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

Bi-invariant types, reliably invariant types, and the comb tree property

James E. Hanson [email protected]
Abstract.

We introduce and examine some special classes of invariant types—bi-invariant, strongly bi-invariant, extendibly invariant, and reliably invariant types—and show that they are related to certain model-theoretic tree properties.

We show that the comb tree property (recently introduced by Mutchnik) is equivalent to the failure of Kim’s lemma for bi-invariant types and is implied by the failure of Kim’s lemma for reliably invariant types over invariance bases. We show that every type over an invariance base extends to a reliably invariant type—generalizing an unpublished result of Kruckman and Ramsey—and use this to show that, under a reasonable definition of Kim-dividing, Kim-forking coincides with Kim-dividing over invariance bases in theories without the comb tree property. Assuming a measurable cardinal, we characterize the comb tree property in terms of a form of dual local character.

We also show that the antichain tree property (introduced by Ahn and Kim) seems to have a somewhat similar relationship to strong bi-invariance. In particular, we show that NATP theories satisfy Kim’s lemma for strongly bi-invariant types and (assuming a measurable cardinal) satisfy a different form of dual local character. Furthermore, we examine a mutual generalization of the local character properties satisfied by NTP2 and NSOP1 theories and show that it is satisfied by all NATP theories.

Finally, we give some related minor results—a strengthened local character characterization of NSOP1 and a characterization of coheirs in terms of invariant extensions in expansions—as well as a pathological example of Kim-dividing.

Key words and phrases:
invariant types, model-theoretic tree properties
2020 Mathematics Subject Classification:
03C45

Introduction

In neostability theory, it is often useful when a combinatorial tameness property is found to be characterized by some form of Kim’s lemma, which states that dividing along some indiscernible sequence in some class AA of indiscernible sequences entails dividing along all indiscernible sequences in some other class BB. Some standard examples are the following:

  • (TT simple) If φ(x,b)\varphi(x,b) divides over MM, then φ(x,b)\varphi(x,b) divides along any non-forking Morley sequence in tp(b/M)\operatorname{tp}(b/M).

  • (TT NTP2) If φ(x,b)\varphi(x,b) divides over MM, then φ(x,b)\varphi(x,b) divides along any strict Morley sequence over MM.

  • (TT NSOP1) If φ(x,b)\varphi(x,b) Kim-divides over MM, then φ(x,b)\varphi(x,b) divides along any sequence generated by an MM-invariant type extending tp(b/M)\operatorname{tp}(b/M) and along any tree Morley sequence in tp(b/M)\operatorname{tp}(b/M).

An important additional consideration is the existence of the relevant special indiscernible sequences in the second class, possibly with extra restrictions, such as being indiscernible relative to some set of parameters. The modern proof of the symmetry of non-forking in simple theories uses the following fact: (TT simple) If aMfba\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{f}}_{M}b, then there is a Morley sequence (ai)i<ω(a_{i})_{i<\omega} which is MbMb-indiscernible with a0=aa_{0}=a. Likewise, in the context of NSOP1 theories, symmetry of Kim-independence was shown by Kaplan and Ramsey using a similar but significantly harder to prove fact: (TT NSOP1) If aMKba\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{K}}_{M}b, then there is a tree Morley sequence (ai)i<ω(a_{i})_{i<\omega} that is MbMb-indiscernible with a0=aa_{0}=a [7]. For NTP2 theories, on the other hand, while symmetry is not expected, similar machinery was used by Chernikov and Kaplan to show that forking and dividing coincide over extension bases. In particular, as part of this argument, they showed that in NTP2 theories, any type over an invariance base extends to a strictly invariant type [5]. Kruckman and Ramsey observed that Chernikov and Kaplan’s proof almost doesn’t rely on the assumption of NTP2 and can be adapted to show that any type over a model extends to a Kim-strictly invariant type [10].

Definition 0.1.

An MM-invariant type p(x)p(x) is strictly invariant if whenever apMba\models p\operatorname{\upharpoonright}Mb, bMfab\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{f}}_{M}a. p(x)p(x) is Kim-strictly invariant if whenever apMba\models p\operatorname{\upharpoonright}Mb, bMKab\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{K}}_{M}a.

Fact 0.2 (Kruckman, Ramsey [10]).

(TT arbitrary) Any type over a model MM has a Kim-strictly invariant extension.

In 2.15, we generalize 0.2 by showing that any type over an invariance base extends to a Kim-strictly invariant type.

0.2 prompted Kruckman and Ramsey to investigate the following variant of Kim’s lemma as possibly characterizing a good mutual generalization of NTP2 and NSOP1: A theory TT satisfies new Kim’s lemma if whenever φ(x,b)\varphi(x,b) Kim-divides over a model MM, then it Kim-divides with regards to any Kim-strictly invariant type extending tp(b/M)\operatorname{tp}(b/M). As described in an upcoming note by Kruckman and Ramsey, a failure of new Kim’s lemma entails the existence of a certain combinatorial configuration mutually generalizing TP2 and SOP1, which they refer to tentatively as the bizarre tree property or BTP, but whether the converse holds is unclear at the moment [10].

The only other previously known general construction of (Kim-)strictly invariant types seems to have been the following fact:

Definition 0.3.

A global type p(x)p(x) is an MM-coheir or a coheir over MM if it is finitely satisfiable in MM. p(x)p(x) is an MM-heir or an heir over MM if for every MM-formula φ(x,y)\varphi(x,y), if there is a bb in the monster such that φ(x,b)p(x)\varphi(x,b)\in p(x), then there is a cMc\in M such that φ(x,c)M\varphi(x,c)\in M.

Fact 0.4.

If p(x)p(x) is MM-invariant and NMN\succeq M is (|M|+||)+(|M|+|\mathcal{L}|)^{+}-saturated, then pnp^{\otimes n} is an NN-heir for every n<ωn<\omega.

In particular, if p(x)p(x) is an MM-heir and apMba\models p\operatorname{\upharpoonright}Mb, then tp(b/Ma)\operatorname{tp}(b/Ma) extends to an MM-coheir, implying that any type that is both invariant and an heir is strictly invariant. But the property in 0.4 is ostensibly stronger than mere strict invariance.

Definition 0.5.

A global type p(x)p(x) is an MM-heir-coheir if p(x)p(x) is an MM-heir and an MM-coheir. p(x)p(x) is MM-bi-invariant if it is MM-invariant and whenever apMba\models p\operatorname{\upharpoonright}Mb, then tp(b/Ma)\operatorname{tp}(b/Ma) extends to a global MM-invariant type. p(x)p(x) is strongly MM-bi-invariant if pnp^{\otimes n} is MM-bi-invariant for every n<ωn<\omega.

Note that in an NIP theory, any strictly invariant type is bi-invariant.

One of the contributions of this paper will be to present a couple of novel methods for constructing heir-coheirs, and therefore bi-invariant types. Unlike 0.4, the heir-coheirs we construct will not be strongly bi-invariant. In fact, there seems to be a significant difference between the tasks of constructing bi-invariant and strongly bi-invariant types. In particular, bi-invariance is something that can be accomplished generically on the level of formulas. This can be seen in the following proposition (which we will not use elsewhere in this paper but is motivating and may be of independent interest).

Proposition 0.6.

Let TT be a countable theory. Let MTM\models T be a countable model with the following weak saturation property:

  • For any MM-formula φ(x,y)\varphi(x,y), if there is an element bb in the monster such that φ(M,b)\varphi(M,b) is infinite, then there is a cMc\in M such that φ(M,c)\varphi(M,c) is infinite.

Let S1nr(M)S_{1}^{\mathrm{nr}}(M) be the set of non-realized 11-types over MM. There is a dense GδG_{\delta} set XS1nr(M)X\subseteq S_{1}^{\mathrm{nr}}(M) such that for any pXp\in X, any MM-coheir extending pp is an MM-heir.

Proof.

For any MM-formulas φ(x)\varphi(x) and ψ(x,y)\psi(x,y), if there is a bb in the monster such that φ(M)ψ(M,b)\varphi(M)\wedge\psi(M,b) is infinite, find a cMc\in M such that φ(M)ψ(M,c)\varphi(M)\wedge\psi(M,c) is infinite and let Uφ,ψU_{\varphi,\psi} be the set of types in S1nr(M)S^{\mathrm{nr}}_{1}(M) containing φ(x)ψ(x,c)\varphi(x)\wedge\psi(x,c). (Note that this set is non-empty since φ(M)ψ(M,c)\varphi(M)\wedge\psi(M,c) is infinite.) If there is no such cc, let Uφ,ψU_{\varphi,\psi} be the set of types in S1nr(M)S^{\mathrm{nr}}_{1}(M) containing φ(x)\varphi(x). Let Uψ=φ(M)Uφ,ψU_{\psi}=\bigcup_{\varphi\in\mathcal{L}(M)}U_{\varphi,\psi}. UψU_{\psi} is a dense open subset of S1nr(M)S^{\mathrm{nr}}_{1}(M) for each ψ(x,y)\psi(x,y). Let X=ψ(M)UψX=\bigcap_{\psi\in\mathcal{L}(M)}U_{\psi}. XX is a dense GδG_{\delta} set.

Now fix p(x)Xp(x)\in X. We need to argue that any MM-coheir q(x)p(x)q(x)\supset p(x) is an MM-heir. Suppose that ψ(x,b)q(x)\psi(x,b)\in q(x). Since q(x)q(x) is an MM-coheir that is not realized in MM, φ(M)ψ(M,b)\varphi(M)\wedge\psi(M,b) must be infinite for every MM-formula φ(x)p(x)\varphi(x)\in p(x). Since p(x)Uψp(x)\in U_{\psi}, we have that for some cMc\in M, ψ(x,c)p(x)\psi(x,c)\in p(x). Since we can do this for any MM-formula ψ(x,y)\psi(x,y), we have that q(x)q(x) is an MM-heir. ∎

Note that the saturation property in 0.6 is satisfied by any computably saturated model and by any model of a theory that eliminates \exists^{\infty}. It’s also relatively easy to see that 0.6 can fail in models at least as large as the covering number of the ideal of meager sets. Specifically, no coheir of a type over (,<)(\mathbb{R},<) is also an heir.

In [11], Mutchnik introduced a certain combinatorial configuration he called ω\omega-DCTP2 as part of his proof that NSOP1 and NSOP2 (or equivalently NTP1) are the same. For the sake of simplicity, we will just refer to this condition as the comb tree property or CTP (1.1). CTP will be a major focus of this paper.

In [2], Ahn and Kim introduced the antichain tree property or ATP, which, like BTP, is intended to be a good mutual generalization of TP2 and SOP1. In [3], they show that ATP is always witnessed by a formula with a single free variable and is always witnessed by the 22-inconsistent version of the condition. Furthermore, in [9], Kim and Lee show that in NATP theories, Kim-forking and Kim-dividing coincide over models and that if new Kim’s lemma holds (in the sense of Kruckman and Ramsey), then coheirs are universal witnesses of Kim-dividing if and only if they are Kim-strict.

It is not hard to show111The fact that ATP implies CTP is immediate from the definition, as discussed in [3], and the fact that CTP implies BTP is a corollary of unpublished results of Kruckman and Ramsey or separately of some of our results here. Specifically, since bi-invariant types are strictly invariant, 1.8 implies that any CTP theory fails the new Kim’s lemma given in [10], whose failure implies BTP. that these three conditions are related. In particular (and in an alphabetically frustrating way), ATP implies CTP, which implies BTP.

The main contribution of this paper will be a characterization of NCTP in terms of a variant of new Kim’s lemma (where Kim-strict invariance is replaced with bi-invariance). We will also give an argument that NATP implies the analogous property with strong bi-invariance, although the converse is unclear.

In Section 2, we will argue that a certain definition of Kim-dividing is natural over invariance bases and extend a result of Mutchnik’s [11, Thm. 4.9] and a result of Kim and Lee’s [9, Rem. 5.10] by showing that under this definition in NCTP theories, Kim-forking coincides with Kim-dividing over invariance bases. We do this by introducing the notions of reliably invariant types and reliable coheirs and show that these always exist over invariance bases and models, respectively, and that these always witness Kim-dividing in CTP theories. This partially remedies a deficiency of our characterization which is that not all types over models extend to heir-coheirs or even strictly invariant types (see [5, Sec. 5.1]). A remaining issue with this is that it is unclear whether NCTP is actually characterized by Kim’s lemma for reliable coheirs.

In Section 3, we give a dual local character characterization of NCTP assuming the existence of a measurable cardinal. We show that NATP theories satisfy a similar form of dual local character, but again the converse is unclear. In Section 3.2, we also discuss a notion of local character mutually generalizing the local character properties that characterize NTP2 and NSOP1 theories and show that it is implied by NATP.

Finally, we would like to thank Alex Kruckman and Nicholas Ramsey for many valuable discussions regarding the ideas in this paper.

1. The comb tree property

Given σ2<ω\sigma\in 2^{<\omega}, we write σ{\mathopen{\llbracket}\sigma\mathclose{\rrbracket}} for the set of τ2<ω\tau\in 2^{<\omega} such that τσ\tau\trianglerighteq\sigma.

Definition 1.1.

Given an ordinal α\alpha, a set X2<αX\subseteq 2^{<\alpha} is a right-comb222In the existing literature, sets satisfying (the mirror image of) this condition are referred to as ‘descending combs,’ but this use of the term relies on an implicit assumption that the comb is enumerated in lexicographic order, which is not normally part of the definition of being a descending comb in a tree. if it is an antichain and satisfies that for any σ2<α\sigma\in 2^{<\alpha}, if there a τX\tau\in X with τσ1\tau\trianglerighteq\sigma\frown 1, then there is at most one τX\tau\in X with τσ0\tau\trianglerighteq\sigma\frown 0.

A theory TT has kk-CTP if there is a binary tree {bσ}σ2<ω\{b_{\sigma}\}_{\sigma\in 2^{<\omega}} and a formula φ(x,y)\varphi(x,y) such that for any path α2ω\alpha\in 2^{\omega}, {φ(x,bαn):n<ω}\{\varphi(x,b_{\alpha\operatorname{\upharpoonright}n}):n<\omega\} is kk-inconsistent but for any right-comb X2<ωX\subseteq 2^{<\omega}, {φ(x,bσ):σX}\{\varphi(x,b_{\sigma}):\sigma\in X\} is consistent.

TT has CTP if it has kk-CTP for some k<ωk<\omega.

Note that since any right-comb is an antichain, any theory with ATP has CTP, as observed in [9, Rem. 5.7]. Another thing to note is that the dual condition of CTP (where paths are consistent and right-combs are kk-inconsistent), was shown to be equivalent to NSOP1 in [11]. This means that many of our results dualize to give analogous results for NSOP1 theories. We collect these in Appendix A.

