(x¯)>0\pi\vdash\varphi(\bar{x})>0. Let Y=[φ=0]Sn(C)Y=[\varphi=0]\subseteq\operatorname{S}_{n}(C) and let ZZ be its projection to Sn(B)\operatorname{S}_{n}(B). Then ZZ is compact, Z[π]=Z\cap[\pi]=\varnothing, so U=Sn(B)ZU=\operatorname{S}_{n}(B)\smallsetminus Z is a neighbourhood of [π][\pi]. By assumption there exists a¯A\bar{a}\in A such that tp(a¯/B)XU\operatorname{tp}(\bar{a}/B)\in X\cap U. Then tp(a¯/C)Y\operatorname{tp}(\bar{a}/C)\notin Y, i.e., φ(a¯)>0\varphi(\bar{a})>0, as desired.

For the third item, any p[π]X¯p\in[\pi]\cap\overline{X} will do. For the fourth, use the fact that a type over a model is approximately realised there. ∎

Fact 3.4.

Let 𝒩\mathcal{N} be a model saturated over ANA\subseteq N. If pSn(N)p\in\operatorname{S}_{n}(N) or pSφ(N)p\in\operatorname{S}_{\varphi}(N) is approximately realised in AA then it is AA-invariant.

Proof.

We only consider the case pSφ(N)p\in\operatorname{S}_{\varphi}(N), since the case pSn(N)p\in\operatorname{S}_{n}(N) follows from it. Say b¯,c¯N\bar{b},\bar{c}\in N, b¯Ac¯\bar{b}\equiv_{A}\bar{c}, and let ε>0\varepsilon>0 be given. By assumption there is a¯A\bar{a}\in A such that

|φ(a¯,b¯)φ(x¯,b¯)p|<ε/2,|φ(a¯,c¯)φ(x¯,c¯)p|<ε/2.\displaystyle|\varphi(\bar{a},\bar{b})-\varphi(\bar{x},\bar{b})^{p}|<\varepsilon/2,\qquad|\varphi(\bar{a},\bar{c})-\varphi(\bar{x},\bar{c})^{p}|<\varepsilon/2.

As we assumed that b¯Ac¯\bar{b}\equiv_{A}\bar{c} we have in particular φ(a¯,b¯)=φ(a¯,c¯)\varphi(\bar{a},\bar{b})=\varphi(\bar{a},\bar{c}) and thus |φ(x¯,b¯)pφ(x¯,c¯)p|<ε|\varphi(\bar{x},\bar{b})^{p}-\varphi(\bar{x},\bar{c})^{p}|<\varepsilon, for every ε>0\varepsilon>0. We conclude that φ(x¯,b¯)p=φ(x¯,c¯)p\varphi(\bar{x},\bar{b})^{p}=\varphi(\bar{x},\bar{c})^{p}, as desired. ∎

Lemma 3.5.

Let ABA\subseteq B, p(x¯)Sn(B)p(\bar{x})\in\operatorname{S}_{n}(B) approximately realised in AA, and assume φ(x¯,y¯)\varphi(\bar{x},\bar{y}) is stable. Then pφSφ(B)p{\restriction}_{\varphi}\in\operatorname{S}_{\varphi}(B) does not fork over AA.

Proof.

Let 𝒩B\mathcal{N}\supseteq B be saturated over AA and let qSn(N)q\in\operatorname{S}_{n}(N) extend pp, still approximately realised in AA. Then qq, and thus qφq{\restriction}_{\varphi}, are AA-invariant, so qφq{\restriction}_{\varphi} does not fork over AA and neither does pφp{\restriction}_{\varphi}. ∎

Proposition 3.6.

Let φ(x¯,y¯)\varphi(\bar{x},\bar{y}) be a stable formula, \mathcal{M} a model, AMA\supseteq M. Let also p(x¯)Sφ(A)p(\bar{x})\in\operatorname{S}_{\varphi}(A) be a complete φ\varphi-type, and q(x¯)Sn(M)q(\bar{x})\in\operatorname{S}_{n}(M) a complete type over MM such that pM=qφSφ(M)p{\restriction}_{M}=q{\restriction}_{\varphi}\in\operatorname{S}_{\varphi}(M). Then the following are equivalent:

  1. (i)

    pqp\cup q is approximately realised in MM.

  2. (ii)

    pp is approximately realised in MM.

  3. (iii)

    pp does not fork over MM.

Proof.
  • (i) \Longrightarrow (ii).

    Immediate.

  • (iii) \Longrightarrow (iv).

    Find p(x¯)Sn(A)p^{\prime}(\bar{x})\in\operatorname{S}_{n}(A) extending pp which is approximately realised in MM and use Lemma 3.5.

  • (v) \Longrightarrow (i).

    Find q(x¯)Sn(A)q^{\prime}(\bar{x})\in\operatorname{S}_{n}(A) extending qq which is approximately realised in MM. Then qφq^{\prime}{\restriction}_{\varphi} is non forking over MM by Lemma 3.5, so it must be the unique non forking extension of pM=qφp{\restriction}_{M}=q{\restriction}_{\varphi}. Therefore qpqq\cup p\subseteq q^{\prime} is approximately realised in MM. ∎

Similarly,

Proposition 3.7.

Assume TT is stable. Let \mathcal{M} be a model of TT, AMA\supseteq M, p(x¯)Sn(A)p(\bar{x})\in\operatorname{S}_{n}(A). Then the following are equivalent:

  1. (i)

    pp does not fork over MM.

  2. (ii)

    pp is approximately realised in MM.

If A=𝒩A=\mathcal{N}\succeq\mathcal{M} is saturated over MM then these are further equivalent to

  1. (iii)

    pp is MM-invariant.

Definition 3.8.

Let \mathcal{M} be a model, B\mathcal{M}\subseteq B. A type pSn(B)p\in\operatorname{S}_{n}(B) is said to be an heir of its restriction to MM if for every formula φ(x¯,b¯,m¯)\varphi(\bar{x},\bar{b},\bar{m}) with b¯B\bar{b}\in B and m¯M\bar{m}\in M, and for every ε>0\varepsilon>0, there are b¯M\bar{b}^{\prime}\in M such that |φ(x¯,b¯,m¯)φ(x¯,b¯,m¯)|p<ε|\varphi(\bar{x},\bar{b},\bar{m})-\varphi(\bar{x},\bar{b}^{\prime},\bar{m})|^{p}<\varepsilon.

Clearly every type over a model is an heir of itself. Also, it is not difficult to check that if \mathcal{M} is a model and a¯\bar{a}, b¯\bar{b} are two tuples possibly outside \mathcal{M} then

tp(a¯/Mb¯) is an heir of tp(a¯/M)tp(b¯/Ma¯) is a co-heir of tp(b¯/M).\displaystyle\operatorname{tp}(\bar{a}/M\bar{b})\text{ is an heir of }\operatorname{tp}(\bar{a}/M)\quad\Longleftrightarrow\quad\operatorname{tp}(\bar{b}/M\bar{a})\text{ is a co-heir of }\operatorname{tp}(\bar{b}/M).

Finally, a standard compactness argument yields that if BC\mathcal{M}\subseteq B\subseteq C and pSn(B)p\in\operatorname{S}_{n}(B) is an heir of pMp{\restriction}_{M} then it admits an extension to CC which is an heir as well.

Lemma 3.9.

Let \mathcal{M} be a model, p(x¯)Sn(M)p(\bar{x})\in\operatorname{S}_{n}(M). Then pp is definable if and only if it has a unique heir to every superset BB\supseteq\mathcal{M}.

Proof.

(We follow Poizat [Poi85, Théorème 11.07].) For left to right, assume pp is definable and let qSn(B)q\in\operatorname{S}_{n}(B) be an heir of pp, where BMB\supseteq M. Let φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be a formula over BB and let dpφ(y¯,c)d_{p}\varphi(\bar{y},c) be the φ\varphi-definition of pp, cMc\in M. Assume that dpφ(b¯,c)φ(x¯,b¯)qd_{p}\varphi(\bar{b},c)\neq\varphi(\bar{x},\bar{b})^{q}, i.e., |dpφ(b¯,c)φ(x¯,b¯)|q>0|d_{p}\varphi(\bar{b},c)-\varphi(\bar{x},\bar{b})|^{q}>0. Then there is b¯M\bar{b}^{\prime}\in M such that |dpφ(b¯,c)φ(x¯,b¯)|q>0|d_{p}\varphi(\bar{b}^{\prime},c)-\varphi(\bar{x},\bar{b}^{\prime})|^{q}>0, a contradiction.

Conversely, assume pp admits a unique heir to every structure. let \mathcal{L}^{\prime} be \mathcal{L} along with a new predicate symbol Dφ(y¯)D_{\varphi}(\bar{y}) for each formula φ(x¯,y¯)\varphi(\bar{x},\bar{y}) (here x¯\bar{x} is fixed, y¯\bar{y} may vary with φ\varphi). We define an expansion \mathcal{M}^{\prime} of \mathcal{M} by interpreting Dφ(b¯)=φ(x¯,b¯)pD_{\varphi}(\bar{b})=\varphi(\bar{x},\bar{b})^{p}. Assume now that 𝒩\mathcal{N}^{\prime}\succeq_{\mathcal{L}^{\prime}}\mathcal{M}^{\prime} and let

𝒩=𝒩,q={φ(x¯,b¯)=Dφ(b¯)}b¯N.\displaystyle\mathcal{N}=\mathcal{N}^{\prime}{\restriction}_{\mathcal{L}},\qquad q=\bigl{\{}\varphi(\bar{x},\bar{b})=D_{\varphi}(\bar{b})\bigr{\}}_{\bar{b}\in N}.

It is not difficult to see that q(x¯)Sn(N)q(\bar{x})\in\operatorname{S}_{n}(N) is a complete, consistent type, and that it is moreover an heir of pp. Since, by assumption, the heir is unique, 𝒩\mathcal{N}^{\prime} is the unique expansion of 𝒩\mathcal{N} which is an elementary extension of \mathcal{M}^{\prime}. In other words, a model of Diag()\operatorname{Diag}(\mathcal{M}) admits at most one expansion to a model of Diag()\operatorname{Diag}(\mathcal{M}^{\prime}). By Beth’s Theorem (see [Ben09]) for each formula φ(x¯,y¯)\varphi(\bar{x},\bar{y}) there exists an \mathcal{M}-definable predicate dpφ(y¯)d_{p}\varphi(\bar{y}) such that Diag()Dφ=dpφ\operatorname{Diag}(\mathcal{M}^{\prime})\vdash D_{\varphi}=d_{p}\varphi. In particular, φ(x¯,m¯)p=Dφ(m¯)=dpφ(m¯)\varphi(\bar{x},\bar{m})^{p}=D_{\varphi}(\bar{m})=d_{p}\varphi(\bar{m}) for every m¯M\bar{m}\in M, and pp is definable. ∎

Notice that for a pair of models 𝒩\mathcal{M}\subseteq\mathcal{N} we could have defined a notion of a φ\varphi-type over a 𝒩\mathcal{N} being an heir of its restriction to \mathcal{M}, in which case Lemma 3.9 holds, with the same proof, for local types.