1.1. Failure of Kim’s lemma for heir-coheirs from the comb tree property

Definition 1.2.

A subset X2<ωX\subseteq 2^{<\omega} is dense above σ\sigma if for every τσ\tau\trianglerighteq\sigma, XτX\cap{\mathopen{\llbracket}\tau\mathclose{\rrbracket}} is non-empty. XX is somewhere dense if there is a σ\sigma such that XX is dense above σ\sigma. A filter \mathcal{F} on 2<ω2^{<\omega} is everywhere somewhere dense if every XX\in\mathcal{F} is somewhere dense.

In this paper, we will only ever use the term ‘somewhere dense’ to refer to the above property of a subset of a tree. We will never use it in the topological sense. We will also only use the term ‘filter’ to refer to proper filters (i.e., filters that do not contain \varnothing).

The following is a fairly standard idea in forcing.

Lemma 1.3.

For any somewhere dense X2<ωX\subseteq 2^{<\omega} and any Y2<ωY\subseteq 2^{<\omega}, either XYX\cap Y is somewhere dense or XYX\setminus Y is somewhere dense.

Proof.

Fix σ\sigma such that XX is dense over σ\sigma. Suppose that for every τσ\tau\trianglerighteq\sigma, XYX\cap Y fails to be dense above τ\tau. Then for every τσ\tau\trianglerighteq\sigma, (XY)τ(X\setminus Y)\cap{\mathopen{\llbracket}\tau\mathclose{\rrbracket}} is non-empty, so XYX\setminus Y is dense above σ\sigma. ∎

Lemma 1.4.

Every everywhere somewhere dense filter extends to some everywhere somewhere dense ultrafilter.

Proof.

By transfinite induction it is sufficient to show that if \mathcal{F} is an everywhere somewhere dense filter and Y2<ωY\subseteq 2^{<\omega}, then either {Y}\mathcal{F}\cup\{Y\} or {2<ωY}\mathcal{F}\cup\{2^{<\omega}\setminus Y\} generates an everywhere somewhere dense filter. So fix some such YY. By 1.3 we have that for each XX\in\mathcal{F}, either XYX\cap Y is somewhere dense or XYX\setminus Y is somewhere dense. One of the sets {X:XYsomewhere dense}\{X\in\mathcal{F}:X\cap Y~{}\text{somewhere dense}\} and {X:XYsomewhere dense}\{X\in\mathcal{F}:X\setminus Y~{}\text{somewhere dense}\} must be cofinal in \mathcal{F}. Assume without loss of generality that {X:XYsomewhere dense}\{X\in\mathcal{F}:X\cap Y~{}\text{somewhere dense}\} is cofinal in \mathcal{F}. This implies that actually XYX\cap Y is somewhere dense for all XX\in\mathcal{F}. Therefore {Y}\mathcal{F}\cup\{Y\} generates an everywhere somewhere dense filter. ∎

We will see later in 3.1 that the following 1.5 holds even for uncountable languages. That said, the proof in uncountable languages is a bit more technical, so we feel it is appropriate to present the friendlier proof for countable theories first.

Proposition 1.5.

Let TT be a countable theory. If TT has CTP, then there is a countable model MM, a formula φ(x,y)\varphi(x,y), an MM-heir-coheir p(y)p(y), and an MM-coheir q(y)q(y) such that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide.

Proof.

Let φ(x,y)\varphi(x,y) and (bσ)σ2<ω(b_{\sigma})_{\sigma\in 2^{<\omega}} witness that TT has CTP.

Let (k(n),m(n))n<ω(k(n),m(n))_{n<\omega} be an enumeration of ω2\omega^{2} with the property that for each pair ,o<ω\ell,o<\omega, the set {n<ω:k(n),m(n)=,o}\{n<\omega:\langle k(n),m(n)\rangle=\langle\ell,o\rangle\} is infinite.

Fix a countable model M0{bσ:σ2<ω}M_{0}\supseteq\{b_{\sigma}:\sigma\in 2^{<\omega}\}. Let σ0=\sigma_{0}=\varnothing and let X0=1X_{0}={\mathopen{\llbracket}1\mathclose{\rrbracket}}. Note that X0X_{0} is dense over σ01\sigma_{0}\frown 1 and X0σ01X_{0}\subseteq{\mathopen{\llbracket}\sigma_{0}\frown 1\mathclose{\rrbracket}}. (This will be our induction hypothesis.)

At stage n<ωn<\omega, suppose we have a countable model MnM_{n} and σn\sigma_{n} and XnX_{n} such that XnX_{n} is dense over σn1\sigma_{n}\frown 1 and Xnσn1X_{n}\subseteq{\mathopen{\llbracket}\sigma_{n}\frown 1\mathclose{\rrbracket}}. Let (ψk()n(y,z))<ω(\psi_{k(\ell)}^{n}(y,z))_{\ell<\omega} be an enumeration of all MnM_{n}-formulas. To get Mn+1M_{n+1}, σn+1\sigma_{n+1}, and Xn+1X_{n+1}, perform the following construction:

  • If ψk(n)m(n)(y,z)\psi^{m(n)}_{k(n)}(y,z) has already been defined and there is a cc in the monster such that Xn{σ:ψk(n)m(n)(bσ,c)}X_{n}\cap\{\sigma:\psi^{m(n)}_{k(n)}(b_{\sigma},c)\} is somewhere dense: Fix some such cc. Let Mn+1MncM_{n+1}\supseteq M_{n}c be a countable model. Find τσn1\tau\trianglerighteq\sigma_{n}\frown 1 such that Xn+12Xn{σ:ψk(n)m(n)(bσ,c)}X_{n+\frac{1}{2}}\coloneqq X_{n}\cap\{\sigma:\psi^{m(n)}_{k(n)}(b_{\sigma},c)\} is dense over τ\tau. Find σn+1τ\sigma_{n+1}\trianglerighteq\tau such that bσn+1Xn+12b_{\sigma_{n+1}}\in X_{n+\frac{1}{2}}, and let Xn+1=Xn+12σn+11X_{n+1}=X_{n+\frac{1}{2}}\cap{\mathopen{\llbracket}\sigma_{n+1}\frown 1\mathclose{\rrbracket}}.

  • If the condition in the previous bullet point fails: Let Mn+1=MnM_{n+1}=M_{n}. Find σn+1σn1\sigma_{n+1}\trianglerighteq\sigma_{n}\frown 1 such that bσn+1Xnb_{\sigma_{n+1}}\in X_{n}. Let Xn+1=Xnσn+11X_{n+1}=X_{n}\cap{\mathopen{\llbracket}\sigma_{n+1}\frown 1\mathclose{\rrbracket}}.

Note that in both cases we have ensured that bσn+1Xnb_{\sigma_{n+1}}\in X_{n}.

After the construction is completed, let M=n<ωMnM=\bigcup_{n<\omega}M_{n}. Let \mathcal{F} be the filter generated by {Xn:n<ω}{Y}\{X_{n}:n<\omega\}\cup\left\{Y\right\} where Yn<ωσn0Y\coloneqq\bigcup_{n<\omega}{\mathopen{\llbracket}\sigma_{n}\frown 0\mathclose{\rrbracket}}. Note that since each XnX_{n} is dense above σn+1\sigma_{n+1}, this filter is everywhere somewhere dense. Let 𝒰\mathcal{U} be an everywhere somewhere dense ultrafilter extending \mathcal{F}. Let p(y)p(y) be the global MM-coheir corresponding to 𝒰\mathcal{U} (i.e., p(y)={ψ(y,c):{σ2<ω:ψ(bσ,c)}𝒰}p(y)=\{\psi(y,c):\{\sigma\in 2^{<\omega}:\psi(b_{\sigma},c)\}\in\mathcal{U}\}). Let q(y)q(y) be any non-realized global MM-coheir finitely satisfiable in {bσn:n<ω}\{b_{\sigma_{n}}:n<\omega\}. Note that {φ(x,bσn):n<ω}\{\varphi(x,b_{\sigma_{n}}):n<\omega\} is uniformly inconsistent (since the σn\sigma_{n}’s form a path in 2<ω2^{<\omega}). Therefore φ(x,y)\varphi(x,y) must qq-Kim-divide.

Claim. pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M.

Proof of claim. Fix an MM-formula ψ(y)p(y)\psi(y)\in p(y). ψ(y)\psi(y) is actually an MkM_{k}-formula for some k<ωk<\omega. By the choice of our enumeration, this means that there was a stage nn at which ψk(n)m(n)(y,z)\psi^{m(n)}_{k(n)}(y,z) was defined and equal to ψ(y)\psi(y) (where zz is a dummy variable). Since ψ(y)p(y)\psi(y)\in p(y) and since 𝒰\mathcal{U} is everywhere somewhere dense, we must have chosen the first bullet point at this stage. Hence ψ(bσ)\psi(b_{\sigma}) holds for all σXn+1\sigma\in X_{n+1}, whereby ψ(bσ)\psi(b_{\sigma_{\ell}}) holds for all >n+1\ell>n+1. Therefore ψ(y)q(y)\psi(y)\in q(y) as well. Since we can do this for every MM-formula ψ(y)p(y)\psi(y)\in p(y), we have that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M. \blacksquare

Claim. pp is an heir over MM.

Proof of claim. Fix an MM-formula ψ(y,z)\psi(y,z). Once again, there must have been a stage nn at which ψk(n)m(n)\psi^{m(n)}_{k(n)} was defined and equal to ψ\psi. Suppose that there is a dd in the monster such that ψ(y,d)p(y)\psi(y,d)\in p(y). Since 𝒰\mathcal{U} is everywhere somewhere dense, this implies that we chose the first bullet point at stage nn, so we found some cc, added this cc to Mn+1M_{n+1}, and moved to a set Xn+1X_{n+1} satisfying ψ(bσ,c)\psi(b_{\sigma},c) for all σXn+1\sigma\in X_{n+1}. Therefore ψ(y,c)p(y)\psi(y,c)\in p(y). Since we can do this for any MM-formula ψ(y,z)\psi(y,z), we have that pp is an heir over MM. \blacksquare

Claim. φ(x,y)\varphi(x,y) does not pp-Kim-divide.

Proof of claim. Recall that Yn<ωσn0Y\coloneqq\bigcup_{n<\omega}{\mathopen{\llbracket}\sigma_{n}\frown 0\mathclose{\rrbracket}} is in 𝒰\mathcal{U}. Also note that by construction, σn𝒰{\mathopen{\llbracket}\sigma_{n}\mathclose{\rrbracket}}\in\mathcal{U} for each n<ωn<\omega. Let Yn=YσnY_{n}=Y\cap{\mathopen{\llbracket}\sigma_{n}\mathclose{\rrbracket}} for each n<ωn<\omega. Note that if n0,n1,,nk1n_{0},n_{1},\dots,n_{k-1} is an increasing sequence of integers and if τiσni0\tau_{i}\trianglerighteq\sigma_{n_{i}}\frown 0 for each i<ki<k, then {τi:i<k}\{\tau_{i}:i<k\} is a right-comb.

Let (ek)k<ω(e_{k})_{k<\omega} be a Morley sequence generated by pp. We have by assumption that for any finite right-comb ZYZ\subseteq Y, {φ(x,bτ):τZ}\{\varphi(x,b_{\tau}):\tau\in Z\} is consistent. Suppose that for some k<ωk<\omega (possibly 0), we’ve shown that for every finite right-comb ZYZ\subseteq Y, {φ(x,bτ):τZ}{φ(x,ei):i<k}\{\varphi(x,b_{\tau}):\tau\in Z\}\cup\{\varphi(x,e_{i}):i<k\} is consistent. For any such ZZ, there is an nn such that for any ηYn\eta\in Y_{n}, Z{η}Z\cup\{\eta\} is a right-comb. (In particular, this will be true for any sufficiently large nn.) This implies that {φ(x,bτ):τZ}{φ(x,ei):i<k+1}\{\varphi(x,b_{\tau}):\tau\in Z\}\cup\{\varphi(x,e_{i}):i<k+1\} is consistent. Hence, by induction, we have that {φ(x,ei):i<ω}\{\varphi(x,e_{i}):i<\omega\} is consistent and so φ(x,y)\varphi(x,y) does not pp-Kim-divide. \blacksquare

Therefore pp and qq satisfy the required conditions. ∎

One thing to note is that instead of building the model in the proof of 1.5 at the same time as the filter, we could instead have built a fixed model satisfying the following weak saturation property that is analogous to the one found in 0.6:

  • For every MM-formula φ(x,y)\varphi(x,y) and σ2<ω\sigma\in 2^{<\omega}, if there is a cc in the monster such that {τ2<ω:φ(bτ,c)}\{\tau\in 2^{<\omega}:\varphi(b_{\tau},c)\} is dense above σ\sigma, then there is a dMd\in M such that {τ2<ω:φ(bτ,d)}\{\tau\in 2^{<\omega}:\varphi(b_{\tau},d)\} is dense above σ\sigma.

This ends up being a bit more work to state than the given proof of 1.5, but this perspective highlights the similarity between 1.5 and 0.6.

Something that is frustrating and interesting is that, at least to the present author, there doesn’t seem to be a clear way to prove the ATP analog of 1.5. That is to say, it is unclear if one can use an instance of ATP to build a failure of Kim’s lemma for strongly bi-invariant types.

Question 1.6.

If TT has ATP, does it follow that there is a model MM, a formula φ(x,y)\varphi(x,y), a strongly MM-bi-invariant type p(y)p(y), and an MM-invariant type q(y)q(y) such that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide?

1.2. Characterization of the comb tree property

Proposition 1.7.

Suppose there is a formula φ(x,y)\varphi(x,y) and two AA-invariant types p(y)p(y) and q(y)q(y) such that pA=qAp\operatorname{\upharpoonright}A=q\operatorname{\upharpoonright}A and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide.

  1. (1)

    If pp is bi-invariant, then TT has CTP.

  2. (2)

    If pp is strongly bi-invariant, then TT has ATP.

Proof.

For 1, suppose that {φ(x,ei):i<ω}\{\varphi(x,e_{i}):i<\omega\} is kk-inconsistent for any Morley sequence e<ωe_{<\omega} generated by qq.

We will argue by induction that for any n<ωn<\omega, there is a family (bσ)σ2n(b_{\sigma})_{\sigma\in 2^{\leq n}} of parameters realizing pAp\operatorname{\upharpoonright}A such that for each σ2<n\sigma\in 2^{<n},

  • for each τσ0\tau\trianglerighteq\sigma\frown 0, bτpA{bη:ησ1}b_{\tau}\models p\operatorname{\upharpoonright}A\cup\{b_{\eta}:\eta\trianglerighteq\sigma\frown 1\} and

  • bσqA{bη:ησ}b_{\sigma}\models q\operatorname{\upharpoonright}A\cup\{b_{\eta}:\eta\triangleright\sigma\}.

Note that the second condition clearly implies that any path through such a tree is a (reverse) Morley sequence generated by qq (so in particular, {φ(x,bηi):in}\{\varphi(x,b_{\eta\operatorname{\upharpoonright}i}):i\leq n\} is kk-inconsistent for any η2n\eta\in 2^{n}). The first condition, moreover, implies that any right-comb in the tree is a Morley sequence generated by pp (in some enumeration). Therefore for any right-comb X2nX\subseteq 2^{\leq n}, {φ(x,bσ):σX}\{\varphi(x,b_{\sigma}):\sigma\in X\} is consistent. So if we can show that these trees exist for all nn, we will have established that TT has kk-CTP.

For n=1n=1, the condition is trivial, since there is only a single σ\sigma in the tree.

Suppose we have built such a tree (cσ)σ2n(c_{\sigma})_{\sigma\in 2^{\leq n}} for some nn. Start the next tree by setting b0σ=cσb_{0\frown\sigma}=c_{\sigma} for each σ2n\sigma\in 2^{\leq n}. Find dpA{bσ:σ2n+1,σ0}d\models p\operatorname{\upharpoonright}A\cup\{b_{\sigma}:\sigma\in 2^{\leq n+1},~{}\sigma\trianglerighteq 0\}. Since pp is AA-bi-invariant, we can find an AA-invariant type r(x¯)r(\bar{x}) extending tp(b0/Ad)\operatorname{tp}(b_{\trianglerighteq 0}/Ad). Now find b1b_{\trianglerighteq 1} such that b1rAb0b_{\trianglerighteq 1}\models r\operatorname{\upharpoonright}Ab_{\trianglerighteq 0}. By AA-invariance of rr, we have that for each σ0\sigma\trianglerighteq 0, bσpAb1b_{\sigma}\models p\operatorname{\upharpoonright}Ab_{\trianglerighteq 1}. Finally, let bqAbb_{\varnothing}\models q\operatorname{\upharpoonright}Ab_{\triangleright\varnothing}.

Since we can do this for any nn, by induction, we have that TT has kk-CTP.

For 2, then we can build similar trees (bσ)σ2n(b_{\sigma})_{\sigma\in 2^{\leq n}} satisfying the additional property that

  • for each antichain BB of elements of 2n2^{\leq n}, {bσ:σB}\{b_{\sigma}:\sigma\in B\} is a Morley sequence in pp over A{bη:ησ1}A\cup\{b_{\eta}:\eta\trianglerighteq\sigma\frown 1\}.

To see that this is possible, we just need to modify the induction step. Suppose we have built such a tree (cσ)σ2n(c_{\sigma})_{\sigma\in 2^{\leq n}} for some nn. Start the next tree by setting b0σ=cσb_{0\frown\sigma}=c_{\sigma} for each σ2n\sigma\in 2^{\leq n}. Find d¯pωA{bσ:σ2n+1,σ0}\bar{d}\models p^{\otimes\omega}\operatorname{\upharpoonright}A\cup\{b_{\sigma}:\sigma\in 2^{\leq n+1},~{}\sigma\trianglerighteq 0\}. Since pωp^{\otimes\omega} is AA-bi-invariant, we can find an AA-invariant type r(x¯)r(\bar{x}) extending tp(b0/Ad¯)\operatorname{tp}(b_{\trianglerighteq 0}/A\bar{d}). Then we can find b1b_{\trianglerighteq 1} such that b1rAb0b_{\trianglerighteq 1}\models r\operatorname{\upharpoonright}Ab_{\trianglerighteq 0}. Now, let BB be an antichain of elements of 2n+12^{\leq n+1}. If B={}B=\{\varnothing\}, then the statement is trivial. Otherwise, we have that B0B\cap{\mathopen{\llbracket}0\mathclose{\rrbracket}} and B1B\cap{\mathopen{\llbracket}1\mathclose{\rrbracket}} are each antichains. By the induction hypothesis, this means that b¯0={bσ:σB0}\bar{b}^{0}=\{b_{\sigma}:\sigma\in B\cap{\mathopen{\llbracket}0\mathclose{\rrbracket}}\} is a Morley sequence in pp and b¯1={bσ:σB1}\bar{b}^{1}=\{b_{\sigma}:\sigma\in B\cap{\mathopen{\llbracket}1\mathclose{\rrbracket}}\} is a Morley sequence in pp. By construction, this means that b¯0\bar{b}^{0} realizes the same type over AA as some initial segment of d¯\bar{d}. Therefore b¯0p|B0|Ab¯1\bar{b}^{0}\models p^{\otimes|B\cap{\mathopen{\llbracket}0\mathclose{\rrbracket}}|}\operatorname{\upharpoonright}A\bar{b}^{1}. Since b¯1\bar{b}^{1} is a Morley sequence in pp over AA, this implies that b¯0b¯1\bar{b}^{0}\bar{b}^{1} is a Morley sequence in pp over AA.

Finally, since we can do this for any nn, we have that TT has kk-ATP. ∎

Theorem 1.8.

A theory TT has CTP if and only if there is a model MTM\models T, a formula φ(x,y)\varphi(x,y), an MM-heir-coheir p(y)p(y), and an MM-coheir q(y)q(y) such that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide.

Proof.

By 1.5 (or 3.1 if \mathcal{L} is uncountable), we have that if TT has CTP, then the required configuration exists. The converse follows by 1.7. ∎

One issue which we have been ignoring up until now is whether kk-CTP implies \ell-CTP for <k\ell<k. This is an important structural property of SOP1 and ATP. The proofs going into the proof of 1.8 clearly preserve the relevant degree of inconsistency. We have been unable to resolve this question, and likewise we have been unable to show that CTP is always witnessed by a formula in a single free variable. Despite the similarity between CTP and ATP, the proofs of these facts for ATP in [3] seem to rely pretty heavily on nice structural properties of antichains that right-combs do no share (e.g., the fact that an ‘antichain of antichains’ is an antichain, which is used in [3, Lem. 3.20]).

Question 1.9.

If TT has CTP, does it follow that TT has 22-CTP?

Question 1.10.

If TT has CTP, does it have CTP witnessed by a formula with a single free variable?

One thing to note is that the proofs of these kinds of facts often make good use of indiscernible trees. As observed in [9, Rem. 5.6], CTP is always witnessed by a strongly indiscernible tree. While this will almost certainly be an important tool for studying NCTP theories at some point, we nevertheless find it interesting that tree indiscernibility plays no role in any of the proofs in this paper.

2. Reliably invariant types

2.1. What should Kim-dividing over invariance bases be?

In [7], Kaplan and Ramsey originally defined Kim-dividing in terms of dividing along sequences generated by arbitrary invariant types. Pretty quickly in their analysis, however, it becomes clear that the natural concept is dividing along sequences generated by coheirs. Ramsey has in conversation consistently expressed the opinion that the definition of Kim-dividing in terms of coheirs is more natural. Coheirs satisfy two important properties that arbitrary invariant types do not:

  • (Expansion) If pp is an MM-coheir and MM^{\dagger} is an expansion of MM, then there is an MM^{\dagger}-coheir qq extending pp.

  • (Left extension) If p(x)p(x) is an MM-coheir and q(x,y)q(x,y) is a type over MM extending pMp\operatorname{\upharpoonright}M, then there is an MM-coheir r(x,y)r(x,y) extending p(x)q(x,y)p(x)\cup q(x,y).

In the context of NSOP1 theories (over models), the distinction between these two definitions becomes immaterial given the relevant Kim’s lemma. Dividing along some invariant Morley sequence implies dividing along all invariant Morley sequences. In SOP1 theories, however, a reasonable question, frequently asked by Kruckman, is whether the original definition of Kim-dividing is formula independent, i.e., if φ(x,b)\varphi(x,b) and ψ(x,c)\psi(x,c) are logically equivalent and φ(x,b)\varphi(x,b) Kim-divides over a model MM, then ψ(x,c)\psi(x,c) Kim-divides over MM. It follows from [9, Thm. 3.10] that in an NATP theory, this is always the case. In Appendix C, we give an example showing that this can fail for ATP theories.

When attempting to define a robust notion of Kim-dividing over invariance bases, one ideally would like to retain the two nice properties of coheirs mentioned above. The issue with expansion, however, is that it characterizes coheirs over invariance bases.

Proposition 2.1.

Fix a set of parameters AA.

  1. (1)

    If AA is an invariance base,333Note that we are only considering invariance with regards to ordinary type, rather than Kim-Pillay or Lascar strong type, although over a set that is an invariance base in this sense, these notions collapse by essentially the same argument as in the proof of 1. then acl(A)=dcl(A)\operatorname{acl}(A)=\operatorname{dcl}(A).

  2. (2)

    Assume that A=acl(A)A=\operatorname{acl}(A). Fix an AA-invariant type p(x)p(x). The following are equivalent.

    1. (a)

      For any model MAM\supseteq A and any expansion MM^{\dagger} of MM, pMp\operatorname{\upharpoonright}M has a completion in S(M)S(M^{\dagger}) that is Aut(M/A)\operatorname{Aut}(M^{\dagger}/A)-invariant.

    2. (b)

      p(x)p(x) is finitely satisfiable in AA.

Proof.

For 1, suppose that AA is an invariance base. Fix an algebraic type p(x)S(A)p(x)\in S(A). Since AA is an invariance base, p(x)p(x) has an AA-invariant global extension q(x)q(x). Let a0,,an1a_{0},\dots,a_{n-1} be the realizations of pp. Let bb be a realization of qacl(A)q\operatorname{\upharpoonright}\operatorname{acl}(A). It must be the case that b=aib=a_{i} for some i<ni<n, but by invariance, this implies that n=1n=1.

2b \Rightarrow 2a. If p(x)p(x) is finitely satisfiable in AA, then there is an ultrafilter 𝒰\mathcal{U} on AA whose average type is p(x)p(x). The average type of 𝒰\mathcal{U} will be an invariant type extending p(x)p(x) in any expansion of the theory.

¬\neg2b \Rightarrow ¬\neg2a. Our proof of this is somewhat technical, so we have opted to put it in Appendix B. ∎

In particular, if AA is an invariance base, then every type over AA has an AA-invariant extension satisfying 2a if and only dcl(A)\operatorname{dcl}(A) is a model.

Fortunately, though, expansion seems to be more of a convenience than a necessity. The second property, however, seems to be fairly significant. All of this might suggest focusing on the following special class of invariant types.

Definition 2.2.

A global type p(x)p(x) is extendibly AA-invariant if for any q(x,y)S(A)q(x,y)\in S(A) extending pAp\operatorname{\upharpoonright}A, p(x)q(x,y)p(x)\cup q(x,y) extends to an AA-invariant type.

As we already said, coheirs over models are always extendibly invariant. The example given in Appendix C shows that not all invariant types over models are extendibly invariant.

While extendibly invariant types might seem relatively special, we get them for free over invariance bases.

Proposition 2.3.

Let AA be a set of parameters.

  1. (1)

    If there is an extendibly AA-invariant type, then AA is an invariance base.

  2. (2)

    An AA-invariant type p(x)p(x) is extendibly AA-invariant if and only if for every formula φ(x,b)p(x)\varphi(x,b)\in p(x) and every AA-formula ψ(x,y)\psi(x,y), if φ(x,b)ψ(x,y)\varphi(x,b)\wedge\psi(x,y) quasi-forks444Recall that a formula χ(x,c)\chi(x,c) quasi-forks over AA if there is no AA-invariant type containing χ(x,c)\chi(x,c). This is equivalent to implying a disjunction of formulas that quasi-divide over AA. over AA, then p(x)¬yψ(x,y)p(x)\vdash\neg\exists y\psi(x,y).

  3. (3)

    If p(x,z)p(x,z) is extendibly AA-invariant and r(x)r(x) is the restriction of p(x,z)p(x,z) to xx, then r(x)r(x) is extendibly AA-invariant.

  4. (4)

    If p(x¯)p(\bar{x}) is the type of an (|A|+|T|)+(|A|+|T|)^{+}-saturated model, then any AA-invariant extension of p(x¯)p(\bar{x}) is extendibly AA-invariant.

  5. (5)

    If AA is an invariance base, then every type over AA extends to an extendibly AA-invariant type.

  6. (6)

    If p(x)p(x) is an extendibly AA-invariant type, then for any q(x,y)S(A)q(x,y)\in S(A) extending pAp\operatorname{\upharpoonright}A, p(x)q(x,y)p(x)\cup q(x,y) extends to an extendibly AA-invariant type.

Proof.

1 is immediate. For 2, let p(x)p(x) be an AA-invariant type. Suppose that there is some formula φ(x,b)p(x)\varphi(x,b)\in p(x) and some AA-formula ψ(x,y)\psi(x,y) such that φ(x,b)ψ(x,y)\varphi(x,b)\wedge\psi(x,y) quasi-forks over AA and p(x)yψ(x,y)p(x)\vdash\exists y\psi(x,y). Let q(x,y)q(x,y) be a complete type extending pA{ψ(x,y)}p\operatorname{\upharpoonright}A\cup\{\psi(x,y)\}. We now have that p(x)q(x,y)p(x)\cup q(x,y) has no AA-invariant global extension, so p(x)p(x) is not extendibly AA-invariant.

Now assume that p(x)p(x) satisfies the condition in 2. Fix q(x,y)S(A)q(x,y)\in S(A) extending pAp\operatorname{\upharpoonright}A. We have that for every φ(x,b)p(x)\varphi(x,b)\in p(x) and every ψ(x,y)q(x,y)\psi(x,y)\in q(x,y), φ(x,b)ψ(x,y)\varphi(x,b)\wedge\psi(x,y) does not quasi-fork over AA. By compactness, this implies that there is an AA-invariant type r(x,y)r(x,y) containing each of these formulas, which implies that r(x,y)p(x)q(x,y)r(x,y)\supseteq p(x)\cup q(x,y).

For 3, let q(x,y)q(x,y) be a type over AA extending rAr\operatorname{\upharpoonright}A. (pA)(x,z)q(x,y)(p\operatorname{\upharpoonright}A)(x,z)\cup q(x,y) is consistent. Let s(x,y,z)s(x,y,z) be a completion of (pA)(x,z)q(x,y)(p\operatorname{\upharpoonright}A)(x,z)\cup q(x,y). We now have that there is an AA-invariant type t(x,y,z)t(x,y,z) extending p(x,z)s(x,y,z)p(x,z)\cup s(x,y,z). The restriction of tt to xzxz is now the required type.