Theorem 3.10.

The following are equivalent for a theory TT:

  1. (i)

    The theory TT is stable.

  2. (ii)

    Every type over a model has a unique co-heir to any superset.

  3. (iii)

    Every type over a model has a unique heir to any superset.

Proof.
  • (i) \Longrightarrow (ii).

    Assume TT is stable, B\mathcal{M}\subseteq B, and qSn(B)q\in\operatorname{S}_{n}(B) is a co-heir of p=qMp=q{\restriction}_{M}. Let 𝒩B\mathcal{N}\supseteq B be saturated over \mathcal{M} and let qSn(N)q^{\prime}\in\operatorname{S}_{n}(N) extend qq, also a co-heir of pp. Then qq^{\prime} is MM-invariant and therefore the unique non forking extension of pp to NN. Thus qq is the unique non forking extension of pp to BB.

  • (iii) \Longrightarrow (iv).

    Let \mathcal{M} be a model, pSn(M)p\in\operatorname{S}_{n}(M). In order to show that pp has a unique heir to every BMB\supseteq M it is enough to consider the case B=Mb¯B=M\bar{b} where b¯\bar{b} is a finite tuple. So indeed, assume that a¯\bar{a} realises an heir of pp to Mb¯M\bar{b}. Then tp(b¯/Ma¯)\operatorname{tp}(\bar{b}/M\bar{a}) is a co-heir of tp(b¯/M)\operatorname{tp}(\bar{b}/M) and by assumption it is uniquely determined by tp(b¯/M)\operatorname{tp}(\bar{b}/M) and by a¯\bar{a}. It follows that tp(a¯/Mb¯)\operatorname{tp}(\bar{a}/M\bar{b}) is uniquely determined by b¯\bar{b} and tp(a¯/M)\operatorname{tp}(\bar{a}/M), as desired.

  • (v) \Longrightarrow (i).

    The assumption and Lemma 3.9 yield that every type is definable, so TT is stable. ∎

Using the local version of Lemma 3.9 alluded to above we can prove a local version of Theorem 3.10, namely that φ(x¯,y¯)\varphi(\bar{x},\bar{y}) is stable if and only if every φ\varphi-type over a model admits a unique co-heir to larger sets if and only if every φ\varphi-type over models admits a unique heir to larger models.

4. Invariant types, indiscernible sequences and dividing

Fact 4.1.

Let \mathcal{M} be a model saturated over AMA\subseteq M, and let pSn(M)p\in\operatorname{S}_{n}(M) be AA-invariant. Let (a¯n)nM(\bar{a}_{n})_{n\in\mathbb{N}}\subseteq M be a sequence constructed inductively, choosing each a¯n\bar{a}_{n} to realise pAa¯<np{\restriction}_{A\bar{a}_{<n}}.

Then the sequence (a¯n)n(\bar{a}_{n})_{n\in\mathbb{N}} is AA-indiscernible, and its type over AA depends only on pp.

Proof.

Standard. ∎

The common type over AA of such sequences will be denoted by p(ω)Ap^{(\omega)}{\restriction}_{A}. For every finite or countable BMB\subseteq M we may construct p(ω)ABp^{(\omega)}{\restriction}_{A\cup B} just as well. By a gluing argument, p(ω)={p(ω)AB:B[M]0}p^{(\omega)}=\bigcup\bigl{\{}p^{(\omega)}{\restriction}_{A\cup B}\colon B\in[M]^{\aleph_{0}}\bigr{\}} is a complete type of an MM-indiscernible sequence in pp, and is of course AA-invariant.

Lemma 4.2.

Let AA be a set, φ(x¯,y¯)\varphi(\bar{x},\bar{y}) a stable formula, pSφ(A)p\in\operatorname{S}_{\varphi}(A) a stationary φ\varphi-type. Let A\mathcal{M}\supseteq A be saturated over AA, and let pqSn(M)p\subseteq q\in\operatorname{S}_{n}(M), qq invariant over AA. Let (c¯n)nq(ω)A(\bar{c}_{n})_{n\in\mathbb{N}}\vDash q^{(\omega)}{\restriction}_{A} be an AA-indiscernible sequence as constructed in Fact 4.1.

Then the sequence (φ~n(y¯,c¯2n))n\bigl{(}\tilde{\varphi}^{n}(\bar{y},\bar{c}_{\leq 2n})\bigr{)}_{n\in\mathbb{N}} converges uniformly to the definition dpφ(y¯)d_{p}\varphi(\bar{y}) at a rate which only depends on φ\varphi.

Proof.

Since qφq{\restriction}_{\varphi} is AA-invariant, it does not fork over AA, so dpφ(y¯)=dqφ(y¯)d_{p}\varphi(\bar{y})=d_{q}\varphi(\bar{y}).

Fix ε>0\varepsilon>0. By Fact 2.1 there is k=k(φ,ε)k=k(\varphi,\varepsilon) and a sequence (c¯n)n2kM(\bar{c}^{\prime}_{n})_{n\leq 2k}\subseteq M such that |dpφ(y¯)φ~k(y¯,c¯2k)|ε|d_{p}\varphi(\bar{y})-\tilde{\varphi}^{k}(\bar{y},\bar{c}^{\prime}_{\leq 2k})|\leq\varepsilon, and such that furthermore c¯nqA,c¯<n\bar{c}^{\prime}_{n}\vDash q{\restriction}_{A,\bar{c}^{\prime}_{<n}}. By Fact 4.1 we have c¯2kAc¯2k\bar{c}_{\leq 2k}\equiv_{A}\bar{c}^{\prime}_{\leq 2k}. In addition, dpφd_{p}\varphi is over AA, so |dpφ(y¯)φ~k(y¯,c¯2k)|ε|d_{p}\varphi(\bar{y})-\tilde{\varphi}^{k}(\bar{y},\bar{c}_{\leq 2k})|\leq\varepsilon.

Consider now n>kn>k. First of all, by exactly the same argument as above, for every w[2n+1]2k+1w\in[2n+1]^{2k+1} we have |dpφ(y¯)φ~k(y¯,c¯w)|ε|d_{p}\varphi(\bar{y})-\tilde{\varphi}^{k}(\bar{y},\bar{c}_{\in w})|\leq\varepsilon. In addition, for any b¯\bar{b} there exists a subset w[2n+1]2k+1w\in[2n+1]^{2k+1} such that φ~n(b¯,c¯2n)=φ~k(b¯,c¯w)\tilde{\varphi}^{n}(\bar{b},\bar{c}_{\leq 2n})=\tilde{\varphi}^{k}(\bar{b},\bar{c}_{\in w}) (from any set of 2n+12n+1 reals one can choose a subset of size 2k+12k+1 with the same median value). Thus |dpφ(y¯)φ~n(y¯,c¯2n)|ε|d_{p}\varphi(\bar{y})-\tilde{\varphi}^{n}(\bar{y},\bar{c}_{\leq 2n})|\leq\varepsilon for all nkn\geq k, where kk depends only on ε\varepsilon and φ\varphi, as desired. ∎

Proposition 4.3.

Let φ(x¯,b¯)\varphi(\bar{x},\bar{b}) be an instance of a stable formula, AA a set of parameters. Then the following are equivalent:

  1. (i)

    The condition φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not fork over AA.

  2. (ii)

    If (b¯n)n(\bar{b}_{n})_{n\in\mathbb{N}} is an AA-indiscernible sequence, b¯0=b¯\bar{b}_{0}=\bar{b}, then the set of conditions {φ(x¯,b¯n)r}n\{\varphi(\bar{x},\bar{b}_{n})\leq r\}_{n\in\mathbb{N}} is consistent (i.e., the condition φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not divide over AA).

Proof.
  • (i) \Longrightarrow (ii).

    If (b¯n)n(\bar{b}_{n})_{n\in\mathbb{N}} is an AA-indiscernible sequence and b¯0=b¯\bar{b}_{0}=\bar{b} then each b¯n\bar{b}_{n} is an acleq(A)\operatorname{acl}^{eq}(A)-conjugate of b¯\bar{b}.

  • (iii) \Longrightarrow (i).

    Fix models 𝒩A\mathcal{N}\succeq\mathcal{M}\supseteq A where 𝒩\mathcal{N} is saturated over MM. Let q0=tp(b¯/acleq(A))q_{0}=\operatorname{tp}(\bar{b}/\operatorname{acl}^{eq}(A)). By Lemma 2.2 there exists qSm(M)q\in\operatorname{S}_{m}(M) extending q0q_{0} such that qφ~q{\restriction}_{\tilde{\varphi}} does not fork over AA, i.e., such that dqφ~=dq0φ~d_{q}\tilde{\varphi}=d_{q_{0}}\tilde{\varphi}. Let q1Sm(N)q_{1}\in\operatorname{S}_{m}(N) be an MM-invariant extension of qq. Finally, let (b¯n)nq1(ω)M(\bar{b}_{n})_{n\in\mathbb{N}}\vDash q_{1}^{(\omega)}{\restriction}_{M}. Then (b¯n)n(\bar{b}_{n})_{n\in\mathbb{N}} is an \mathcal{M}-indiscernible sequence, and a fortiori AA-indiscernible, in tp(b¯/A)\operatorname{tp}(\bar{b}/A). Thus by assumption there exists a¯\bar{a} such that φ(a¯,b¯n)r\varphi(\bar{a},\bar{b}_{n})\leq r for all nn. In addition, by Lemma 4.2 we have

    dqφ~(a¯)=limnmedn(φ(a¯,b¯i))i2nr.\displaystyle d_{q}\tilde{\varphi}(\bar{a})=\lim_{n}\operatorname{med}_{n}\bigl{(}\varphi(\bar{a},\bar{b}_{i})\bigr{)}_{i\leq 2n}\leq r.

    Let pSφ(M)p\in\operatorname{S}_{\varphi}(M) be a non forking extension of tpφ(a¯/acleq(A))\operatorname{tp}_{\varphi}(\bar{a}/\operatorname{acl}^{eq}(A)). Then φ(x¯,b¯)p=dqφ~(x¯)pr\varphi(\bar{x},\bar{b})^{p}=d_{q}\tilde{\varphi}(\bar{x})^{p}\leq r, witnessing that φ(x¯,b¯)r\varphi(\bar{x},\bar{b})\leq r does not fork over AA, as desired. ∎

5. Canonical bases

Recall that the canonical base of a stationary type pSn(A)p\in\operatorname{S}_{n}(A) in a stable theory is Cb(p)={Cb(pφ):φ(x¯,)}\mathrm{Cb}(p)=\{\mathrm{Cb}(p{\restriction}_{\varphi})\colon\varphi(\bar{x},\ldots)\in\mathcal{L}\}, namely the set of all canonical parameters of φ\varphi-definitions of pp.