For 4, let r(x¯)r(\bar{x}) be an AA-invariant extension of p(x¯)p(\bar{x}). Note that 2 implies that it is sufficient to check that every restriction of r(x¯)r(\bar{x}) to finitely many variables is extendibly AA-invariant. Let s(x)s(x) be a restriction to finitely many variables. Let q(x,y)q(x,y) be some type over AA extending sAs\operatorname{\upharpoonright}A. Let MM be a realization of p(x¯)p(\bar{x}) and let bb be the finite tuple of elements corresponding to the variable xx. By saturation, there is some cMc\in M such that bcq(x,y)bc\models q(x,y). Therefore, the restriction of p(x¯)p(\bar{x}) to the variables corresponding to bcbc witnesses that s(x)q(x,y)s(x)\cup q(x,y) has an AA-invariant extension. Since we can do this for any finite tuple of variables, we have that r(x¯)r(\bar{x}) is extendibly AA-invariant by 2.

5 and 6 now follow immediately from 3 and 4. ∎

Given 2.3, we propose that defining Kim-dividing over invariance bases in terms of extendibly invariant types is likely to be a more robust notion than Kim-dividing defined in terms of arbitrary invariant types. Our strongest evidence that this is a good definition (at least in the context of NCTP theories) is the results of the next section, specifically 2.6. As we discuss at the end of Section 2.2, however, it unclear that this evidence is completely solid. Nevertheless, we find 2.16 fairly compelling.

2.2. Kim’s lemma for reliably invariant types

Definition 2.4.

A sequence (bi)i<n(b_{i})_{i<n} is an invariant sequence over AA if biAbjb_{i}\equiv_{A}b_{j} for each i<j<ni<j<n and biAib<ib_{i}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{i}}_{A}b_{<i} for each i<ni<n.

Definition 2.5.

Given a class of AA-invariant types \mathcal{I}, an AA-invariant type p(x)p(x) is reliably in \mathcal{I} if it is in the largest class \mathcal{R}\subseteq\mathcal{I} satisfying that

  1. (1)

    for any p(x,y)p(x,y)\in\mathcal{R}, the restriction p(x)p(x) is in \mathcal{R},

  2. (2)

    for any p(x)p(x)\in\mathcal{R} and q(x,y)S(A)q(x,y)\in S(A) extending pAp\operatorname{\upharpoonright}A, there is an r(x,y)r(x,y)\in\mathcal{R} extending p(x)q(x,y)p(x)\cup q(x,y), and

  3. (3)

    for any p(x0)p(x_{0})\in\mathcal{R} and q(x0,,xn1)S(A)q(x_{0},\dots,x_{n-1})\in S(A) extending (pA)(x0)(p\operatorname{\upharpoonright}A)(x_{0}), if q(x¯)q(\bar{x}) is the type of an invariant sequence over AA, then there is an r(x¯)r(\bar{x})\in\mathcal{R} extending p(x0)p(xn1)q(x0,,xn1)p(x_{0})\cup\dots\cup p(x_{n-1})\cup q(x_{0},\dots,x_{n-1}).

If \mathcal{I} is the class of all AA-invariant types and p(x)p(x) is reliably in \mathcal{I}, then we say that p(x)p(x) is reliably AA-invariant. If \mathcal{I} is the class of AA-coheirs and p(x)p(x) is reliably in \mathcal{I}, we say that p(x)p(x) is a reliable AA-coheir.

The class \mathcal{R} exists by the Knaster-Tarski theorem (although it could be empty). Note that reliability clearly implies Kim-strictness. Given 2.15 and the fact that there are types over models with no strictly invariant extensions, we know that it does not always imply strictness.

Proposition 2.6.

If there is a formula φ(x,y)\varphi(x,y), a reliably AA-invariant type p(y)p(y), and an extendibly AA-invariant type q(y)q(y) such that pA=qAp\operatorname{\upharpoonright}A=q\operatorname{\upharpoonright}A and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide, then TT has CTP.

Proof.

Suppose that φ(x,y)\varphi(x,y) is kk-inconsistent on Morley sequences generated by qq. For each n<ωn<\omega, we will build a certain configuration (bσ)σ3n(b_{\sigma})_{\sigma\in 3^{n}} from which we can extract a kk-CTP tree of height nn. By compactness, we’ll have that TT has kk-CTP.

We will build the arrays (bσ)σ3n(b_{\sigma})_{\sigma\in 3^{n}} inductively. For n=1n=1, let b0b_{0} be any realization of pAp\operatorname{\upharpoonright}A. Then let b1pAb0b_{1}\models p\operatorname{\upharpoonright}Ab_{0} and b2qAb0b1b_{2}\models q\operatorname{\upharpoonright}Ab_{0}b_{1}. Note that b0b1b2b_{0}b_{1}b_{2} is an invariant sequence over AA. Let p1(x¯1)=p(x)p_{1}(\bar{x}_{1})=p(x).

Now suppose we are given the array (bσ)σ3n+1(b_{\sigma})_{\sigma\in 3^{n+1}} and a type pn+1(y¯n+1)p_{n+1}(\bar{y}_{n+1}) satisfying the following induction hypotheses:

  • The 33-element sequence (b0σ)σ3n,(b1σ)σ3n,(b2σ)σ3n\langle(b_{0\frown\sigma})_{\sigma\in 3^{n}},(b_{1\frown\sigma})_{\sigma\in 3^{n}},(b_{2\frown\sigma})_{\sigma\in 3^{n}}\rangle is invariant over AA.

  • pn+1(y¯n+1)p_{n+1}(\bar{y}_{n+1}) is reliably AA-invariant.

  • (b1σ)σ3npn+1A(b0σ)σ3n(b_{1\frown\sigma})_{\sigma\in 3^{n}}\models p_{n+1}\operatorname{\upharpoonright}A\cup(b_{0\frown\sigma})_{\sigma\in 3^{n}}.

  • b2n+1qA(biσ)i<2,σ3nb_{\langle 2\rangle^{n+1}}\models q\operatorname{\upharpoonright}A\cup(b_{i\frown\sigma})_{i<2,\sigma\in 3^{n}} (where 2n+1\langle 2\rangle^{n+1} is the sequence of length n+1n+1 consisting entirely of 22’s).

Let c0σ=bσc_{0\frown\sigma}=b_{\sigma} for each σ3n+1\sigma\in 3^{n+1}. Since pn+1(y¯n+1)p_{n+1}(\bar{y}_{n+1}) is reliably AA-invariant and since (c0σ)σ3n+1pn+1A(c_{0\frown\sigma})_{\sigma\in 3^{n+1}}\models p_{n+1}\operatorname{\upharpoonright}A, we can find a reliably AA-invariant type pn+2(y¯n+2)p_{n+2}(\bar{y}_{n+2}) extending pn+1(y¯n+1)tp((c0σ)σ3n+1/A)p_{n+1}(\bar{y}_{n+1})\cup\operatorname{tp}((c_{0\frown\sigma})_{\sigma\in 3^{n+1}}/A). Let (c1σ)σ3n+1pn+1A(c0σ)σ3n+1(c_{1\frown\sigma})_{\sigma\in 3^{n+1}}\models p_{n+1}\operatorname{\upharpoonright}A\cup(c_{0\frown\sigma})_{\sigma\in 3^{n+1}}. Then find (c2σ)σ3n+1(c_{2\frown\sigma})_{\sigma\in 3^{n+1}} such that (c2σ)σ3n+1Ai(ciσ)i<2,σ3n+1(c_{2\frown\sigma})_{\sigma\in 3^{n+1}}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{i}_{A}(c_{i\frown\sigma})_{i<2,\sigma\in 3^{n+1}} and c2n+1qA(ciσ)i<2,σ3n+1c_{\langle 2\rangle^{n+1}}\models q\operatorname{\upharpoonright}A\cup(c_{i\frown\sigma})_{i<2,\sigma\in 3^{n+1}} (which we can always do, since qq is extendibly AA-invariant). Finally, let (bσ)σ3n+2(b_{\sigma})_{\sigma\in 3^{n+2}} be (cσ)σ3n+2(c_{\sigma})_{\sigma\in 3^{n+2}}.

By induction we can perform this procedure for as many steps as we like. Fix some nn and consider the cube (bσ)σ3n(b_{\sigma})_{\sigma\in 3^{n}}. Let ff be the map from 2<n2^{<n} into 3n3^{n} taking τ2<n\tau\in 2^{<n} to the unique σ3n\sigma\in 3^{n} satisfying that σ(i)=0\sigma(i)=0 if and only if τ(i)=1\tau(i)=1, σ(i)=1\sigma(i)=1 if and only if τ(i)=0\tau(i)=0, and σ(i)=2\sigma(i)=2 if and only if τ(i)\tau(i) is not defined. Note that for any τ2<n\tau\in 2^{<n}, f(τ)f(\tau) only has 22’s at its end and either has a final segment of 22’s or has no 22’s at all.

We would like to argue that (bf(τ))τ2<n(b_{f(\tau)})_{\tau\in 2^{<n}} is a kk-CTP tree of height nn (i.e., satisfies the definition of kk-CTP except for the requirement that the tree have infinite height).

Claim. For any τ2<n\tau\in 2^{<n}, bf(τ)qA(bf(η))ητb_{f(\tau)}\models q\operatorname{\upharpoonright}A\cup(b_{f(\eta)})_{\eta\triangleright\tau}.

Proof of claim. Let γ\gamma be the longest initial segment of f(τ)f(\tau) not containing any 22’s. Let mm be the number of 22’s in f(τ)f(\tau). It follows by one of the induction hypotheses in the construction of the cube that bf(τ)qA(bγiε)i<2,ε3m1b_{f(\tau)}\models q\operatorname{\upharpoonright}A\cup(b_{\gamma\frown i\frown\varepsilon})_{i<2,\varepsilon\in 3^{m-1}}. The family (bγiε)i<2,ε3m1(b_{\gamma\frown i\frown\varepsilon})_{i<2,\varepsilon\in 3^{m-1}} contains bf(η)b_{f(\eta)} for any η2<n\eta\in 2^{<n} with ητ\eta\triangleright\tau, so the claim follows. \blacksquare

So, in particular, we have that for any path η2n\eta\in 2^{n}, {φ(x,bf(ηi)):i<n}\{\varphi(x,b_{f(\eta\operatorname{\upharpoonright}i)}):i<n\} is kk-inconsistent.

Claim. For any τ2<n\tau\in 2^{<n} and ητ0\eta\trianglerighteq\tau\frown 0, bf(η)pA(bf(γ))γτ1b_{f(\eta)}\models p\operatorname{\upharpoonright}A\cup(b_{f(\gamma)})_{\gamma\trianglerighteq\tau\frown 1}.

Proof of claim. Let δ\delta be the longest initial segment of f(τ)f(\tau) not containing any 22’s. We then have that f(γ)f(\gamma) has δ0\delta\frown 0 as an initial segment for any γτ1\gamma\trianglerighteq\tau\frown 1 (since ff switches 0 and 11). Let mm be the number of 22’s in f(τ)f(\tau). By the construction of the cube, we have that for any ε\varepsilon with initial segment δ1\delta\frown 1, bεpA(bδ0ξ)ξ3m1b_{\varepsilon}\models p\operatorname{\upharpoonright}A\cup(b_{\delta\frown 0\frown\xi})_{\xi\in 3^{m-1}}. The family (bδ0ξ)ξ3m1(b_{\delta\frown 0\frown\xi})_{\xi\in 3^{m-1}} includes bf(γ)b_{f(\gamma)} for any γτ1\gamma\trianglerighteq\tau\frown 1, so the claim follows. \blacksquare

This last claim implies that if X2nX\subseteq 2^{\leq n} is a right-comb, then it is a Morley sequence generated by p(y)p(y) when enumerated in decreasing lexicographic order. In particular, {φ(x,bf(τ)):τX}\{\varphi(x,b_{f(\tau)}):\tau\in X\} is consistent for any right-comb XX.

Therefore, by compactness, TT has kk-CTP.555We could also just build the kk-CTP tree directly with part of a cube of the form 3ω3^{\omega} (taken as a direct limit of the cubes 3n3^{n}), but this is a bit more annoying to actually write out.

One notable thing about the above proof is that, despite the advocacy for the concept of extendibly invariant types given in Section 2.1, extendibility is playing a relatively minor role in the proof. We never use the first or second property in 2.5 for the type p(y)p(y) and extendibility for q(y)q(y) seems to barely matter in that we throw away all of the ‘siblings’ of the realization of qq that we build at each step. This raises the question of whether the restriction to extendibly invariant types is necessary.

Question 2.7.

Say that an AA-invariant type is semi-reliably AA-invariant if it satisfies 2.5 with the first and second conditions removed (and with \mathcal{I} the class of all AA-invariant types). Is it true that if there is a set of parameters AA, a formula φ(x,y)\varphi(x,y), a semi-reliably AA-invariant type p(y)p(y), and an AA-invariant type q(y)q(y) such that pA=qAp\operatorname{\upharpoonright}A=q\operatorname{\upharpoonright}A and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide, then TT has CTP?

Another interesting and frustrating thing is that, while reliably AA-invariant types always exist over extension bases (as we will show in the next section), the construction of reliably invariant types in 2.15 and the construction of bi-invariant types in 1.5 seem to be rather incompatible. 2.15 relies heavily on knowing the precise type over the set AA, whereas 1.5 is a forcing construction that explicitly needs to avoid committing to a specific complete type before the end. It’s not even clear that the bi-invariant types constructed in 1.5 are extendibly bi-invariant (i.e., have the property that for any type q(x,y)q(x,y) over AA extending pAp\operatorname{\upharpoonright}A, there is an AA-bi-invariant type r(x,y)r(x,y) extending p(x)q(x,y)p(x)\cup q(x,y)).

Question 2.8.

Does the converse of 2.6 hold? Moreover, if TT has CTP, does there exist a reliable heir-coheir witnessing this? A reliably bi-invariant type? Extendibly bi-invariant?

It’s far from clear that the definition we’ve given in 2.5 is in some sense the ‘correct’ one. Its motivation is primarily that it is the strongest property that we can guarantee over invariance bases. In particular, while coheirs are always extendibly invariant, it is not clear whether heir-coheirs are always reliably invariant. If this were true we would be able to combine 1.8 and 2.6 and give a statement of Kim’s lemma that both characterizes NCTP and is non-trivial over arbitrary extension bases (and therefore over arbitrary models).

Question 2.9.

Are all heir-coheirs extendible heir-coheirs? In other words, if p(x)p(x) is an MM-heir-coheir and q(x,y)q(x,y) is a type over MM extending pMp\operatorname{\upharpoonright}M, is there an heir-coheir extending p(x)q(x,y)p(x)\cup q(x,y)?

Question 2.10.

Are all heir-coheirs reliable coheirs? Reliably invariant?