Proposition 5.1.

Assume TT is stable, and let p(x¯)Sn(A)p(\bar{x})\in\operatorname{S}_{n}(A) be stationary. Then:

  1. (i)

    Cb(p)dcleq(A)\mathrm{Cb}(p)\subseteq\operatorname{dcl}^{eq}(A).

  2. (ii)

    pp does not fork over Cb(p)\mathrm{Cb}(p).

  3. (iii)

    pCb(p)p{\restriction}_{\mathrm{Cb}(p)} is stationary.

  4. (iv)

    Cb(p)\mathrm{Cb}(p) is minimal for the three previous properties, meaning that if Bdcleq(A)B\subseteq\operatorname{dcl}^{eq}(A) and pBp{\restriction}_{B} is a stationary non forking restriction then Cb(p)dcleq(B)\mathrm{Cb}(p)\subseteq\operatorname{dcl}^{eq}(B).

Proof.

The first two items are immediate, while the third is by Corollary 2.8. Under the assumptions of the fourth we have Cb(p)=Cb(pB)dcleq(B)\mathrm{Cb}(p)=\mathrm{Cb}(p{\restriction}_{B})\subseteq\operatorname{dcl}^{eq}(B). ∎

The four properties listed in Proposition 5.1 determine the canonical base up to inter-definability. Indeed, if BB has all four then Cb(p)dcleq(B)\mathrm{Cb}(p)\subseteq\operatorname{dcl}^{eq}(B) but also Bdcleq(Cb(p))B\subseteq\operatorname{dcl}^{eq}(\mathrm{Cb}(p)), whereby dcleq(B)=dcleq(Cb(p))\operatorname{dcl}^{eq}(B)=\operatorname{dcl}^{eq}(\mathrm{Cb}(p)). In this case we say that BB is a canonical base for pp.

Proposition 5.2.

Assume TT is stable, and let p(x¯)Sn(A)p(\bar{x})\in\operatorname{S}_{n}(A) be stationary. Let qSn(M)q\in\operatorname{S}_{n}(M) be the unique non forking extension of pp, where \mathcal{M} is saturated over AA. Then a (small) set BMB\subseteq M is a canonical base for pp if and only if, for every fAut()f\in\operatorname{Aut}(\mathcal{M}): fB=idBf(q)=qf{\restriction}_{B}=\operatorname{id}_{B}\Longleftrightarrow f(q)=q.

Proof.

Let C=Cb(p)=Cb(q)C=\mathrm{Cb}(p)=\mathrm{Cb}(q). It follows directly from the definitions that an automorphism of \mathcal{M} fixes qq if and only if it fixes qφq{\restriction}_{\varphi} for every formula φ(x¯,)\varphi(\bar{x},\ldots), if and only if it fixes every member of CC. A small set BB is another canonical base for pp if and only if dcleq(B)=dcleq(C)\operatorname{dcl}^{eq}(B)=\operatorname{dcl}^{eq}(C) which is further equivalent to BB and CC 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 p(x¯)Sm(A)p(\bar{x})\in\operatorname{S}_{m}(A) is a sequence I=(a¯n)nI=(\bar{a}_{n})_{n\in\mathbb{N}} of realisations of pp which is independent over AA, i.e., such that a¯nAa¯<n\bar{a}_{n}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}\bar{a}_{<n} for all nn\in\mathbb{N}. It follows by standard independence calculus that a¯sAa¯t\bar{a}_{\in s}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}\bar{a}_{\in t} for every two disjoint index sets s,ts,t\subseteq\mathbb{N}. From stationarity of pp it follows that the sequence II is indiscernible over AA, and its type over AA, which we may denote by p(ω)p^{(\omega)}, is uniquely determined by pp.

It is not difficult to check that if pp satisfies the assumptions of Fact 4.1 then the definition of p(ω)p^{(\omega)} which appears thereafter agrees with the one given here. In the general case, let \mathcal{M} be saturated over AA and let qSm(M)q\in\operatorname{S}_{m}(M) be the non forking extension of pp. Then by construction, p(ω)=qA(ω)p^{(\omega)}=q^{(\omega)}_{A}, 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 I=(a¯n)nI=(\bar{a}_{n})_{n\in\mathbb{N}} be a sequence of tuples, or, for that matter, even of sets. Let IkI^{\geq k} denote the tail (a¯n)nk(\bar{a}_{n})_{n\geq k}. We define the tail definable closure of II as

tdcleq(I)=kdcleq(Ik).\displaystyle\operatorname{tdcl}^{eq}(I)=\bigcap_{k\in\mathbb{N}}\operatorname{dcl}^{eq}(I^{\geq k}).

It is not difficult to see that for an indiscernible sequence II, tdcleq(I)\operatorname{tdcl}^{eq}(I) consists precisely of all cdcleq(I)c\in\operatorname{dcl}^{eq}(I) over which II is indiscernible.

Lemma 5.4.

Let I=(a¯n)nI=(\bar{a}_{n})_{n\in\mathbb{N}} and J=(b¯n)nJ=(\bar{b}_{n})_{n\in\mathbb{N}} be indiscernible sequences such that the concatenation IJI{{}^{\frown}}J is indiscernible as well. Then tdcleq(I)=tdcleq(J)\operatorname{tdcl}^{eq}(I)=\operatorname{tdcl}^{eq}(J). Moreover, every automorphism which sends II to JJ necessarily fixes tdcleq(I)\operatorname{tdcl}^{eq}(I).

Proof.

For kk\in\mathbb{N} let JkJ_{k} be the sequence a¯0,,a¯k1,b¯k,b¯k+1,\bar{a}_{0},\ldots,\bar{a}_{k-1},\bar{b}_{k},\bar{b}_{k+1},\ldots, namely the sequence obtained by replacing the first kk elements of JJ with the corresponding elements from II. Since IJI{{}^{\frown}}J is indiscernible so is JkJ_{k} for each kk, and there exists an automorphism fkf_{k} sending JJkJ\mapsto J_{k}. Now let ctdcleq(J)c\in\operatorname{tdcl}^{eq}(J). Since cc is definable over JkJ^{\geq k} it is fixed by fkf_{k}, so cJcJkcJ\equiv cJ_{k}. This holds for all kk, whence cIcJcI\equiv cJ.

Fix an automorphism ff which sends II to JJ (which must necessarily exist). Then f(c)JcIcJf(c)J\equiv cI\equiv cJ, so f(c)=cf(c)=c. Thus ff fixes tdcleq(J)\operatorname{tdcl}^{eq}(J). Applying f1f^{-1} we obtain that tdcleq(I)=tdcleq(J)\operatorname{tdcl}^{eq}(I)=\operatorname{tdcl}^{eq}(J), as desired. ∎

Theorem 5.5.

Let pSm(A)p\in\operatorname{S}_{m}(A) be a stationary type and let I=(a¯n)nI=(\bar{a}_{n})_{n\in\mathbb{N}} be a Morley sequence in pp. Then tdcleq(I)\operatorname{tdcl}^{eq}(I) is a canonical base of pp.

Proof.

First of all, we have seen that pCb(p)p{\restriction}_{\mathrm{Cb}(p)} is stationary, with the same canonical base as pp. It is also not difficult to check that a Morley sequence in pp is also a Morley sequence in pCb(p)p{\restriction}_{\mathrm{Cb}(p)}. It is therefore enough to prove for pCb(p)p{\restriction}_{\mathrm{Cb}(p)}, i.e., we may assume that A=Cb(p)A=\mathrm{Cb}(p).

So let \mathcal{M} be saturated over A=Cb(p)A=\mathrm{Cb}(p) and let qSm(M)q\in\operatorname{S}_{m}(M) be the non forking extension of pp. As pointed above, Ip(ω)=qA(ω)I\vDash p^{(\omega)}=q^{(\omega)}_{A}. By Lemma 4.2 pp is definable over II, so Cb(p)dcleq(I)\mathrm{Cb}(p)\subseteq\operatorname{dcl}^{eq}(I). Also, every tail of a Morley sequence is a Morley sequence, whence Cb(p)tdcleq(I)\mathrm{Cb}(p)\subseteq\operatorname{tdcl}^{eq}(I).

Conversely, let ff be an automorphism fixing A=Cb(p)A=\mathrm{Cb}(p). Then ff fixes pp and therefore sends II to another Morley sequence in pp, say JJ. Let KK be a third Morley sequence in pp, KAI,JK\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}I,J. Then both IKI{{}^{\frown}}K and JKJ{{}^{\frown}}K can be verifies to be Morley sequences in pp (of length ω+ω\omega+\omega), and in particular indiscernible. We can decompose f=hgf=h\circ g where g(I)=Kg(I)=K and h(K)=Jh(K)=J. By the Lemma tdcleq(I)=tdcleq(K)=tdcleq(J)\operatorname{tdcl}^{eq}(I)=\operatorname{tdcl}^{eq}(K)=\operatorname{tdcl}^{eq}(J) and this set is fixed by gg, hh and therefore by ff. Thus tdcleq(I)dcl(Cb(p))\operatorname{tdcl}^{eq}(I)\subseteq\operatorname{dcl}\bigl{(}\mathrm{Cb}(p)\bigr{)}, 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 I=(a¯n)nI=(\bar{a}_{n})_{n\in\mathbb{N}} is a Morley sequence in some type, say qq. Let A=tdcleq(I)A=\operatorname{tdcl}^{eq}(I) and p=tp(a¯n/A)p=\operatorname{tp}(\bar{a}_{n}/A), which does not depend on nn. By the Theorem, A=Cb(q)=Cb(p)A=\mathrm{Cb}(q)=\mathrm{Cb}(p) and II is a Morley sequence in pp.

In the case of probability theory this is a well known fact. Indeed, in probability algebras or in spaces of random variables (say [0,1][0,1]-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 (Xn)n(X_{n})_{n\in\mathbb{N}} is sequence of random variables which is indiscernible (i.e., exchangeable) and 𝒜\mathscr{A} is its tail algebra then the sequence (Xn)n(X_{n})_{n\in\mathbb{N}} is i.i.d. over 𝒜\mathscr{A}, meaning that the random variables XnX_{n} are independent over 𝒜\mathscr{A} and have the same conditional distribution over 𝒜\mathscr{A}.

Corollary 5.6.

Assume TT is stable, and let p(x¯)Sm(A)p(\bar{x})\in\operatorname{S}_{m}(A) be stationary. Let I=(a¯n)nI=(\bar{a}_{n})_{n\in\mathbb{N}} be a Morley sequence in pp, J=Ia¯0J=I\smallsetminus\bar{a}_{0}. Then a¯0AJ\bar{a}_{0}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}J and a¯0JA\bar{a}_{0}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{J}A.

Proof.

The first independence is immediate and implies a¯0Cb(p)AJ\bar{a}_{0}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{\mathrm{Cb}(p)}AJ. By Theorem 5.5 we have Cb(p)dcleq(J)\mathrm{Cb}(p)\subseteq\operatorname{dcl}^{eq}(J) 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 G,S\langle G,S\rangle be a homogeneous space, type-definable in models of a stable theory TT. This is to say that GG is a type-definable group and SS a type-definable set, equipped with a type-definable (and therefore definable) transitive group action G×SSG\times S\to S. For convenience let us assume that both are defined without parameters. We shall identify GG and SS with their sets of realisations in a monster model 𝔐\mathfrak{M}. We are particularly interested in the case where S=GS=G where GG acts on itself either on the left (g,h)gh(g,h)\mapsto gh or on the right (g,h)hg1(g,h)\mapsto hg^{-1}.

Given a partial type π(x)\pi(x) in the sort of SS we let π(S)\pi(S) denote the subset of SS defined by π\pi.

Definition 6.1.
  1. (i)

    A generic set in SS is a subset XSX\subseteq S finitely many GG-translates of which cover SS.

  2. (ii)

    A generic partial type in SS is a partial type π(x)\pi(x) such that every logical neighbourhood of π\pi (as per Definition 1.5) defines in SS a generic set. Single conditions as well as complete types are generic if they are generic as partial types.

  3. (iii)

    We say that an element sSs\in S is generic over a set AA if tp(s/A)\operatorname{tp}(s/A) is generic.

  4. (iv)

    A left-generic set in GG is a subset XGX\subseteq G which is generic under the action of GG on itself on the left. We define partial types in the sort of GG to be left-generic accordingly. Similarly for right-generic.

Let π(x)\pi(x) be a partial type. Clearly, if π(S)\pi(S) is a generic set then π\pi is a generic partial type, but the converse is not always true. In classical logic, if π\pi consists of a single formula (i.e., if π(S)\pi(S) is a relatively definable subset of SS, and so is its complement), then π\pi is its own logical neighbourhood and the two notions coincide. Unfortunately, this will generally never happen in continuous logic (except for π(S)=S\pi(S)=S or π(S)=\pi(S)=\varnothing).

Lemma 6.2.

The following are equivalent for a partial type π(x)\pi(x) in the sort of SS, with parameters in a set AA:

  1. (i)

    The partial type π\pi is generic in SS.

  2. (ii)

    For every formula φ(x,a¯)\varphi(x,\bar{a}) over AA, if the condition φ(x,a¯)=0\varphi(x,\bar{a})=0 is a logical neighbourhood of π\pi then it is a generic condition.

Proof.

One direction is immediate, the other follow from Lemma 1.6. ∎

Let SS(A)\operatorname{S}_{S}(A) denote the set of all complete types over AA implying xSx\in S. Equipped with the induced topology from Sx(A)\operatorname{S}_{x}(A), it is a compact space, and the set of all generic complete types over AA is closed. Closed subsets of SS(A)\operatorname{S}_{S}(A) are in bijection with partial types over AA implying xSx\in S, i.e., with type-definable subsets of SS using parameters in AA. If X,YSX,Y\subseteq S are two such sets, say that YY is a logical neighbourhood of XX relative to SS, in symbols Y>SXY>^{S}X, if [X][Y][X]\subseteq[Y]^{\circ} where the interior is calculated in SS(A)\operatorname{S}_{S}(A). This is equivalent to saying that there exists a true logical neighbourhood Y>XY^{\prime}>X such that Y=YSY=Y^{\prime}\cap S. Thus a type-definable set XSX\subseteq S is defined by a generic partial type in SS if and only if every relative logical neighbourhood of XX in SS defines a generic set.

For gGg\in G and XSX\subseteq S, let Lg[X]=gX={gs}sXL_{g}[X]=gX=\{gs\}_{s\in X}. Somewhat superfluously, we may also define Lg1[X]={sS:gsX}=Lg1[X]L_{g}^{-1}[X]=\{s\in S\colon gs\in X\}=L_{g^{-1}}[X].

Lemma 6.3.

Let AA be a set of parameters, gG(A)=Gdcl(A)g\in G(A)=G\cap\operatorname{dcl}(A).

  1. (i)

    If XSX\subseteq S is type-definable over AA, say by a partial type π\pi, then Lg[X]L_{g}[X] is also type-definable over AA by a partial type which will be denoted LgπL_{g}\pi (or gπg\pi). Moreover, π\pi is generic if and only if gπg\pi is.

  2. (ii)

    If p=tp(s/A)SS(A)p=\operatorname{tp}(s/A)\in\operatorname{S}_{S}(A) is a complete type then Lgp=gp=tp(gp/A)L_{g}p=gp=\operatorname{tp}(gp/A), and Lg:SS(A)SS(A)L_{g}\colon\operatorname{S}_{S}(A)\to\operatorname{S}_{S}(A) 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 π\pi is generic, and let gX<SYgX<^{S}Y. Then X<SLg1[Y]X<^{S}L_{g}^{-1}[Y], so Lg1[Y]L_{g}^{-1}[Y] is a generic subset of SS. It follows immediately that so is YY. Thus gπg\pi is a generic partial type. For the converse replace gg with g1g^{-1}. ∎

Similarly, for sSs\in S and XGX\subseteq G we define Rs[X]=Xs={gs}gXR_{s}[X]=Xs=\{gs\}_{g\in X}. For XSX\subseteq S we define Rs1[X]={gG:gsX}R_{s}^{-1}[X]=\{g\in G\colon gs\in X\}.

Lemma 6.4.

Let AA be a set of parameters, sS(A)=Sdcl(A)s\in S(A)=S\cap\operatorname{dcl}(A).

  1. (i)

    If XGX\subseteq G is type-definable over AA, say by a partial type π\pi, then Rs[X]R_{s}[X] is also type-definable over AA by a partial type which will be denoted RsπR_{s}\pi (or πs\pi s). Moreover, if π\pi is left-generic then RsπR_{s}\pi is generic.

  2. (ii)

    If p=tp(g/A)SG(A)p=\operatorname{tp}(g/A)\in\operatorname{S}_{G}(A) is a complete type then Rsp=ps=tp(gs/A)R_{s}p=ps=\operatorname{tp}(gs/A), and Rs:SG(A)SS(A)R_{s}\colon\operatorname{S}_{G}(A)\to\operatorname{S}_{S}(A) is a continuous surjection, sending left-generic types to generic types.

Notice that we do not claim that every generic type in SS(A)\operatorname{S}_{S}(A) is the image under RsR_{s} of a left-generic type in SG(A)\operatorname{S}_{G}(A) (this is true if TT is stable).

Proof.

Essentially identical to that of Lemma 6.3. ∎

Under the assumption that the theory T=Th(𝔐)T=\operatorname{Th}(\mathfrak{M}) 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 𝔐^=G,S,\hat{\mathfrak{M}}=\langle G,S,\ldots\rangle in a language ^\hat{\mathcal{L}} (in addition to sorts GG and SS, ^\hat{\mathcal{L}} consists of additional sorts which we shall described later). We define the distance on the first two sorts by

dG𝔐^(g,g)=suphGd𝔐(hg,hg),dS𝔐^(s,s)=suphGd𝔐(hs,hs).\displaystyle d_{G}^{\hat{\mathfrak{M}}}(g,g^{\prime})=\sup_{h\in G}\,d^{\mathfrak{M}}(hg,hg^{\prime}),\qquad d_{S}^{\hat{\mathfrak{M}}}(s,s^{\prime})=\sup_{h\in G}\,d^{\mathfrak{M}}(hs,hs^{\prime}).

This coincides with the original distance in 𝔐\mathfrak{M} if the latter is invariant under the action of GG (on the left). In any case, dG𝔐^d_{G}^{\hat{\mathfrak{M}}} is a distance function, invariant under the action of GG, and satisfies dG𝔐^d𝔐d_{G}^{\hat{\mathfrak{M}}}\geq d^{\mathfrak{M}}. On the other hand, if gngg_{n}\to g in d𝔐d^{\mathfrak{M}} then gngg_{n}\to g in dG𝔐^d_{G}^{\hat{\mathfrak{M}}} as well (if not, then by a compactness argument, for some ε>0\varepsilon>0 there would exist hGh\in G such that d𝔐(hg,hg)εd^{\mathfrak{M}}(hg,hg)\geq\varepsilon, an absurd). It follows that (G,dG𝔐^)(G,d_{G}^{\hat{\mathfrak{M}}}) is a complete metric space. The same observations hold for (S,dS𝔐^)(S,d_{S}^{\hat{\mathfrak{M}}}).

Let now ΦS\Phi_{S} consist of all \mathcal{L}-formulae of the form φ(x,y¯)\varphi(x,\bar{y}) where xx is in the sort of SS. For each φΦS\varphi\in\Phi_{S}, there will be a sort CφC_{\varphi}, consisting of all canonical parameters of instances of φ\varphi in 𝔐\mathfrak{M}. The canonical parameter of φ(x,b¯)\varphi(x,\bar{b}) will be denote [b¯]φ[\bar{b}]_{\varphi}, or [b¯][\bar{b}] if there is no ambiguity. We put on it the standard metric, namely

dφ([b¯]φ,[b¯]φ)=supa𝔐|φ(a,b¯)φ(a,b¯)|.\displaystyle d_{\varphi}([\bar{b}]_{\varphi},[\bar{b}^{\prime}]_{\varphi})=\sup_{a\in\mathfrak{M}}\,|\varphi(a,\bar{b})-\varphi(a,\bar{b}^{\prime})|.

The only symbols in the language ^\hat{\mathcal{L}}, in addition to the distance symbols of the various sorts, are a predicate symbol φ^(xS,yG,zφ)\hat{\varphi}(x_{S},y_{G},z_{\varphi}) for each formula φΦS\varphi\in\Phi_{S}, interpreted by

φ^(s,g,[b¯])𝔐^=φ(g1s,b¯)𝔐.\displaystyle\hat{\varphi}(s,g,[\bar{b}])^{\hat{\mathfrak{M}}}=\varphi(g^{-1}s,\bar{b})^{\mathfrak{M}}.

Since φ\varphi is uniformly continuous in all its variables, so is φ^\hat{\varphi}. These definitions make 𝔐^\hat{\mathfrak{M}} a continuous ^\hat{\mathcal{L}}-structure.