2.3. Existence of reliably invariant types

Our approach to building reliably invariant types is to build invariant types that are ‘as bi-invariant as possible.’ In order to manage our bookkeeping, we will use a fixed tuples of variables indexed by the monster model.

Definition 2.11.

Let 𝕆\mathbb{O} be a fixed monster model of TT. Let 𝕏=(xa)a𝕆\mathbb{X}=(x_{a})_{a\in\mathbb{O}} be a fixed enumeration of distinct variables indexed by 𝕆\mathbb{O}. Given a formula φ(𝕏)\varphi(\mathbb{X}) (with parameters in the monster) and σAut(𝕆)\sigma\in\operatorname{Aut}(\mathbb{O}), we write σvarφ\sigma^{\operatorname{var}}\cdot\varphi to represent the formula φ(𝕏)\varphi(\mathbb{X}) with its variables permuted by σ\sigma in the obvious way (but not its parameters). We write σparφ\sigma^{\operatorname{par}}\cdot\varphi to represent the formula φ(𝕏)\varphi(\mathbb{X}) with its parameters permuted by σ\sigma (but not its variables). We will use the same notation for partial types to indicate the image under the corresponding map.

The minimal monster type (over MM), ΘM(𝕏)\Theta_{M}(\mathbb{X}), is tp(𝕆/M)\operatorname{tp}(\mathbb{O}/M) with the obvious variable assignment. A monster type (over MM) is any consistent type extending ΘM(𝕏)\Theta_{M}(\mathbb{X}).

A partial type Σ(𝕏)\Sigma(\mathbb{X}) is AA-invariant if for any σAut(𝕆/A)\sigma\in\operatorname{Aut}(\mathbb{O}/A), σparΣ=Σ\sigma^{\operatorname{par}}\cdot\Sigma=\Sigma. Σ(𝕏)\Sigma(\mathbb{X}) is AA-co-invariant if for any σAut(𝕆/A)\sigma\in\operatorname{Aut}(\mathbb{O}/A), σvarΣ=Σ\sigma^{\operatorname{var}}\cdot\Sigma=\Sigma.

Note in particular that the minimal monster type over MM is MM-invariant (trivially) and MM-co-invariant. Also note that (the deductive closure of) the union of any two AA-invariant types is AA-invariant and of any two AA-co-invariant types is AA-co-invariant.

When a¯\bar{a} is a tuple of elements of the monster, we may write xa¯x_{\bar{a}} to represent the corresponding tuple of variables in 𝕏\mathbb{X}.

Lemma 2.12.

Fix a set AA. Let Ξ(𝕏)\Xi(\mathbb{X}) be a maximal666Whenever we talk about maximal partial types among some class, we mean maximal consistent partial types among that class. AA-co-invariant partial type. For any tuple a¯𝕆\bar{a}\in\mathbb{O} and any formula φ(xa¯)\varphi(x_{\bar{a}}), if φ(xa¯)\varphi(x_{\bar{a}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}), then there are a¯0,,a¯n1Aa¯\bar{a}_{0},\dots,\bar{a}_{n-1}\equiv_{A}\bar{a} such that Ξ(𝕏)i<nφ(xa¯i)\Xi(\mathbb{X})\vdash\bigvee_{i<n}\varphi(x_{\bar{a}_{i}}).

Proof.

Suppose that no such a¯0,,a¯n1\bar{a}_{0},\dots,\bar{a}_{n-1} exist. Then we’d have that {Ξ(𝕏){¬φ(xa¯):a¯Aa¯}\{\Xi(\mathbb{X})\cup\{\neg\varphi(x_{\bar{a}^{\prime}}):\bar{a}^{\prime}\equiv_{A}\bar{a}\} is consistent and AA-co-invariant, but then by maximality, we’d have that Ξ(𝕏)¬φ(xa¯)\Xi(\mathbb{X})\vdash\neg\varphi(x_{\bar{a}}), contradicting the fact that φ(xa¯)\varphi(x_{\bar{a}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}). ∎

Lemma 2.13.

Let Ξ(𝕏)\Xi(\mathbb{X}) be a maximal AA-co-invariant partial type, and let a¯0,,a¯n1\bar{a}_{0},\dots,\allowbreak\bar{a}_{n-1} be an invariant sequence over AA.

  1. (1)

    If a formula φ(xa¯0)\varphi(x_{\bar{a}_{0}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}), then i<nφ(xa¯i)\bigwedge_{i<n}\varphi(x_{\bar{a}_{i}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}).

  2. (2)

    If a global type p(xa¯0)p(x_{\bar{a}_{0}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}), then i<np(xa¯i)\bigwedge_{i<n}p(x_{\bar{a}_{i}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}).

Proof.

For 1, since φ(xa¯0)\varphi(x_{\bar{a}_{0}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}), we have by 2.12 that there are b¯0,,b¯k1\bar{b}_{0},\dots,\bar{b}_{k-1} such that Ξ(𝕏)j<kφ(b¯i)\Xi(\mathbb{X})\vdash\bigvee_{j<k}\varphi(\bar{b}_{i}) and b¯jAa¯0\bar{b}_{j}\equiv_{A}\bar{a}_{0} for each j<kj<k.

For each positive i<ni<n, let p(x¯)p(\bar{x}) be an MM-invariant type such that a¯ipAa¯<i\bar{a}_{i}\models p\operatorname{\upharpoonright}A\bar{a}_{<i}.

By induction, we can build a forest (d¯η)ηkn{}(\bar{d}_{\eta})_{\eta\in k^{\leq n}\setminus\{\varnothing\}} such that

  • for each ηk<n\eta\in k^{<n}, d¯η0d¯η1d¯η(n1)Mc¯0c¯1c¯n1\bar{d}_{\eta\frown 0}\bar{d}_{\eta\frown 1}\dots\bar{d}_{\eta\frown(n-1)}\equiv_{M}\bar{c}_{0}\bar{c}_{1}\dots\bar{c}_{n-1} and

  • for each ηk<n{}\eta\in k^{<n}\setminus\{\varnothing\}, d¯ηpn(η)d¯η\bar{d}_{\eta}\models p_{n-\ell(\eta)}\operatorname{\upharpoonright}\bar{d}_{\triangleright\eta} (where (η)\ell(\eta) is the length of η\eta).

Let q(𝕏)q(\mathbb{X}) be a completion of Ξ(𝕏)\Xi(\mathbb{X}). Since Ξ(𝕏)j<kφ(xd¯ηj)\Xi(\mathbb{X})\vdash\bigvee_{j<k}\varphi(x_{\bar{d}_{\eta\frown j}}) for each ηk<n\eta\in k^{<n}, we can find a path βkn\beta\in k^{n} such that q(𝕏)0<inφ(xd¯βi)q(\mathbb{X})\vdash\bigwedge_{0<i\leq n}\varphi(x_{\bar{d}_{\beta\operatorname{\upharpoonright}i}}). This implies that 0<inφ(xd¯βi)\bigwedge_{0<i\leq n}\varphi(x_{\bar{d}_{\beta\operatorname{\upharpoonright}i}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}), so by AA-co-invariance, i<nφ(xa¯i)\bigwedge_{i<n}\varphi(x_{\bar{a}_{i}}) is consistent with Ξ(𝕏)\Xi(\mathbb{X}) as well.

2 follows by compactness. ∎

The following proposition says that we can rely on the existence of reliable types.

Theorem 2.14.

Fix a set of parameters AA and let \mathcal{I} be a class of AA-invariant types. Suppose furthermore that

  1. (1)

    every type over AA extends to a member of \mathcal{I},

  2. (2)

    for each tuple of variables x¯\bar{x}, Sx¯(𝕆)S_{\bar{x}}(\mathbb{O})\cap\mathcal{I} is closed,

  3. (3)

    \mathcal{I} is invariant under renaming variables, and

  4. (4)

    any AA-invariant type p(x¯)p(\bar{x}) is in \mathcal{I} if and only if every restriction of it to finitely many variables is in \mathcal{I}.

Then for every q(x¯)S(A)q(\bar{x})\in S(A), there is an AA-invariant p(x¯)q(x¯)p(\bar{x})\supset q(\bar{x}) that is reliably in \mathcal{I}.

Proof.

Let FS𝕏(𝕆)F\subseteq S_{\mathbb{X}}(\mathbb{O}) be the set of extensions of ΘA(𝕏)\Theta_{A}(\mathbb{X}) that are in \mathcal{I}. Note that since FF is the intersection of two closed sets, it is closed. Let Σ(𝕏)\Sigma(\mathbb{X}) be the global partial type corresponding to the closed set FF.

Claim. Σ(𝕏)\Sigma(\mathbb{X}) is AA-co-invariant.

Proof of claim. Fix σAut(𝕆/A)\sigma\in\operatorname{Aut}(\mathbb{O}/A). By 3 and 4, S𝕏(𝕆)\mathcal{I}\cap S_{\mathbb{X}}(\mathbb{O}) is fixed by the action of σ\sigma on the variables 𝕏\mathbb{X}. Since the set of global completions of ΘA(𝕏)\Theta_{A}(\mathbb{X}) is as well, this implies that FF and therefore Σ(𝕏)\Sigma(\mathbb{X}) is fixed under the action of σ\sigma. Since we can do this for any σ\sigma, Σ(𝕏)\Sigma(\mathbb{X}) is AA-co-invariant. \blacksquare

Let Ξ(𝕏)\Xi(\mathbb{X}) be a maximal AA-co-invariant extension of Σ(𝕏)\Sigma(\mathbb{X}) (which always exists by Zorn’s lemma). Note that Ξ(𝕏)\Xi(\mathbb{X}) is also AA-invariant.777Normally we couldn’t require a maximal AA-co-invariant partial type to be AA-invariant, but since every complete type in FF is AA-invariant, any partial type extending Σ(𝕏)\Sigma(\mathbb{X}) is AA-invariant. Also note that since Ξ(𝕏)\Xi(\mathbb{X}) is a monster type, we have that if aa is a realization of p(x)S(A)p(x)\in S(A), then Ξ(𝕏)p(xa)\Xi(\mathbb{X})\vdash p(x_{a}).

Now let 𝒮\mathcal{S} be the class of AA-invariant types p(x¯)p(\bar{x}) satisfying the following property:

  • ()(\ast)

    For any finite sub-tuple y0,,yn1y_{0},\dots,y_{n-1} of x¯\bar{x}, let q(y¯)q(\bar{y}) be the restriction of pp to the variables y¯\bar{y}. If a0a1an1qAa_{0}a_{1}\dots a_{n-1}\models q\operatorname{\upharpoonright}A, then q(xa0,xa1,,xan1)Ξ(𝕏)q(x_{a_{0}},x_{a_{1}},\dots,x_{a_{n-1}})\cup\Xi(\mathbb{X}) is consistent.

Note that since Ξ(𝕏)\Xi(\mathbb{X}) is AA-co-invariant, the above property doesn’t depend on the choice of a¯\bar{a}. We would like to show that the class 𝒮\mathcal{S} satisfies the closure properties in 2.5 relative to \mathcal{I} and therefore every type in 𝒮\mathcal{S} is reliably in \mathcal{I}.

The first closure property is immediate. Note that for small tuples of variables, the second closure property is also immediate. If p(x¯)p(\bar{x}) is an AA-invariant type in 𝒮\mathcal{S} and q(x¯,y¯)q(\bar{x},\bar{y}) is a type in Sx¯y¯(A)S_{\bar{x}\bar{y}}(A) extending pAp\operatorname{\upharpoonright}A, then given a¯b¯q\bar{a}\bar{b}\models q, we have that Ξ(𝕏)q(xa¯,xb¯)\Xi(\mathbb{X})\vdash q(x_{\bar{a}},x_{\bar{b}}) and so p(xa¯)q(xa¯,xb¯)p(x_{\bar{a}})\cup q(x_{\bar{a}},x_{\bar{b}}) is automatically consistent. For types with large tuples of variables, the result follows from the fact that for a fixed type q(x¯)q(\bar{x}) over AA, the set of global extensions of q(x¯)q(\bar{x}) satisfying ()(\ast) is closed and so the required result follows by compactness.

Now we need to show that 𝒮\mathcal{S} satisfies the third closure property. Again, we first show this for small tuples of variables.

Claim. For any p(x¯)S(A)p(\bar{x})\in S(A) (with x¯\bar{x} a small tuple of variables) and invariant sequence (a¯i)i<n(\bar{a}_{i})_{i<n} over AA, if a¯0p\bar{a}_{0}\models p and q(xa¯0)q(x_{\bar{a}_{0}}) is a global extension of p(xa¯0)p(x_{\bar{a}_{0}}) such that Ξ(𝕏)q(xa¯0)\Xi(\mathbb{X})\cup q(x_{\bar{a}_{0}}) is consistent, then Ξ(𝕏)q(xa¯0)q(xa¯1)q(xa¯n1)r(xa¯0,,xa¯n1)\Xi(\mathbb{X})\cup q(x_{\bar{a}_{0}})\cup q(x_{\bar{a}_{1}})\cup\dots\cup q(x_{\bar{a}_{n-1}})\cup r(x_{\bar{a}_{0}},\dots,x_{\bar{a}_{n-1}}) is consistent, where r(xa¯0,,xa¯n1)=tp(a¯0a¯n1/A)r(x_{\bar{a}_{0}},\dots,x_{\bar{a}_{n-1}})=\operatorname{tp}(\bar{a}_{0}\dots\bar{a}_{n-1}/A).

Proof of claim. Since (a¯i)i<n(\bar{a}_{i})_{i<n} is an invariant sequence over AA, we have by 2.13 that Ξ(𝕏)q(xa¯0)q(xa¯1)q(xa¯n1)\Xi(\mathbb{X})\cup q(x_{\bar{a}_{0}})\cup q(x_{\bar{a}_{1}})\cup\dots\cup q(x_{\bar{a}_{n-1}}) is consistent. Since Ξ(𝕏)\Xi(\mathbb{X}) is a monster type, we have that Ξ(𝕏)r(xa¯0,,xa¯n1)\Xi(\mathbb{X})\vdash r(x_{\bar{a}_{0}},\dots,x_{\bar{a}_{n-1}}), so the required partial type is consistent. \blacksquare

Therefore if q(xa¯0)q(x_{\bar{a}_{0}}) is any global completion of p(xa¯0)=tp(a0/A)p(x_{\bar{a}_{0}})=\operatorname{tp}(a_{0}/A) that is consistent with Ξ(𝕏)\Xi(\mathbb{X}) and (a¯i)i<n(\bar{a}_{i})_{i<n} is an invariant sequence over AA, we have that for any b¯\bar{b}, if a¯0qAb¯\bar{a}_{0}\models q\operatorname{\upharpoonright}A\bar{b}, then we can find a global completion s(xa¯0,,xa¯n1)s(x_{\bar{a}_{0}},\dots,x_{\bar{a}_{n-1}}) consistent with Ξ(𝕏)\Xi(\mathbb{X}) and a¯1a¯n1Aa¯0a¯1a¯n1\bar{a}^{\prime}_{1}\dots\bar{a}^{\prime}_{n-1}\equiv_{A\bar{a}_{0}}\bar{a}_{1}\dots\bar{a}_{n-1} such that a¯0a¯1a¯n1sAb¯\bar{a}_{0}\bar{a}^{\prime}_{1}\dots\bar{a}^{\prime}_{n-1}\models s\operatorname{\upharpoonright}A\bar{b}.