If G,S\langle G,S\rangle is definable then 𝔐^\hat{\mathfrak{M}} is interpretable in 𝔐\mathfrak{M} and T^=Th^(𝔐^)\hat{T}=\operatorname{Th}_{\hat{\mathcal{L}}}(\hat{\mathfrak{M}}) is stable (assuming TT is). In the general case, all we know is that 𝔐^\hat{\mathfrak{M}} is saturated for quantifier-free types in which only φ^\hat{\varphi} appear. It follows from stability in TT that each formula φ^(x,y,z)\hat{\varphi}(x,y,z), with any partition of the variables, is stable.

For hGh\in G define a mapping θh:𝔐^𝔐^\theta_{h}\colon\hat{\mathfrak{M}}\to\hat{\mathfrak{M}} by sending gGg\in G to hghg, sSs\in S to hshs, and fixing all the auxiliary sorts. This is easily verified to be an automorphism of 𝔐^\hat{\mathfrak{M}}. Since the action of GG on SS is assumed to be transitive, if AφCφA\subseteq\bigcup_{\varphi}C_{\varphi} then all elements of SS have the same type over AA in 𝔐^\hat{\mathfrak{M}}, and similarly all elements of GG.

Lemma 6.5.

Assume that φ(x,y¯)ΦS\varphi(x,\bar{y})\in\Phi_{S} is stable. Then the following are equivalent for an instance φ(x,b¯)\varphi(x,\bar{b}):

  1. (i)

    The condition φ(x,b¯)=0\varphi(x,\bar{b})=0 is generic in SS.

  2. (ii)

    The condition φ^(x,e,[b¯])=0\hat{\varphi}(x,e,[\bar{b}])=0 does not fork in 𝔐^\hat{\mathfrak{M}} over \varnothing.

  3. (iii)

    The condition φ^(x,e,[b¯])=0\hat{\varphi}(x,e,[\bar{b}])=0 does not fork in 𝔐^\hat{\mathfrak{M}} over [b¯][\bar{b}].

Proof.

Recall that the ^\hat{\mathcal{L}}-formula φ^(xS,yGzφ)\hat{\varphi}(x_{S},y_{G}z_{\varphi}) with this (or any other) partition of the variables is stable in 𝔐^\hat{\mathfrak{M}}. For ε>0\varepsilon>0 let Xε={sS:φ(s,b¯)ε}X_{\varepsilon}=\{s\in S\colon\varphi(s,\bar{b})\leq\varepsilon\}. By Lemma 6.2, the condition φ(x,b¯)=0\varphi(x,\bar{b})=0 is generic if and only if XεX_{\varepsilon} is a generic set for all ε>0\varepsilon>0.

  • (i) \Longrightarrow (ii).

    Assume first that φ(x,b¯)=0\varphi(x,\bar{b})=0 is generic in SS, i.e., that the set XεX_{\varepsilon} is generic for every ε>0\varepsilon>0. Find giGg_{i}\in G such that S=i<ngiXεS=\bigcup_{i<n}g_{i}X_{\varepsilon}, and find sSs\in S such that tpφ^(s/[b¯]g<n)\operatorname{tp}_{\hat{\varphi}}(s/[\bar{b}]g_{<n}) does not fork over \varnothing (in symbols sφ^[b¯]g<ns\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{${}^{\hat{\varphi}}$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{${}^{\hat{\varphi}}$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{${}^{\hat{\varphi}}$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{${}^{\hat{\varphi}}$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}\,[\bar{b}]g_{<n}). Since si<ngiXεs\in\bigcup_{i<n}g_{i}X_{\varepsilon} we may assume that sg0Xεs\in g_{0}X_{\varepsilon}, so φ^(s,g0,[b¯])=φ(g01s,b¯)ε\hat{\varphi}(s,g_{0},[\bar{b}])=\varphi(g_{0}^{-1}s,\bar{b})\leq\varepsilon. Thus φ^(x,g0,[b¯])ε\hat{\varphi}(x,g_{0},[\bar{b}])\leq\varepsilon does not fork over \varnothing. Applying θg01\theta_{g_{0}^{-1}} we see that φ^(x,e,[b¯])ε\hat{\varphi}(x,e,[\bar{b}])\leq\varepsilon does not fork over \varnothing either. It follows that φ^(x,e,[b¯])=0\hat{\varphi}(x,e,[\bar{b}])=0 does not fork over \varnothing.

  • (iii) \Longrightarrow (iv).

    Immediate.

  • (v) \Longrightarrow (i).

    Assume now that φ^(x,e,[b¯])=0\hat{\varphi}(x,e,[\bar{b}])=0 does not fork over [b¯][\bar{b}]. By Proposition 2.17 there are gnGg_{n}\in G for nn\in\mathbb{N} and a faithful combination ψ(x,[b¯])=θ(φ^(x,gn,[b¯]))n\psi(x,[\bar{b}])=\theta\bigl{(}\hat{\varphi}(x,g_{n},[\bar{b}])\bigr{)}_{n\in\mathbb{N}} which is definable over [b¯][\bar{b}] and such that ψ(x,[b¯])=0\psi(x,[\bar{b}])=0 is consistent. Since 𝔐^\hat{\mathfrak{M}} is saturated for quantifier-free types involving only φ^\hat{\varphi}, there is sSs\in S such that ψ(s,[b¯])=0\psi(s,[\bar{b}])=0. Since all elements of SS have the same type over [b¯][\bar{b}] in 𝔐^\hat{\mathfrak{M}}, we see that ψ(s,[b¯])=0\psi(s,[\bar{b}])=0 for all sSs\in S. Assume (toward a contradiction) that there exists ε>0\varepsilon>0 such that φ(x,b¯)ε\varphi(x,\bar{b})\leq\varepsilon is not generic. By compactness we can find sSs\in S such that φ(gn1s,b¯)ε\varphi(g_{n}^{-1}s,\bar{b})\geq\varepsilon for all nn, i.e., φ^(s,gn,[b¯])ε\hat{\varphi}(s,g_{n},[\bar{b}])\geq\varepsilon. Since the combination above was faithful we get ψ(s,[b¯])ε>0\psi(s,[\bar{b}])\geq\varepsilon>0, a contradiction. ∎

Lemma 6.6.

Assume that φ(x,y¯)ΦS\varphi(x,\bar{y})\in\Phi_{S} is stable and that φ(x,b¯)=0\varphi(x,\bar{b})=0 is a generic condition in SS. Then it does not fork over \varnothing.

Proof.

By Proposition 4.3 it will be enough to show that φ(x,b¯)=0\varphi(x,\bar{b})=0 does not divide over \varnothing. For this purpose let (b¯n)n(\bar{b}_{n})_{n\in\mathbb{N}} be any indiscernible sequence with b¯0=b¯\bar{b}_{0}=\bar{b}. Since edcl()e\in\operatorname{dcl}(\varnothing), the sequence (e,b¯n)n(e,\bar{b}_{n})_{n\in\mathbb{N}} is indiscernible as well, and thus the sequence (e,[b¯n])n(e,[\bar{b}_{n}])_{n\in\mathbb{N}} is indiscernible in 𝔐^\hat{\mathfrak{M}}. On the other hand, since the condition φ(x,b¯)=0\varphi(x,\bar{b})=0 is generic, by Lemma 6.5 the condition φ^(x,e,[b¯])=0\hat{\varphi}(x,e,[\bar{b}])=0 does not fork over \varnothing, so {φ^(x,e,[b¯n])}n\{\hat{\varphi}(x,e,[\bar{b}_{n}])\}_{n\in\mathbb{N}} is consistent. Since 𝔐^\hat{\mathfrak{M}} is saturated for such formulae, there is sSs\in S such that φ^(s,e,[b¯n])=0\hat{\varphi}(s,e,[\bar{b}_{n}])=0, i.e., φ(s,b¯i)=0\varphi(s,\bar{b}_{i})=0, for all nn, as desired. ∎

From now on we assume that TT is stable.

Proposition 6.7.

Let π(x)\pi(x) be a partial type over AA. Then π\pi is generic if and only if it extends to a complete generic type over AA, i.e., if and only if [π]SS(A)[\pi]\subseteq\operatorname{S}_{S}(A) 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 π\pi 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 π\pi contains a generic type, and we may further restrict our attention to logical neighbourhoods defined by a single condition φ(x,b¯)=0\varphi(x,\bar{b})=0. Since π\pi is generic in SS so is φ(x,b¯)=0\varphi(x,\bar{b})=0. By Lemma 6.5 φ^(x,e,[b¯])=0\hat{\varphi}(x,e,[\bar{b}])=0 does not fork over \varnothing in 𝔐^\hat{\mathfrak{M}}. By Corollary 2.4 there exists a type p^Sx(𝔐^)\hat{p}\in\operatorname{S}_{x}(\hat{\mathfrak{M}}) such that φ^(x,e,[b¯])p^=0\hat{\varphi}(x,e,[\bar{b}])^{\hat{p}}=0 and in addition pψ^p{\restriction}_{\hat{\psi}} does not fork over \varnothing for every formula ψΦS\psi\in\Phi_{S}. Let

p(x)={ψ(x,c¯)=ψ^(x,e,[c¯])p^}ψΦS,c¯𝔐.\displaystyle p(x)=\bigl{\{}\psi(x,\bar{c})=\hat{\psi}(x,e,[\bar{c}])^{\hat{p}}\bigr{\}}_{\psi\in\Phi_{S},\bar{c}\in\mathfrak{M}}.

This type is approximately finitely realised in 𝔐\mathfrak{M} (since p^\hat{p} is in 𝔐^\hat{\mathfrak{M}}) and therefore consistent. By Lemma 6.5 every condition in pp is generic (since p^\hat{p} does not fork over \varnothing), and by Lemma 6.2, pp is generic, and so is pAp{\restriction}_{A}. We have thus found a generic type pA[φ(x,b¯)=0]p{\restriction}_{A}\in[\varphi(x,\bar{b})=0] and the proof is complete. ∎

Proposition 6.8.

Assume ABA\subseteq B. Then a type pSS(B)p\in\operatorname{S}_{S}(B) is generic if and only if it does not fork over AA and pAp{\restriction}_{A} is generic. In particular, a generic type does not fork over \varnothing.

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 pSS(B)p\in\operatorname{S}_{S}(B) is generic then clearly so is pAp{\restriction}_{A}, and by the previous paragraph pp does not fork over AA. For the converse, assume that pSS(B)p\in\operatorname{S}_{S}(B) does not fork over AA and p0=pAp_{0}=p{\restriction}_{A} is generic. Replacing pp with a non forking extension we may assume that B=𝔐B=\mathfrak{M}. By Proposition 6.7 there is p1SS(𝔐)p_{1}\in\operatorname{S}_{S}(\mathfrak{M}) extending p0p_{0} which is generic, and by what we have just shown it is also non forking over AA. Since pA=p0=p1Ap{\restriction}_{A}=p_{0}=p_{1}{\restriction}_{A} there is fAut(𝔐/A)f\in\operatorname{Aut}(\mathfrak{M}/A) sending p1acleq(A)p_{1}{\restriction}_{\operatorname{acl}^{eq}(A)} to pacleq(A)p{\restriction}_{\operatorname{acl}^{eq}(A)}, and therefore p1p_{1} to pp. Thus pp is generic as well. ∎

We can also complement Lemma 6.3:

Proposition 6.9.

The action of GG on the set of generic types in SS(𝔐)\operatorname{S}_{S}(\mathfrak{M}) is transitive.

Proof.

Let p,qSS(𝔐)p,q\in\operatorname{S}_{S}(\mathfrak{M}) be two generic types. Define

p^={φ^(x,g,[b¯])=φ(g1x,[b¯])p}φΦS,b¯𝔐,gG,\displaystyle\hat{p}=\{\hat{\varphi}(x,g,[\bar{b}])=\varphi(g^{-1}x,[\bar{b}])^{p}\}_{\varphi\in\Phi_{S},\bar{b}\in\mathfrak{M},g\in G},

and define q^\hat{q} similarly. Let C^=(acleq())𝔐^\hat{C}=\bigl{(}\operatorname{acl}^{eq}(\varnothing)\bigr{)}^{\hat{\mathfrak{M}}} and let p^0=p^C\hat{p}_{0}=\hat{p}{\restriction}_{C}, q^0=q^C\hat{q}_{0}=\hat{q}{\restriction}_{C}. Since 𝔐^\hat{\mathfrak{M}} is saturated for formulae of this form we may realise p^0\hat{p}_{0} and q^0\hat{q}_{0} in 𝔐^\hat{\mathfrak{M}}, and by transitivity there exists hGh\in G such that θhp^0q^0\theta_{h}\hat{p}_{0}\cup\hat{q}_{0} is realised. Since θh\theta_{h} is an automorphism of 𝔐^\hat{\mathfrak{M}} we must have q^0=θhp^0=(θhp^)C\hat{q}_{0}=\theta_{h}\hat{p}_{0}=(\theta_{h}\hat{p}){\restriction}_{C}. In addition, neither of p^\hat{p}, q^\hat{q} or θhp^\theta_{h}\hat{p} forks over \varnothing, whereby θhp^=q^\theta_{h}\hat{p}=\hat{q}, i.e., hp=qhp=q. ∎

Theorem 6.10.

Let GG be a type-definable group in a stable theory TT, acting type-definably and transitively on a type-definable set SS.

  1. (i)

    If gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s (where gGg\in G, sSs\in S) and gg is left-generic over AA then gsgs is generic over AA and gsAsgs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s.

  2. (ii)

    An element sSs\in S is generic if and only if gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s implies gsAggs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}g for every gGg\in G. Moreover, in this case gsgs is generic over AA as well.

  3. (iii)

    An element gGg\in G is left-generic over AA if and only if g1g^{-1} is.

  4. (iv)

    An element gGg\in G is left-generic if and only if it is right-generic (over AA). From now on we shall only speak of generic elements and types in GG.

  5. (v)

    An element gGg\in G is generic over AA if and only if it is generic over \varnothing and gAg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}A.

Proof.

We use Proposition 6.8 repeatedly.

For the first item, let sSs\in S, gGg\in G, and assume that gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s. If gg is left-generic over AA then it is left-generic over A,sA,s. By Lemma 6.4 gsgs is generic over A,sA,s. It follows that gsgs is generic over AA and that gsA,sgs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}A,s, 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 sAgs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}g implies gsA,ggs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}A,g for all gg. We may choose gg which is left-generic over AA such that gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s. Then g1Agsg^{-1}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}gs by assumption, gsgs is generic over AA by the first item, and s=g1gss=g^{-1}gs is generic over AA by the moreover part.

For the third item, let gGg\in G be left-generic over AA. Choose hGh\in G left-generic over AA such that gAhg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}h. By the first item ghgh is generic over AA and ghAhgh\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}h. This can be re-written as hAh1g1h\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}h^{-1}g^{-1}. By the first item again, g1=hh1g1g^{-1}=hh^{-1}g^{-1} is left-generic over AA. Notice that g1g^{-1} is left-generic if and only if gg 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 AA, a group element gG(A)g\in G(A) induces a homeomorphism Lg:pgpL_{g}\colon p\mapsto gp on SS(A)\operatorname{S}_{S}(A). It is also not difficult to check that LgLh=LghL_{g}\circ L_{h}=L_{gh}, whence a group action of G(A)G(A) on SS(A)\operatorname{S}_{S}(A). In addition, we have seen that it restricts to an action by homeomorphism of G(A)G(A) on the set of generic types in SS(A)\operatorname{S}_{S}(A).

Specifically, we obtain an action of G=G(𝔐)G=G(\mathfrak{M}) on SS(𝔐)\operatorname{S}_{S}(\mathfrak{M}). The stabiliser of a type pSS(𝔐)p\in\operatorname{S}_{S}(\mathfrak{M}) under this action is Stab(p)={gG:gp=p}G\operatorname{Stab}(p)=\{g\in G\colon gp=p\}\leq G. For a stationary type pSS(A)p\in\operatorname{S}_{S}(A) we define Stab(p)=Stab(p𝔐)\operatorname{Stab}(p)=\operatorname{Stab}(p{\restriction}^{\mathfrak{M}}).

Proposition 6.11.

Let pSS(A)p\in\operatorname{S}_{S}(A) be stationary. Then stabiliser Stab(p)\operatorname{Stab}(p) is a sub-group of GG type-definable over Cb(p)\mathrm{Cb}(p).

Moreover, assume that sps\vDash p, gGg\in G and gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s. Then gStab(p)g\in\operatorname{Stab}(p) if and only if gspgs\vDash p.

Proof.

We may assume that pSS(𝔐)p\in\operatorname{S}_{S}(\mathfrak{M}).

Let φ(x,z¯)\varphi(x,\bar{z}) be a formula, xx in the sort of SS. Let yy be a variable in the sort of GG. Then φ(yx,z¯)\varphi(yx,\bar{z}) is a definable predicate on G×S×sort of z¯G\times S\times\langle\text{sort of }\bar{z}\rangle, i.e., a continuous function SG,S,z¯(T)[0,1]\operatorname{S}_{G,S,\bar{z}}(T)\to[0,1]. By Tietze’s Extension Theorem this extends to a continuous function Sx,y,z¯(T)[0,1]\operatorname{S}_{x,y,\bar{z}}(T)\to[0,1]. For clarity we shall use φ(yx,z¯)\varphi(yx,\bar{z}) to denote the corresponding definable predicate.

Once this technical preliminary is taken care of we see that Stab(p)\operatorname{Stab}(p) is defined by the following axiom scheme:

π(y)={supz¯|dpφ(x,z¯)dpφ(yx,z¯)|=0}φΦS.\displaystyle\pi(y)=\Bigl{\{}\sup_{\bar{z}}|d_{p}\varphi(x,\bar{z})-d_{p}\varphi(yx,\bar{z})|=0\Bigr{\}}_{\varphi\in\Phi_{S}}.

The moreover part easily follows. ∎

Lemma 6.12.

Let H<GH<G be a type-definable subgroup of bounded index, say with parameters in AA, and let gHg\in H. Then gg is generic over AA in GG if and only if it is generic over AA in HH.

Proof.

Naming AA in the language we may assume that A=A=\varnothing. Since HH has bounded index we may enumerate its cosets {giH}i<λ\{g_{i}H\}_{i<\lambda}. Let h0Gh_{0}\in G be generic over {gi}i<λ\{g_{i}\}_{i<\lambda}. Then h0giHh_{0}\in g_{i}H for some ii, and h1=gi1h0Hh_{1}=g_{i}^{-1}h_{0}\in H is generic in GG. Now let h2h_{2} be generic in HH. Without loss of generality we may assume that h1h2h_{1}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}h_{2}. Then h1h2Hh_{1}h_{2}\in H is generic both in HH and in GG and h1h2h1h_{1}h_{2}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}h_{1}. Thus h2=h11h1h2h_{2}=h_{1}^{-1}h_{1}h_{2} is generic in GG as well. We have thus shown that every generic of HH is a generic of GG. A similar argument shows that every generic of GG in HH is generic in HH. ∎

Proposition 6.13.

A type pSS(A)p\in\operatorname{S}_{S}(A) is generic if and only if Stab(p)\operatorname{Stab}(p) has bounded index in GG.

Proof.

There are only boundedly many generic types over 𝔐\mathfrak{M}, since they do not fork over \varnothing and therefore determined by their restriction to acleq()\operatorname{acl}^{eq}(\varnothing). In addition, the action of GG on SS(𝔐)\operatorname{S}_{S}(\mathfrak{M}) restrict to an action of GG on the space of generic types, so the stabiliser of a generic type must be of bounded index.

Conversely, assume Stab(p)\operatorname{Stab}(p) has bounded index, and let sps\vDash p. Then there exists gStab(p)g\in\operatorname{Stab}(p) which is generic in GG over AA, and we may further assume that gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s. Then gspgs\vDash p is generic over AA, i.e., pp is generic. ∎

Since GG acts transitively on the generic types over 𝔐\mathfrak{M}, the stabilisers of generic types are all conjugate. It is also not difficult to check that if pSS(𝔐)p\in\operatorname{S}_{S}(\mathfrak{M}) is generic, qSG(𝔐)q\in\operatorname{S}_{G}(\mathfrak{M}) is a generic type of Stab(p)\operatorname{Stab}(p) (and therefore of GG), and spacleq()s\vDash p{\restriction}_{acl^{eq}(\varnothing)}, then qs=pqs=p. If qq^{\prime} is any other generic of GG then (since GG acts transitively on its own generic types, on the left as well as on the right) there exists gGg\in G such that q=qgq=q^{\prime}g and p=q(gs)p=q^{\prime}(gs). Thus the right action of SS on GG send each and every generic type of GG onto the generic types of SS, complementing Lemma 6.4.

Theorem 6.14.

Let GG be a type-definable group in a stable theory, say over \varnothing. Then GG admits a smallest type-definable group of bounded index (over any set of parameters), called the connected component of GG, and denoted G0G^{0}. It has the following additional properties.

  1. (i)

    The connected component G0G^{0} is a normal subgroup of GG, type-definable over \varnothing.

  2. (ii)

    The stabiliser of every generic type is equal to G0G^{0}.

  3. (iii)

    Each coset gG0gG^{0} contains a unique generic type over 𝔐\mathfrak{M}.

  4. (iv)

    The generic type of G0G^{0} is definable over \varnothing.