Again the third closure condition for types with large tuples of variables follows from compactness and the fact that if (a¯ib¯i)i<n(\bar{a}_{i}\bar{b}_{i})_{i<n} is an AA-invariant sequence, then (a¯i)i<n(\bar{a}_{i})_{i<n} is an AA-invariant sequence as well.

Therefore 𝒮\mathcal{S} satisfies the closure conditions in 2.5 relative to \mathcal{I} and so 𝒮\mathcal{S}\subseteq\mathcal{R}, since \mathcal{R} is the largest class of types satisfying this condition. Finally, if q(x¯)q(\bar{x}) is a type in a small number of variables, we can find a global AA-invariant type p(x¯)q(x¯)p(\bar{x})\supseteq q(\bar{x}), which is therefore in the class 𝒮\mathcal{S} and is therefore reliably in \mathcal{I}. For types with large tuples of variables, the extension again exists by compactness. ∎

Corollary 2.15.

Fix a set of parameters AA.

  1. (1)

    If AA is an invariance base, then every type p(x)Sx(A)p(x)\in S_{x}(A) extends to a reliably AA-invariant type.

  2. (2)

    If AA is a model, then every type p(x)Sx(A)p(x)\in S_{x}(A) extends to a reliable AA-coheir.

Proof.

If AA is an invariance base, we can apply 2.14 with \mathcal{I} equal to the class of AA-invariant global types. If AA is a model, we can apply 2.14 with \mathcal{I} equal to the class of AA-coheirs. ∎

Corollary 2.16.

If TT is NCTP, then a formula φ(x,b)\varphi(x,b) Kim-forks with regards to extendibly invariant types over an invariance base AA if and only if it Kim-divides with regards to extendibly invariant types over AA.

Proof.

Suppose that φ(x,b)\varphi(x,b) Kim-forks over AA. Let φ(x,b)i<nψi(x,ci)\varphi(x,b)\vdash\bigvee_{i<n}\psi_{i}(x,c_{i}) such that ψi(x,ci)\psi_{i}(x,c_{i}) Kim-divides over AA for each i<ni<n. Let p(y,z0,,zn1)p(y,z_{0},\dots,z_{n-1}) be a reliably AA-invariant type extending tp(bc0cn1/A)\operatorname{tp}(bc_{0}\dots c_{n-1}/A). Let (djeij)i<n,j<ω(d^{j}e_{i}^{j})_{i<n,j<\omega} be a Morley sequence generated by pp. Note that d<ωd^{<\omega} and ei<ωe^{<\omega}_{i} for each i<ni<n are Morley sequences in reliably AA-invariant types. By 2.6, we have that {ψi(x,eij):j<ω}\{\psi_{i}(x,e_{i}^{j}):j<\omega\} is inconsistent for each i<ni<n. Therefore, by the standard argument, we have that {i<nψi(x,eij):j<ω}\{\bigvee_{i<n}\psi_{i}(x,e_{i}^{j}):j<\omega\} is inconsistent. By indiscernibility, this implies that {φ(x,dj):j<ω}\{\varphi(x,d^{j}):j<\omega\} is inconsistent, or, in other words, that φ(x,b)\varphi(x,b) Kim-divides over AA with regards to a reliably AA-invariant type. Any reliably AA-invariant type is extendibly AA-invariant, so we are done. ∎

3. Local Character

3.1. Uncountable languages and dual local character

In the following we will give a characterization of NCTP in terms of a weak dual local character under the assumption that there is a measurable cardinal κ>|T|\kappa>|T|. We don’t expect this to be necessary, but given the unclear usefulness of this characterization and how straightforward the proof with a measurable cardinal is, we have decided to not pursue a stronger result for the time being. In the next proposition we will also extend 1.5 to uncountable languages.

Proposition 3.1.

Let TT be a theory in a (possibly uncountable) language \mathcal{L}. Assume TT has kk-CTP witness by the formula φ(x,y)\varphi(x,y).

  1. (1)

    There is a model MM with |M|=|||M|=|\mathcal{L}|, an MM-heir-coheir p(y)p(y), and an MM-coheir q(y)q(y) such that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide.

  2. (2)

    For any regular cardinal κ>||\kappa>|\mathcal{L}|, there is a model MM with |M|=κ|M|=\kappa, a continuous chain (Nε)ε<κ(N_{\varepsilon})_{\varepsilon<\kappa} of elementary substructures with M=ε<κNεM=\bigcup_{\varepsilon<\kappa}N_{\varepsilon} and |Nε|<κ|N_{\varepsilon}|<\kappa for each ε<κ\varepsilon<\kappa, and a sequence (eε)ε<κ(e_{\varepsilon})_{\varepsilon<\kappa} such that eεNεe_{\varepsilon}\notin N_{\varepsilon} for each ε<κ\varepsilon<\kappa, {φ(x,eε):ε<κ}\{\varphi(x,e_{\varepsilon}):\varepsilon<\kappa\} is kk-inconsistent, and for each ε<κ\varepsilon<\kappa, there is an NεN_{\varepsilon}-heir-coheir sεtp(eε/Nε)s_{\varepsilon}\supseteq\operatorname{tp}(e_{\varepsilon}/N_{\varepsilon}) such that φ(x,eε)\varphi(x,e_{\varepsilon}) does not sεs_{\varepsilon}-Kim-divide.

Proof.

The proofs of 1 and 2 are nearly the same. If we are proving 1, let κ=||\kappa=|\mathcal{L}|, and if we are proving 2, let κ\kappa be as in the statement of the proposition. When we say a model NN is small, we mean that |N|κ|N|\leq\kappa if we are proving 1 and |N|<κ|N|<\kappa if we are proving 22.

Let (bσ)σ2<ω(b_{\sigma})_{\sigma\in 2^{<\omega}} witness that the formula φ(x,y)\varphi(x,y) has kk-CTP. Let M0{bσ:σ2<ω}M_{0}\supseteq\{b_{\sigma}:\sigma\in 2^{<\omega}\} be a small model. Expand M0M_{0} with the following predicates:

  • A unary predicate R0R_{0}, which defines the set {bσ:σ2<ω}\{b_{\sigma}:\sigma\in 2^{<\omega}\}.

  • A unary predicate X0X_{0}, which defines the set {bσ:σ1}\{b_{\sigma}:\sigma\trianglerighteq 1\}.

  • A partial order \trianglelefteq on R0R_{0} defined by bσbτb_{\sigma}\trianglelefteq b_{\tau} if and only if στ\sigma\trianglelefteq\tau.

  • The functions bσbσ0b_{\sigma}\mapsto b_{\sigma\frown 0} and bσbσ1b_{\sigma}\mapsto b_{\sigma\frown 1}, which we will write as b0b\frown 0 and b1b\frown 1.

Let 0\mathcal{L}_{0} be the expanded language.

Let (β(α),γ(α))α<κ(\beta(\alpha),\gamma(\alpha))_{\alpha<\kappa} be an enumeration of κ2\kappa^{2} with the property that for each pair δ,ε<κ\delta,\varepsilon<\kappa, the set {α<κ:β(α),γ(α)=δ,ε}\{\alpha<\kappa:\langle\beta(\alpha),\gamma(\alpha)\rangle=\langle\delta,\varepsilon\rangle\} is cofinal in κ\kappa. (Such an enumeration can be defined using any bijection between κ\kappa and κ3\kappa^{3}.)

Let (Rα)α<κ(R_{\alpha})_{\alpha<\kappa} and (Xα)α<κ(X_{\alpha})_{\alpha<\kappa} be two sequences of distinct unary predicates. For each α<κ\alpha<\kappa, let α=0{Rβ,Xβ:βα}\mathcal{L}_{\alpha}=\mathcal{L}_{0}\cup\{R_{\beta},X_{\beta}:\beta\leq\alpha\}. We will inductively build a sequence of elements (bα)α<κ(b_{\alpha})_{\alpha<\kappa} and a chain of small structures (Mα)α<κ(M_{\alpha})_{\alpha<\kappa} (with MαM_{\alpha} an α\mathcal{L}_{\alpha}-structure) where for any β<α\beta<\alpha, MβM_{\beta} is an β\mathcal{L}_{\beta}-elementary substructure of MαM_{\alpha}. Each RαR_{\alpha} will be downwards closed under \trianglerighteq and closed under the functions bb0b\mapsto b\frown 0 and bb1b\mapsto b\frown 1. We will call any set closed under these a closed tree. Furthermore, for β<α\beta<\alpha, we will have RαRβR_{\alpha}\subseteq R_{\beta}, XαXβX_{\alpha}\subseteq X_{\beta}, and XαRαX_{\alpha}\subseteq R_{\alpha}. (If we are proving 11, the XαX_{\alpha}’s will eventually be part of the filter that we use to define our two coheirs.)

Let c0=bc_{0}=b_{\varnothing}. At stage α<κ\alpha<\kappa, suppose we have our α\mathcal{L}_{\alpha}-structure MαM_{\alpha} with |Mα|=κ|M_{\alpha}|=\kappa. Furthermore, suppose that XαX_{\alpha} is dense above cα1c_{\alpha}\frown 1 in RαR_{\alpha} (i.e., for any cRα(Mα)c\in R_{\alpha}(M_{\alpha}) with cbαc\trianglerighteq b_{\alpha}, there is a dRα(Mα)d\in R_{\alpha}(M_{\alpha}) such that dcd\trianglerighteq c and dXα(Mα)d\in X_{\alpha}(M_{\alpha})). Let (ψβ(δ)α(y,z))δ<κ(\psi^{\alpha}_{\beta(\delta)}(y,z))_{\delta<\kappa} be an enumeration of all α(Mα)\mathcal{L}_{\alpha}(M_{\alpha})-formulas. To get Mα+1M_{\alpha+1} and cα+1c_{\alpha+1}, perform the following construction:

  • If ψβ(α)γ(α)(y,z)\psi^{\gamma(\alpha)}_{\beta(\alpha)}(y,z) has already been defined and there is an α\mathcal{L}_{\alpha}-elementary extension NMαN\succeq M_{\alpha}, a closed tree SRα(N)S\subseteq R_{\alpha}(N) with SRα(Mα)S\supseteq R_{\alpha}(M_{\alpha}), and a dd (in the Th(Mα)\operatorname{Th}(M_{\alpha})-monster) such that Xβ(N)X_{\beta}(N) is dense above cβ+1c_{\beta+1} in SS for each β<α\beta<\alpha and {bSXα(N):ψβ(α)γ(α)(b,d)}\{b\in S\cap X_{\alpha}(N):\psi^{\gamma(\alpha)}_{\beta(\alpha)}(b,d)\} is somewhere dense above cα1c_{\alpha}\frown 1 in SS: By Löwenheim-Skolem, we can assume that NN is small. Let Mα+1NdM_{\alpha+1}\supseteq Nd be a small elementary extension. Find cα+1Xα(N)c_{\alpha+1}\in X_{\alpha}(N) with cα+1cα1c_{\alpha+1}\trianglerighteq c_{\alpha}\frown 1 such that ψβ(α)γ(α)(cα+1,d)\psi^{\gamma(\alpha)}_{\beta(\alpha)}(c_{\alpha+1},d) holds. By compactness, we may assume that for any bRα(Mα)b\in R_{\alpha}(M_{\alpha}), ¬(bcα+1)\neg(b\trianglerighteq c_{\alpha+1}). Let Rα+1(Mα+1)=SR_{\alpha+1}(M_{\alpha+1})=S, and let

    Xα+1(Mα+1)={bSXα(Mα+1):bcα+11,ψβ(α)γ(α)(b,d)}.X_{\alpha+1}(M_{\alpha+1})=\{b\in S\cap X_{\alpha}(M_{\alpha+1}):b\trianglerighteq c_{\alpha+1}\frown 1,~{}\psi^{\gamma(\alpha)}_{\beta(\alpha)}(b,d)\}.

    Note that Xα+1(Mα+1)X_{\alpha+1}(M_{\alpha+1}) is dense above cα+11c_{\alpha+1}\frown 1 in Rα+1(Mα+1)R_{\alpha+1}(M_{\alpha+1}).

  • If the condition in the previous bullet point fails: Find a small α\mathcal{L}_{\alpha}-elementary extension Mα+1MαM_{\alpha+1}\succeq M_{\alpha} with a cα+1cα1c_{\alpha+1}\trianglerighteq c_{\alpha}\frown 1 such that cα+1Xα(Mα+1)c_{\alpha+1}\in X_{\alpha}(M_{\alpha+1}) and such that for any bRα(Mα)b\in R_{\alpha}(M_{\alpha}), ¬(bcα+1)\neg(b\trianglerighteq c_{\alpha+1}). (This is always possible because Rα(Mα)R_{\alpha}(M_{\alpha}) is a closed tree.) Let Rα+1(Mα+1)=Rα(Mα+1)R_{\alpha+1}(M_{\alpha+1})=R_{\alpha}(M_{\alpha+1}), and let Xα+1(Mα+1)={bXα(Mα=1):bcα+11}X_{\alpha+1}(M_{\alpha+1})=\{b\in X_{\alpha}(M_{\alpha=1}):b\trianglerighteq c_{\alpha+1}\frown 1\}.

Note that in both cases we have ensured that cα+1Xαc_{\alpha+1}\in X_{\alpha} and that cα+1c_{\alpha+1} is not \trianglelefteq-upper-bounded by any element of Rα(Mα)R_{\alpha}(M_{\alpha}). Also note that for every βα\beta\leq\alpha, we have that Xβ(Mα+1)X_{\beta}(M_{\alpha+1}) is dense above cβ+1c_{\beta+1} in Rα+1(Mα+1)R_{\alpha+1}(M_{\alpha+1}).