  5. (v)

    If pSG(A)p\in\operatorname{S}_{G}(A) is any stationary generic type over a small set then G0={g1h:g,hp}G^{0}=\{g^{-1}h\colon g,h\vDash p\}.

Proof.

We start by constructing G0G^{0} and proving the second item. Since left generic and right generic are the same, the action of GG on the generic types is transitive on either side. In particular, if p,qSG(𝔐)p,q\in\operatorname{S}_{G}(\mathfrak{M}) are generic then there exists gGg\in G such that q=pgq=pg, and thus Stab(p)=Stab(q)\operatorname{Stab}(p)=\operatorname{Stab}(q). Let this unique stabiliser of generic types be denoted G0G^{0}. Then G0G^{0} is type-definable, and since GG \varnothing-invariant, so is G0G^{0}, and we may conclude that G0G^{0} is type-definable over \varnothing as well. We also already know that G0G^{0} has bounded index in GG. Assume now that HG0H\leq G^{0} is another type-definable subgroup of bounded index, say over \varnothing (otherwise name the parameters in the language). Then there exists pSH(𝔐)p\in\operatorname{S}_{H}(\mathfrak{M}) generic in GG, so Stab(p)=G0\operatorname{Stab}(p)=G^{0}, whereby G0HG^{0}\subseteq H. Thus G0G^{0} is indeed the smallest type-definable subgroup of GG of bounded index. Notice that G0gG0g1G^{0}\cap gG^{0}g^{-1} is also type-definable of bounded index for every gGg\in G, so G0G^{0} is normal in GG. This concludes the proof of the first two items.

Let pSG(𝔐)p\in\operatorname{S}_{G}(\mathfrak{M}) be generic in G0G^{0}. Since G0=Stab(p)G^{0}=\operatorname{Stab}(p) acts transitively on its generic types, pp is the unique generic type in G0G^{0}. It follows that a coset gG0gG^{0} contains a unique generic type gpgp. The uniqueness of the generic type of G0G^{0} implies that it is \varnothing-invariant, and therefore definable over \varnothing.

Finally, let pSG(A)p\in\operatorname{S}_{G}(A) be any stationary generic type over a small set. Then p𝔐p{\restriction}^{\mathfrak{M}} is the unique generic type in some coset gG0gG^{0}. It follows that gG0gG^{0} is AA-invariant, so pxgG0p\vdash x\in gG^{0}. Thus {g1h:g,hp}G0\{g^{-1}h\colon g,h\vDash p\}\subseteq G^{0}. Conversely, let gG0g\in G^{0}, and let hph\vDash p, gAhg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}h. Since G0G^{0} must also be the right-stabiliser of pp we have hgphg\vDash p as well, and g=h1(hg)g=h^{-1}(hg), as desired. ∎

It follows that GG is connected (i.e., G=G0G=G^{0}) 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 SS is generic if and only if the corresponding type in 𝔐^\hat{\mathfrak{M}} is a non forking extension of the unique type over \varnothing, i.e., if its φ^\hat{\varphi}-type has the same Cantor-Bendixson ranks as all of SS for every φΦS\varphi\in\Phi_{S}. Thus the various ε\varepsilon-φ^\hat{\varphi}-Cantor-Bendixson ranks play the role of stratified local ranks characterising genericity. In a superstable (and even more so in an 0\aleph_{0}-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 d(x,y)εd(x,y)\leq\varepsilon via conjunction and existential quantification. It is a general fact that if π(x¯,y¯)\pi(\bar{x},\bar{y}) is a partial type then the property y¯π(x¯,y¯)\exists\bar{y}\,\pi(\bar{x},\bar{y}), 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 π(x¯)\pi(\bar{x}) be a partial type and ε0\varepsilon\geq 0 a real number. We define π(x¯ε)\pi(\bar{x}^{\varepsilon}) to be the partial type expressing that y¯(π(y¯)&d(x¯,y¯)ε)\exists\bar{y}\,\bigl{(}\pi(\bar{y})\,\&\,d(\bar{x},\bar{y})\leq\varepsilon\bigr{)}, i.e., that π\pi is realised in the ε\varepsilon-neighbourhood of x¯\bar{x}. For a tuple a¯\bar{a} we shall use a¯ε\bar{a}^{\varepsilon} as a notational representation for the somewhat vague concept of “a¯\bar{a} known up to distance ε\varepsilon”, so in particular a¯0\bar{a}^{0} is just another representation for a¯\bar{a}. Accordingly, if p(x¯)=tp(a¯/C)p(\bar{x})=\operatorname{tp}(\bar{a}/C) then we define tp(a¯ε/C)=p(x¯ε)\operatorname{tp}(\bar{a}^{\varepsilon}/C)=p(\bar{x}^{\varepsilon}), so in particular tp(a¯/C)=tp(a¯0/C)=ε>0tp(a¯ε/C)\operatorname{tp}(\bar{a}/C)=\operatorname{tp}(\bar{a}^{0}/C)=\bigwedge_{\varepsilon>0}\operatorname{tp}(\bar{a}^{\varepsilon}/C).

Definition 6.15.

To an arbitrary theory TT we define κ(T)\kappa(T) to be the least infinite cardinal, if such exists, such that for every complete type p(x¯)p(\bar{x}) over a set AA, and for every ε>0\varepsilon>0, there is a subset A0AA_{0}\subseteq A such that p(x¯ε)p(\bar{x}^{\varepsilon}) does not divide over A0A_{0}.

We say that TT is simple if κ(T)||+\kappa(T)\leq|\mathcal{L}|^{+}, and that it is supersimple if κ(T)=0\kappa(T)=\aleph_{0}.

It follows from our earlier results that every stable theory is simple. It is true (but we shall not require it) that if TT is not simple then κ(T)=\kappa(T)=\infty.

Definition 6.16.

We say that TT is λ\lambda-stable if for every set AA, |A|λ|A|\leq\lambda, the metric density character of Sn(A)\operatorname{S}_{n}(A) is at most λ\lambda. We define λ0(T)\lambda_{0}(T) to be the least (infinite) cardinal of stability for TT. We say that TT is superstable if it is λ\lambda-stable for all λ\lambda 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 TT is stable if and only if it is λ||\lambda^{|\mathcal{L}|}-stable for all λ\lambda. In particular, TT is stable if and only if λ0(T)<\lambda_{0}(T)<\infty, in which case λ0(T)2||\lambda_{0}(T)\leq 2^{|\mathcal{L}|}. 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 TT is λ\lambda-stable if and only if λ<κ(T)=λλ0(T)\lambda^{<\kappa(T)}=\lambda\geq\lambda_{0}(T).

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 κ(T)\kappa(T) we wish to define global forking ranks SUε\operatorname{SU}_{\varepsilon} for ε>0\varepsilon>0 such that SUε(a¯/C)>SUε(a¯/Cb¯)\operatorname{SU}_{\varepsilon}(\bar{a}/C)>\operatorname{SU}_{\varepsilon}(\bar{a}/C\bar{b}) if and only if a¯\bar{a}ε\varepsilon-depends” on b¯\bar{b} over CC. We have quite a bit of liberty in choosing what “ε\varepsilon-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 tp(a¯ε/Cb¯)\operatorname{tp}(\bar{a}^{\varepsilon}/C\bar{b}) forks over CC, or if tp(a¯ε/Cb¯)tp(a¯/acleq(C))\operatorname{tp}(\bar{a}^{\varepsilon}/C\bar{b})\wedge\operatorname{tp}(\bar{a}/\operatorname{acl}^{eq}(C)) forks over CC. 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 a¯\bar{a} and b¯\bar{b} be tuples, CC a set and ε>0\varepsilon>0. We keep in mind that a¯ε\bar{a}^{\varepsilon} represents “a¯\bar{a} known up to distance ε\varepsilon”.

  1. (i)

    We say that an indiscernible sequence (b¯n)n(\bar{b}_{n})_{n\in\mathbb{N}} could be in tp(b¯/a¯εC)\operatorname{tp}(\bar{b}/\bar{a}^{\varepsilon}C) if there are CC^{\prime} and a sequence (a¯n)n(\bar{a}_{n})_{n} such that (a¯nb¯n)n(\bar{a}_{n}\bar{b}_{n})_{n} is CC^{\prime}-indiscernible, a¯nb¯nCa¯b¯C\bar{a}_{n}\bar{b}_{n}C^{\prime}\equiv\bar{a}\bar{b}C and d(a¯0,a¯1)εd(\bar{a}_{0},\bar{a}_{1})\leq\varepsilon.

  2. (ii)

    We say that a¯εCb¯\bar{a}^{\varepsilon}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b} if every indiscernible sequence in tp(b¯/C)\operatorname{tp}(\bar{b}/C) could be in tp(b¯/a¯εC)\operatorname{tp}(\bar{b}/\bar{a}^{\varepsilon}C).

  3. (iii)

    We define SUε(a¯/C)\operatorname{SU}_{\varepsilon}(\bar{a}/C) as may be expected: SUε(a¯/C)α+1\operatorname{SU}_{\varepsilon}(\bar{a}/C)\geq\alpha+1 if and only if there is b¯\bar{b} such that a¯ε Cb¯\bar{a}^{\varepsilon}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\not$\kern 6.34859pt\hss}\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\not$\kern 5.54167pt\hss}\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b} and SUε(a¯/Cb¯)α\operatorname{SU}_{\varepsilon}(\bar{a}/C\bar{b})\geq\alpha. (In [Ben06] the notation SU(a¯ε/C)\operatorname{SU}(\bar{a}^{\varepsilon}/C) was used.)

A few justifications for these definitions may be in place. Indeed, for ε=0\varepsilon=0 it is easy to see that a sequence (b¯n)(\bar{b}_{n}) could be in tp(b¯/a¯0C)\operatorname{tp}(\bar{b}/\bar{a}^{0}C) if and only if it admits a conjugate which is in tp(b¯/a¯C)\operatorname{tp}(\bar{b}/\bar{a}C) in the ordinary sense. By a compactness argument, this is further equivalent to the property that (b¯n)(\bar{b}_{n}) could be in tp(b¯/a¯εC)\operatorname{tp}(\bar{b}/\bar{a}^{\varepsilon}C) for every ε>0\varepsilon>0.

Next, we wish to justify the notation \mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}. 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 aa, bb and cc 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 bacb\equiv_{a}c. Then abca\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b}c if and only if acba\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{c}b.

Proof.

Assume not, say acba\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{c}b but a bca\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\not$\kern 6.34859pt\hss}\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\not$\kern 5.54167pt\hss}\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b}c. We construct by induction a sequence (bn)n(b_{n})_{n\in\mathbb{N}} such that bnbn+1abcb_{n}b_{n+1}\equiv_{a}bc and abn+1bnb<nab_{n+1}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{n}}b_{<n} for all nn. We start with b0=bb_{0}=b, b1=cb_{1}=c. Then for each nn we can choose bn+2b_{n+2} such that bn+1bn+2abcb_{n+1}b_{n+2}\equiv_{a}bc and bn+2abn+1bnb_{n+2}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{ab_{n+1}}b_{\leq n}. Our induction hypothesis tells us that abn+1bnb<nab_{n+1}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{n}}b_{<n} and abn+1bna\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{n+1}}b_{n}. By standard independence calculus we obtain abn+1bna\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{n+1}}b_{\leq n}, whence abn+2bn+1bnab_{n+2}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{n+1}}b_{\leq n}, as desired.

Since a bca\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\not$\kern 6.34859pt\hss}\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\not$\kern 5.54167pt\hss}\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b}c, there exists a bb-indiscernible sequence (ck)k(c^{k})^{k}, starting with c0=cc^{0}=c, such that there exists no aa^{\prime} satisfying ackbaca^{\prime}c^{k}\equiv_{b}ac for all kk. For each nn we may choose a copy (cnk)k(c_{n}^{k})_{k} such that (cnk)k,bn+1,bn(ck)k,c,b(c_{n}^{k})_{k},b_{n+1},b_{n}\equiv(c^{k})_{k},c,b, so in particular this copy starts with cn0=bn+1c_{n}^{0}=b_{n+1} and is indiscernible over bnb_{n}. Since bn+1bnb<nb_{n+1}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{n}}b_{<n}, we may choose (cnk)k(c_{n}^{k})_{k} to be indiscernible over bnb_{\leq n}.

Let us now define cn=bn+1c_{n}=b_{n+1} for all nn and consider the sequence (bncn)n(b_{n}c_{n})_{n}. Applying compactness, we can find for arbitrarily big λ\lambda a sequence (bici)i<λ(b_{i}c_{i})_{i<\lambda} in tp(bc/a)\operatorname{tp}(bc/a), such that for each ii there exists a sequence (cik)k(c_{i}^{k})_{k} which is indiscernible over bic<ib_{\leq i}c_{<i} such that (cik)k,ci,bi(ck)k,c,b(c_{i}^{k})_{k},c_{i},b_{i}\equiv(c^{k})_{k},c,b, so in particular ci0=cic_{i}^{0}=c_{i}. Each of these sequences witnesses that a b<ic<ibicia\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\not$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\not$\kern 6.34859pt\hss}\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\not$\kern 5.54167pt\hss}\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{b_{<i}c_{<i}}b_{i}c_{i}, contradicting the local character. ∎

We can now prove:

Fact 6.21 ([Ben06, Lemma 1.13]).

Assume that TT is simple (or stable). For a¯ε\bar{a}^{\varepsilon}, CC and b¯\bar{b}, the following imply one another from top to bottom:

  1. (i)

    a¯εCb¯\bar{a}^{\varepsilon}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b}.

  2. (ii)

    There is a Morley sequence for b¯\bar{b} over CC which could be in tp(b¯/a¯εC)\operatorname{tp}(\bar{b}/\bar{a}^{\varepsilon}C).

  3. (iii)

    tp(a¯ε/b¯C)\operatorname{tp}(\bar{a}^{\varepsilon}/\bar{b}C) does not fork over CC.

  4. (iv)

    a¯2εCb¯\bar{a}^{2\varepsilon}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b}.

Proof.

We only need to provide a proof for (ii) \Longrightarrow (iii), for which the original proof used local ranks.

Indeed let (a¯n)n(\bar{a}_{n})_{n} and CC^{\prime} witness that a Morley sequence (b¯n)(\bar{b}_{n}) could be in tp(b¯/a¯εC)\operatorname{tp}(\bar{b}/\bar{a}^{\varepsilon}C). We may assume that (a¯nb¯n)n(\bar{a}_{n}\bar{b}_{n})_{n} is indiscernible over CCCC^{\prime}, and we may further assume that for some a~\tilde{a}, the pair a~b¯\tilde{a}\bar{b} continues the sequence (a¯nb¯n)n(\bar{a}_{n}\bar{b}_{n})_{n} indiscernibly over CCCC^{\prime}. Standard arguments regarding indiscernibility provide that b¯(b¯n)nCCa0\bar{b}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{(\bar{b}_{n})_{n}}CC^{\prime}a_{0}. Since b¯C(b¯n)n\bar{b}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}(\bar{b}_{n})_{n} we obtain b¯CCa¯0\bar{b}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}C^{\prime}\bar{a}_{0}. By Lemma 6.20 we have b¯CC\bar{b}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C^{\prime}}C and thus b¯Ca¯0\bar{b}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C^{\prime}}\bar{a}_{0}. Let ff be an automorphism sending a~b¯C\tilde{a}\bar{b}C^{\prime} to a¯b¯C\bar{a}\bar{b}C. Then b¯Cf(a¯0)\bar{b}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}f(\bar{a}_{0}) and d(f(a¯0),a¯)=d(a¯0,a~)εd(f(\bar{a}_{0}),\bar{a})=d(\bar{a}_{0},\tilde{a})\leq\varepsilon. Thus tp(a¯ε/b¯C)\operatorname{tp}(\bar{a}^{\varepsilon}/\bar{b}C) does not fork over CC, as desired. ∎

Thus in particular a¯0Cb¯\bar{a}^{0}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b} if and only if tp(a¯/b¯C)\operatorname{tp}(\bar{a}/\bar{b}C) does not fork over CC, i.e., if and only if a¯Cb¯\bar{a}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b}, justifying our notation. By earlier observations, this is further equivalent to a¯εCb¯\bar{a}^{\varepsilon}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}\bar{b} holding for all ε>0\varepsilon>0. It is further shown in [Ben06] that TT is supersimple if and only if SUε(a¯/B)\operatorname{SU}_{\varepsilon}(\bar{a}/B) is ordinal for every finite tuple a¯\bar{a} and ε>0\varepsilon>0. Moreover, in a supersimple theory TT, SUε\operatorname{SU}_{\varepsilon} ranks characterise independence: a¯CB\bar{a}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{C}B if and only if SUε(a¯/C)=SUε(a¯/BC)\operatorname{SU}_{\varepsilon}(\bar{a}/C)=\operatorname{SU}_{\varepsilon}(\bar{a}/BC) for all ε>0\varepsilon>0.

Since our global ranks depend (inevitably) on a metric resolution parameter ε\varepsilon we may only hope to characterise genericity in case the metric is invariant under the group action, i.e., if the action of each gGg\in G on SS is an isometry.

We have seen that if gg is generic over s,As,A then gsgs is generic over AA. We now prove a converse:

Lemma 6.22.

Assume G,S\langle G,S\rangle is a type-definable transitive group action in a stable theory TT, sSs\in S generic over a set AA, tSt\in S satisfying tAst\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s. Then there is gGg\in G, gAtg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}t such that gs=tgs=t. Moreover, gg can be chosen generic over AA (i.e., over AtAt).

Proof.

We may assume A=A=\varnothing. First choose gGg\in G generic, gs,tg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}s,t. Then ss is generic over g,tg,t, so gsg,tgs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}g,t By standard independence calculus we obtain ggs,tg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}gs,t. Since the action is transitive we can find hGh\in G such that hgs=thgs=t, and we may take it so that hgs,tgh\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{gs,t}g. Then gg is generic over t,gs,ht,gs,h, and so is ghgh, and in particular hgthg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}t. Then g=hgg^{\prime}=hg is generic over tt as required. ∎

Theorem 6.23.

Assume G,S\langle G,S\rangle is a type-definable transitive group action with an invariant metric in a superstable continuous theory TT, pSS(A)p\in\operatorname{S}_{S}(A). Then pp is generic if and only if SUε(p)=SUε(S)=sup{SUε(q):qSS()}\operatorname{SU}_{\varepsilon}(p)=\operatorname{SU}_{\varepsilon}(S)=\sup\{\operatorname{SU}_{\varepsilon}(q)\colon q\in\operatorname{S}_{S}(\varnothing)\} for all ε>0\varepsilon>0. In particular, types of maximal SUε\operatorname{SU}_{\varepsilon}-rank exist.

Proof.

We may assume that A=A=\varnothing. We shall use the fact that if pSn(B)p\in\operatorname{S}_{n}(B), qSm(B)q\in\operatorname{S}_{m}(B) and f:p(𝔐)q(𝔐)f\colon p(\mathfrak{M})\to q(\mathfrak{M}) is BB-definable and isometric then SUε(p)=SUε(q)\operatorname{SU}_{\varepsilon}(p)=\operatorname{SU}_{\varepsilon}(q) for all ε>0\varepsilon>0. The proof of this fact is left as an exercise to the reader.

Let sps\vDash p, and assume first that pp is generic. Let tSt\in S realise an arbitrary type over \varnothing. We may nonetheless assume that tst\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}s. By the Lemma there exists gtg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}t such that gs=tgs=t. Since multiplication by gg is isometric we obtain SUε(s)SUε(s/g)=SUε(t/g)=SUε(t)=SUε(q)\operatorname{SU}_{\varepsilon}(s)\geq\operatorname{SU}_{\varepsilon}(s/g)=\operatorname{SU}_{\varepsilon}(t/g)=\operatorname{SU}_{\varepsilon}(t)=\operatorname{SU}_{\varepsilon}(q).

Conversely, let sSs\in S and assume that SUε(s)SUε(q)\operatorname{SU}_{\varepsilon}(s)\geq\operatorname{SU}_{\varepsilon}(q) for all qSS()q\in\operatorname{S}_{S}(\varnothing) and all ε>0\varepsilon>0. Let gGg\in G, gAsg\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}_{A}s. Then SUε(gs/g)=SUε(s/g)=SUε(s)SUε(gs)SUε(gs/g)\operatorname{SU}_{\varepsilon}(gs/g)=\operatorname{SU}_{\varepsilon}(s/g)=\operatorname{SU}_{\varepsilon}(s)\geq\operatorname{SU}_{\varepsilon}(gs)\geq\operatorname{SU}_{\varepsilon}(gs/g). Thus equality holds all the way for all ε>0\varepsilon>0, whereby gsggs\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\displaystyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\displaystyle\smile$\hss}\kern 5.71527pt}{\kern 5.71527pt\hbox to0.0pt{\hss$\textstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\textstyle\smile$\hss}\kern 5.71527pt}{\kern 4.53473pt\hbox to0.0pt{\hss$\scriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 2.71246pt\hbox to0.0pt{\hss$\scriptstyle\smile$\hss}\kern 4.53473pt}{\kern 3.95836pt\hbox to0.0pt{\hss$\scriptscriptstyle\mid$\hbox to0.0pt{$$\hss}\hss}\lower 1.93747pt\hbox to0.0pt{\hss$\scriptscriptstyle\smile$\hss}\kern 3.95836pt}}g, so ss is generic. ∎

References