Just before limit stage α<κ\alpha<\kappa, suppose that we have MβM_{\beta} and cβc_{\beta} for all β<α\beta<\alpha. Let N=β<αMβN=\bigcup_{\beta<\alpha}M_{\beta}, which we can regard as an <α\mathcal{L}_{<\alpha}-structure (where <αβ<αβ\mathcal{L}_{<\alpha}\coloneqq\bigcup_{\beta<\alpha}\mathcal{L}_{\beta}). By compactness, we can find a small model MαNM_{\alpha}\succeq N, a closed tree Rα(Mα)MαR_{\alpha}(M_{\alpha})\subseteq M_{\alpha}, a set Xα(Mα)Rα(Mα)X_{\alpha}(M_{\alpha})\subseteq R_{\alpha}(M_{\alpha}), and cαRα(Mα)c_{\alpha}\in R_{\alpha}(M_{\alpha}) such that for all β<α\beta<\alpha, Rα(Mα)Rβ(Mα)R_{\alpha}(M_{\alpha})\subseteq R_{\beta}(M_{\alpha}), Xα(Mα)Xβ(Mα)X_{\alpha}(M_{\alpha})\subseteq X_{\beta}(M_{\alpha}), cαcβc_{\alpha}\trianglerighteq c_{\beta}, and Xα(Mα)X_{\alpha}(M_{\alpha}) is dense above cα1c_{\alpha}\frown 1 in Rα(Mα)R_{\alpha}(M_{\alpha}). If we are proving 2 and {Xβ(y):β<α}\{X_{\beta}(y):\beta<\alpha\} axiomatizes a complete <α\mathcal{L}_{<\alpha}-type rα(y)r_{\alpha}(y) over NN, we may moreover assume that there is a dαMαd_{\alpha}\in M_{\alpha} realizing rα(y)r_{\alpha}(y) such that cβdαcαc_{\beta}\trianglelefteq d_{\alpha}\trianglelefteq c_{\alpha} for all β<α\beta<\alpha.

After the construction is completed, let M=α<κMαM=\bigcup_{\alpha<\kappa}M_{\alpha}. Note that |M|=κ|M|=\kappa by induction.

Proof of 1. Let R=α<κRα(Mα)R=\bigcup_{\alpha<\kappa}R_{\alpha}(M_{\alpha}). Note that RR is a closed tree.

Let \mathcal{F} be the filter generated by {Xα(M):α<κ}{Y}\{X_{\alpha}(M):\alpha<\kappa\}\cup\{Y\}, where Yα<κ{bR:bcα0}Y\coloneqq\bigcup_{\alpha<\kappa}\{b\in R:b\trianglerighteq c_{\alpha}\frown 0\}.

Claim 1. For every β<κ\beta<\kappa, Xβ(M)X_{\beta}(M) is dense above cβ+1c_{\beta+1} in RR.

Proof of claim. Fix a β<κ\beta<\kappa. At each stage α<κ\alpha<\kappa with αβ\alpha\geq\beta, we ensured that Xβ(Mα+1)X_{\beta}(M_{\alpha+1}) is dense above cβ+1c_{\beta+1} in Rα+1(Mα+1)R_{\alpha+1}(M_{\alpha+1}). Fix bRb\in R with bcβ+1b\trianglerighteq c_{\beta+1}. Since RR is the union of a chain, there is an α<κ\alpha<\kappa with αβ\alpha\geq\beta such that bRα+1(Mα+1)b\in R_{\alpha+1}(M_{\alpha+1}). At stage α\alpha, we ensured that Xβ(Mα+1)X_{\beta}(M_{\alpha+1}) is dense above cβ+1c_{\beta+1}. Therefore there is a dXβ(Mα+1)Rα+1(Mα+1)d\in X_{\beta}(M_{\alpha+1})\cap R_{\alpha+1}(M_{\alpha+1}) with dbd\trianglerighteq b. Since Rα+1(Mα+1)RR_{\alpha+1}(M_{\alpha+1})\subseteq R, we have that dRd\in R. By elementarity, we also have that bXβ(M)b\in X_{\beta}(M). Since we can do this for any bRb\in R with bcβ+1b\trianglerighteq c_{\beta+1}, we have that Xβ(M)X_{\beta}(M) is dense above cβ+1c_{\beta+1} in RR. \blacksquare

The previous claim implies that the filter \mathcal{F} is everywhere somewhere dense. Therefore we can extend it to an everywhere somewhere dense ultrafilter 𝒰\mathcal{U} by the same argument as in 1.4.

Let 𝒢\mathcal{G} be the filter generated by {Xα(M):α<κ}{{cα:α<κ}}\{X_{\alpha}(M):\alpha<\kappa\}\cup\{\{c_{\alpha}:\alpha<\kappa\}\}. Since cαXβ(Mβ)Xβ(M)c_{\alpha}\in X_{\beta}(M_{\beta})\subseteq X_{\beta}(M) for any β<α<κ\beta<\alpha<\kappa, we have that 𝒢\mathcal{G} is a proper filter. Let 𝒱\mathcal{V} be an ultrafilter extending 𝒢\mathcal{G}.

Let p(y)p(y) be the global average type of 𝒰\mathcal{U}, and let q(y)q(y) be the global average type of 𝒱\mathcal{V}.

Claim 2. pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M. Moreover, this type is axiomatized by {Xα(y):α<κ}\{X_{\alpha}(y):\alpha<\kappa\}.

Proof of claim. Fix an MM-formula ψ(y)p(y)\psi(y)\in p(y). ψ(y)\psi(y) is actually an MβM_{\beta} formula for some β<κ\beta<\kappa. By the choice of our enumeration (β(α),γ(α))α<κ(\beta(\alpha),\gamma(\alpha))_{\alpha<\kappa}, this means that there was a stage α\alpha at which ψγ(α)β(α)(y,z)\psi^{\beta(\alpha)}_{\gamma(\alpha)}(y,z) was defined and equal to ψ(y)\psi(y) (where zz is a dummy variable). Since ψ(y)p(y)\psi(y)\in p(y) and since 𝒰\mathcal{U} is everywhere somewhere dense, we must have chosen the first bullet point at this stage. (In particular, the required condition is satisfied with N=MN=M and S=RS=R.) Hence ψ(b)\psi(b) holds for all bXα+1(Mα+1)b\in X_{\alpha+1}(M_{\alpha+1}). By elementarity, this implies that ψ(b)\psi(b) holds for all bXα+1(M)b\in X_{\alpha+1}(M) as well, implying that ψ(cδ)\psi(c_{\delta}) holds for all δ>α+1\delta>\alpha+1. Therefore ψ(y)q(y)\psi(y)\in q(y) as well. Since we can do this for every MM-formula ψ(y)p(y)\psi(y)\in p(y), we have that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M and that this type is axiomatized by {Xα(y):α<κ}\{X_{\alpha}(y):\alpha<\kappa\}. \blacksquare

Claim 3. pp is an heir over MM.

Proof of claim. Fix an MM-formula ψ(y,z)\psi(y,z). Once again, there must have been a stage α\alpha at which ψγ(α)β(α)\psi^{\beta(\alpha)}_{\gamma(\alpha)} was defined and equal to ψ\psi. Suppose that there is a dd in the monster such that ψ(y,d)p(y)\psi(y,d)\in p(y). Since 𝒰\mathcal{U} is everywhere somewhere dense, this implies that the conditions for the first bullet point were met (with N=MN=M and S=RS=R and the parameter dd). Therefore we added a parameter ee to Mα+1M_{\alpha+1} such that ψ(b,e)\psi(b,e) holds for all bXα+1(Mα+1)b\in X_{\alpha+1}(M_{\alpha+1}). By elementarity, this implies that ψ(b,e)\psi(b,e) holds for all bXα+1(M)b\in X_{\alpha+1}(M). Therefore ψ(y,e)p(y)\psi(y,e)\in p(y). Since we can do this for any MM-formula ψ(y,z)\psi(y,z), we have that p(y)p(y) is an heir of pMp\operatorname{\upharpoonright}M over MM. \blacksquare

For any nn, the set of finite right-combs of cardinality nn in (bσ)σ<ω(b_{\sigma})_{\sigma<\omega} is definable in the language 0\mathcal{L}_{0}. Therefore, by elementarity, we have that for any right-comb ZRZ\subseteq R, {φ(x,b):bZ}\{\varphi(x,b):b\in Z\} is consistent. Likewise, being a \trianglelefteq-increasing sequence of size nn is a definable property, so for every path ZZ in RR, {φ(x,b):bZ}\{\varphi(x,b):b\in Z\} is kk-inconsistent.

Claim 4. φ(x,y)\varphi(x,y) does not pp-Kim-divide.

Proof of claim. The set Yα<ω{bR:bcα0}Y\coloneqq\bigcup_{\alpha<\omega}\{b\in R:b\trianglerighteq c_{\alpha}\frown 0\} is in 𝒰\mathcal{U}. Likewise, for each α<κ\alpha<\kappa, we have that the set {bR:bcα}\{b\in R:b\trianglerighteq c_{\alpha}\} is in 𝒰\mathcal{U}. Note that if α0,α1,,αn1\alpha_{0},\alpha_{1},\dots,\alpha_{n-1} is an increasing sequence of ordinals and if bicαi0b_{i}\trianglerighteq c_{\alpha_{i}}\frown 0 for each i<ni<n, then {bi:i<n}\{b_{i}:i<n\} is a right-comb.

Let (en)n<ω(e_{n})_{n<\omega} be a Morley sequence generated by pp. By the argument in the paragraph just before Claim 4, we have that for any finite right-comb ZYZ\subseteq Y, {φ(x,b):bZ}\{\varphi(x,b):b\in Z\} is consistent. Suppose that for some n<ωn<\omega (possibly 0), we’ve shown that for every finite right-comb ZYZ\subseteq Y, {φ(x,b):bZ}{φ(x,ei):i<n}\{\varphi(x,b):b\in Z\}\cup\{\varphi(x,e_{i}):i<n\} is consistent. For any such ZZ, there is an α\alpha such that for any dYd\in Y with dcαd\trianglerighteq c_{\alpha}, Z{d}Z\cup\{d\} is a right-comb. This implies that {φ(x,b):bZ}{φ(x,ei):i<n+1}\{\varphi(x,b):b\in Z\}\cup\{\varphi(x,e_{i}):i<n+1\} is consistent. Hence, by induction, we have that {φ(x,ei):i<ω}\{\varphi(x,e_{i}):i<\omega\} is consistent and so φ(x,y)\varphi(x,y) does not pp-Kim-divide. \blacksquare

Finally, note that since {cα:α<κ}𝒱\{c_{\alpha}:\alpha<\kappa\}\in\mathcal{V} and since {φ(x,cα):α<κ}\{\varphi(x,c_{\alpha}):\alpha<\kappa\} is kk-inconsistent, we have that {φ(x,ei):i<ω}\{\varphi(x,e_{i}):i<\omega\} is kk-inconsistent for any Morley sequence (ei)i<ω(e_{i})_{i<\omega} generated by qq. Therefore pp and qq satisfy the required conditions.

Proof of 2. Say that an MβM_{\beta}-formula ψ\psi is covered by α<κ\alpha<\kappa if ψγ(α)β(α)\psi^{\beta(\alpha)}_{\gamma(\alpha)} was defined and equal to ψ\psi at stage α\alpha. Since κ\kappa is regular, we have by a standard argument that there is a club CκC\subseteq\kappa such that for every αC\alpha\in C, every β<α\beta<\alpha, and every MβM_{\beta}-formula ψ\psi, there is a γ<α\gamma<\alpha such that ψ\psi is covered by γ\gamma.

Fix some μC\mu\in C. Let Aμ=α<μMαA_{\mu}=\bigcup_{\alpha<\mu}M_{\alpha}. We can now repeat the proofs of the claims in the proof of 1. In particular, we get from Claim 2 that {Xα(y):y<μ}\{X_{\alpha}(y):y<\mu\} axiomatizes a complete type. Therefore we added an element dμd_{\mu} to MμM_{\mu} realizing this type. From Claims 3 and 4, we get that there is an AμA_{\mu}-heir-coheir pμ(y)p_{\mu}(y) extending tp(dμ/Aμ)\operatorname{tp}(d_{\mu}/A_{\mu}) such that φ(x,y)\varphi(x,y) does not pp-Kim-divide.

Let (μ(ε))ε<κ(\mu(\varepsilon))_{\varepsilon<\kappa} be an enumeration of CC in order. Let Nε=Aμ(ε)N_{\varepsilon}=A_{\mu(\varepsilon)}, eε=dμ(ε)e_{\varepsilon}=d_{\mu(\varepsilon)}, and sε=pμ(ε)s_{\varepsilon}=p_{\mu(\varepsilon)} for each ε<κ\varepsilon<\kappa. Since (eε)ε<κ(e_{\varepsilon})_{\varepsilon<\kappa} is an \trianglelefteq-increasing sequence, we have that {φ(x,eε):ε<κ}\{\varphi(x,e_{\varepsilon}):\varepsilon<\kappa\} is kk-inconsistent, so we are done. ∎

For the following recall that on a measurable cardinal κ\kappa, we can find a normal ultrafilter 𝒰\mathcal{U} (i.e., for any sequence (Xα)α<κ(X_{\alpha})_{\alpha<\kappa} of elements of 𝒰\mathcal{U}, we have that the diagonal intersection Δα<κXα{β<κ:βα<βXα}\Delta_{\alpha<\kappa}X_{\alpha}\coloneqq\{\beta<\kappa:\beta\in\bigcap_{\alpha<\beta}X_{\alpha}\} is an element of 𝒰\mathcal{U}). In particular 𝒰\mathcal{U} is κ\kappa-complete and also has the property that every X𝒰X\in\mathcal{U} is stationary, implying that every club in κ\kappa is in 𝒰\mathcal{U}.

Lemma 3.2.

Let κ>|T|\kappa>|T| be a measurable cardinal. Let MM be a model of TT with |M|=κ|M|=\kappa. Let (Nα)α<κ(N_{\alpha})_{\alpha<\kappa} be a continuous chain of elementary substructures of MM with |Nα|<κ|N_{\alpha}|<\kappa for each α<κ\alpha<\kappa and α<κNα=M\bigcup_{\alpha<\kappa}N_{\alpha}=M. Let (cα)α<κ(c_{\alpha})_{\alpha<\kappa} be any sequence of elements of MM.

For any normal ultrafilter 𝒰\mathcal{U} on κ\kappa, there is an X𝒰X\in\mathcal{U} such that for every αX\alpha\in X, tp(cα/Nα)\operatorname{tp}(c_{\alpha}/N_{\alpha}) is finitely satisfiable in {cβ:β<α}\{c_{\beta}:\beta<\alpha\} and {cβ:β<α}Nα\{c_{\beta}:\beta<\alpha\}\subseteq N_{\alpha}.

Proof.

Let p(y)={ψ(x,b):{α<κ:ψ(cα,b)}𝒰}Sy(M)p(y)=\{\psi(x,b):\{\alpha<\kappa:\psi(c_{\alpha},b)\}\in\mathcal{U}\}\in S_{y}(M). By a standard argument, there is a club CκC\subseteq\kappa such that for any αC\alpha\in C, {cβ:β<α}Nα\{c_{\beta}:\beta<\alpha\}\subseteq N_{\alpha}. Note that since CC is a club, C𝒰C\in\mathcal{U}.

For each α\alpha, let pα=pNαp_{\alpha}=p\operatorname{\upharpoonright}N_{\alpha}. By construction we have that for each ψ(y)pα\psi(y)\in p_{\alpha}, {β<κ:ψ(cβ)}𝒰\{\beta<\kappa:\psi(c_{\beta})\}\in\mathcal{U}. Since |Nα|<κ|N_{\alpha}|<\kappa and since 𝒰\mathcal{U} is κ\kappa-complete, this implies that {β<κ:cβpα}𝒰\{\beta<\kappa:c_{\beta}\models p_{\alpha}\}\in\mathcal{U}.

Since 𝒰\mathcal{U} is normal, we have that XClimκΔα<κ{β<κ:cβpα}𝒰X\coloneqq C\cap\lim\kappa\cap\Delta_{\alpha<\kappa}\{\beta<\kappa:c_{\beta}\models p_{\alpha}\}\in\mathcal{U}. Consider βX\beta\in X. We have that for every α<β\alpha<\beta, cβpαc_{\beta}\models p_{\alpha} and pαp_{\alpha} is finitely satisfiable in {cγ:γ<α}Nγ\{c_{\gamma}:\gamma<\alpha\}\subseteq N_{\gamma}. Since β\beta is a limit ordinal, this implies that tp(cβ/Nβ)\operatorname{tp}(c_{\beta}/N_{\beta}) is finitely satisfiable in {cα:α<β}Nβ\{c_{\alpha}:\alpha<\beta\}\subseteq N_{\beta}. Since we can do this for any βX\beta\in X, we are done. ∎

Definition 3.3 ([8, Def. 5.3]).

A set Γ(x)\Gamma(x) of formulas is a dual type if there is some k<ωk<\omega such that Γ(x)\Gamma(x) is kk-inconsistent.

Proposition 3.4.

Let κ>|T|\kappa>|T| be a measurable cardinal. Let MM be the unique saturated model of TT of cardinality κ\kappa. Suppose that there is a dual type Γ(x)\Gamma(x) over MM with |Γ(x)|=κ|\Gamma(x)|=\kappa and a club CC of small elementary substructures of MM such that for any NCN\in C, there is a φN(x,cN)Γ(x)\varphi_{N}(x,c_{N})\in\Gamma(x) with cNNc_{N}\notin N such that for some NN-invariant pN(y)tp(c/N)p_{N}(y)\supset\operatorname{tp}(c/N), φ(x,c)\varphi(x,c) does not pp-Kim-divide.

  1. (1)

    If pN(y)p_{N}(y) is NN-bi-invariant for each NCN\in C, then TT has CTP.

  2. (2)

    If pN(y)p_{N}(y) is strongly NN-bi-invariant for each NCN\in C, then TT has ATP.

Proof.

The proofs of 1 and 2 are nearly identical. We will write the proof of 1. To get the proof of 2, just insert the word ‘strongly’ in the appropriate places.

By a standard argument we may assume that there is a continuous chain (Nα)α<κ(N_{\alpha})_{\alpha<\kappa} of small elementary substructures of MM such that M=α<κNαM=\bigcup_{\alpha<\kappa}N_{\alpha} and C={Nα:α<κ}C=\{N_{\alpha}:\alpha<\kappa\}. Let 𝒰\mathcal{U} be a normal ultrafilter on κ\kappa. For each α<κ\alpha<\kappa, let φα(x,cα)Γ(x)\varphi_{\alpha}(x,c_{\alpha})\in\Gamma(x) be a formula with cαNαc_{\alpha}\notin N_{\alpha} such that for some NαN_{\alpha}-bi-invariant pα(y)tp(cα/Nα)p_{\alpha}(y)\supseteq\operatorname{tp}(c_{\alpha}/N_{\alpha}), φα(x,cα)\varphi_{\alpha}(x,c_{\alpha}) does not pαp_{\alpha}-Kim-divide.

By κ\kappa-completeness, there is a Y𝒰Y\in\mathcal{U} and a formula φ(x,y)\varphi(x,y) such that for every αY\alpha\in Y, φα=φ\varphi_{\alpha}=\varphi. By 3.2, there is an X𝒰X\in\mathcal{U} such that for every αX\alpha\in X, tp(cα/Nα)\operatorname{tp}(c_{\alpha}/N_{\alpha}) is finitely satisfiable in {cβ:β<α}\{c_{\beta}:\beta<\alpha\} and {cβ:β<α}Nα\{c_{\beta}:\beta<\alpha\}\subseteq N_{\alpha}.

For any αXY\alpha\in X\cap Y, we now have that for any global coheir qq finitely satisfiable in {cβ:β<α}Nα\{c_{\beta}:\beta<\alpha\}\subseteq N_{\alpha}, if (ei)i<ω(e_{i})_{i<\omega} is a Morley sequence generated by qq, then {φ(x,ei):i<ω}\{\varphi(x,e_{i}):i<\omega\} is kk-inconsistent for some k<ωk<\omega. On the other hand, we have that φ(x,e0)\varphi(x,e_{0}) does not pαp_{\alpha}-Kim-divide where pαp_{\alpha} is an NαN_{\alpha}-bi-invariant type satisfying pαNα=qNαp_{\alpha}\operatorname{\upharpoonright}N_{\alpha}=q\operatorname{\upharpoonright}N_{\alpha}. Therefore TT has CTP. ∎

Corollary 3.5.

Fix a complete theory TT. If there is a measurable cardinal κ>|T|\kappa>|T|, then the following are equivalent.

  1. (1)

    TT has CTP.

  2. (2)

    There is a model MTM\models T with |M|=κ|M|=\kappa, a dual type Γ(x)\Gamma(x) over MM, and a club CC of small elementary substructures of MM such that for any NCN\in C, there is a φ(x,c)Γ(x)\varphi(x,c)\in\Gamma(x) with cNc\notin N such that for some NN-bi-invariant p(y)tp(c/N)p(y)\supseteq\operatorname{tp}(c/N), φ(x,c)\varphi(x,c) does not pp-Kim-divide.

  3. (3)

    The same as 2, but with each p(y)p(y) an NN-heir-coheir.

It seems likely that 3.5 does not require a large cardinal, but nevertheless the question needs to be asked.

Question 3.6.

Does 3.5 hold without the existence of a large cardinal?

Finally, we can essentially ask 1.6 again.

Question 3.7.

Does 3.5 hold for ATP if the types p(y)p(y) are assumed to be strongly bi-invariant?

3.2. NATP implies generic stationary local character

Simplicity, NTP2, and NSOP2 all have characterizations in terms of some kind of local character (which we will state in sub-optimal forms for the sake of exposition):

  • TT is simple if and only if it satisfies local character: There is a κ\kappa such that for every global type p(x)p(x), there is an MM with |M|κ|M|\leq\kappa such that p(x)p(x) does not divide over MM.

  • TT is NTP2 if and only if it satisfies generic local character: There is a κ\kappa such that for every global type p(x)p(x) and every MM with |M|κ|M|\leq\kappa, there is an NMN\succeq M with |N|κ|N|\leq\kappa such that for any dd, if dMiNd\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{i}}_{M}N, then pNdp\operatorname{\upharpoonright}Nd does not divide over NN [4].

  • TT is NSOP1 if and only if it satisfies stationary local character: There is a κ\kappa such that for any global type p(x)p(x), the set {M𝕆:|M|κ,p(x)does not Kim-divide overM}\{M\preceq\mathbb{O}:|M|\leq\kappa,~{}p(x)~{}\text{does not Kim-divide over}~{}M\} is stationary in the set of models of size less than κ\kappa [8].

Note that in all three of these, (Kim-)dividing coincides with (Kim-)forking.

Given the existence of these characterizations, when Kruckman visited the logic group at the University of Maryland in 2020, we tried to come up with a reasonable mutual generalization of generic local character and stationary local character. The idea being that this would be a nice complimentary approach to trying to mutually generalize NTP2 and NSOP1. We came to the following definition.

Definition 3.8.

For any type p(x)p(x) and any small M,NTM,N\models T with MNM\preceq N, we write Ξ(p,M,N)\Xi(p,M,N) for the following condition:

  • For any MM-formula φ(x,y)\varphi(x,y) and any dd such that φ(x,d)p(x)\varphi(x,d)\in p(x) and dMiNd\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{i}}_{M}N, φ(x,d)\varphi(x,d) does not Kim-divide over NN.

We say that TT satisfies generic stationary local character888Perhaps stationarily generic local character would be a more correct name, but it doesn’t quite roll off the tongue. It also doesn’t generalize well to the club version discussed in 3.14. if for every λ\lambda, there is a κ\kappa such that for every κ+\kappa^{+}-saturated model OO, every type pS(O)p\in S(O), and every MOM\preceq O with |M|λ|M|\leq\lambda,

{NO:NM,|N|κ,Ξ(p,M,N)}\{N\preceq O:N\succeq M,~{}|N|\leq\kappa,~{}\Xi(p,M,N)\}

is stationary in [O]κ{XO:|X|=κ}[O]^{\kappa}\coloneqq\{X\subseteq O:|X|=\kappa\}.

As it turns out, however, generic stationary local character is probably too strong to characterize NBTP in that its failure actually implies ATP.

Lemma 3.9.

Fix κ|T|\kappa\geq|T|. For any κ+\kappa^{+}-saturated model OO, any MOM\preceq O with |M|κ|M|\leq\kappa, and any MM-invariant type p(x)p(x), the set {N[O]κ:NT,pis an N-heir}\{N\subseteq[O]^{\kappa}:N\models T,~{}p~{}\text{is an }N\text{-heir}\} is a club in [O]κ[O]^{\kappa}.

Proof.

The set in question is clearly closed, so we just need to show that it is unbounded. Fix M0OM_{0}\preceq O with |M0|=κ|M_{0}|=\kappa. Given MnM_{n}, we can find a model Mn+1M_{n+1} with |Mn+1|κ|M_{n+1}|\leq\kappa such that for any MnM_{n}-formula φ(x,y)\varphi(x,y), if there is a bb in the monster such that φ(x,b)p(x¯)\varphi(x,b)\in p(\bar{x}), then there is a cMn+1c\in M_{n+1} such that φ(x¯,c)p\varphi(\bar{x},c)\in p. Let N=n<ωMnN=\bigcup_{n<\omega}M_{n}. We have that pp is an heir over NN. ∎

Proposition 3.10.

Suppose that TT does not satisfy generic stationary local character. Then there is a small model NN, a formula φ(x,d)\varphi(x,d), a strongly NN-bi-invariant type p(y)tp(d/M)p(y)\supset\operatorname{tp}(d/M), and an NN-invariant q(y)tp(d/M)q(y)\supset\operatorname{tp}(d/M) such that φ(x,d)\varphi(x,d) qq-Kim-divides but does not pp-Kim-divide.

Proof.

Let λ\lambda witness that TT fails generic stationary local character. Let κ=22λ\kappa=2^{2^{\lambda}}. Note that there are at most κ\kappa MM-invariant types for any MM with |M|λ|M|\leq\lambda.

There is a κ+\kappa^{+}-saturated model OO, a type p(x)S(O)p(x)\in S(O), and an MOM\preceq O with |M|λ|M|\leq\lambda such that {NO:NM,|N|κ,Ξ(p,M,N)}\{N\preceq O:N\succeq M,~{}|N|\leq\kappa,~{}\Xi(p,M,N)\} is not stationary in [O]κ[O]^{\kappa}. This means that there is a club C[O]κC\subseteq[O]^{\kappa} such that for every model NCN\in C with NMN\succeq M, there is a φ(x,d)p(x)\varphi(x,d)\in p(x) such that dMiNd\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{i}}_{M}N and φ(x,d)\varphi(x,d) Kim-divides over NN. By induction, we can build a continuous chain (Ni)i<κ+(N_{i})_{i<\kappa^{+}} of elementary substructures of OO and a sequence (φi(x,di))i<κ+(\varphi_{i}(x,d_{i}))_{i<\kappa^{+}} of formulas in p(x)p(x) such that for each i<κ+i<\kappa^{+}, diMiNid_{i}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{i}}_{M}N_{i} and φi(x,di)\varphi_{i}(x,d_{i}) Kim-divides over NiN_{i}. For each ii, let qi(y)tp(di/Ni)q_{i}(y)\supseteq\operatorname{tp}(d_{i}/N_{i}) be an MM-invariant type.

For each j<κ+j<\kappa^{+}, let Dj={i<κ+:i>j,(n<ω)qjnis anNi-heir}D_{j}=\{i<\kappa^{+}:i>j,~{}(\forall n<\omega)q_{j}^{\otimes n}~{}\text{is an}~{}N_{i}\text{-heir}\}. By 3.9, each DjD_{j} is a club in κ+\kappa^{+}. Therefore the diagonal intersection Δj<κ+Dj={i<κ+:(j<i)iDj}\Delta_{j<\kappa^{+}}D_{j}=\{i<\kappa^{+}:(\forall j<i)i\in D_{j}\} and also Elimκ+Δj<κ+DjE\coloneqq\lim\kappa^{+}\cap\Delta_{j<\kappa^{+}}D_{j} are clubs. For each iEi\in E, let f(i)f(i) be the least jj such that diNjd_{i}\in N_{j}. This is a regressive function, so by Fodor’s lemma, there is a k<κ+k<\kappa^{+} and a stationary set SES\subseteq E such that f(i)=kf(i)=k for all iSi\in S. Since the number of MM-formulas is less than κ\kappa, there is a stationary set SSS^{\prime}\subseteq S such that for any ii and jj in SS^{\prime}, φi=φj\varphi_{i}=\varphi_{j}. Since the number of MM-invariant types is at most κ\kappa, there is a stationary set S′′SS^{\prime\prime}\subseteq S^{\prime} such that for any ii and jj in S′′S^{\prime\prime}, qi=qjq_{i}=q_{j}. Let j=min(S′′)j=\min(S^{\prime\prime}) and let =min(S′′{j})\ell=\min(S^{\prime\prime}\setminus\{j\}).

We now have that (di)iS′′{j}(d_{i})_{i\in S^{\prime\prime}\setminus\{j\}} is a Morley sequence generated by qq_{\ell} over NN_{\ell}. Since p(x)p(x) is consistent, we have that {φ(x,di):iS′′{j}}\{\varphi_{\ell}(x,d_{i}):i\in S^{\prime\prime}\setminus\{j\}\} is consistent as well. Furthermore, qj=qq_{j}=q_{\ell} has the property that qjnq_{j}^{\otimes n} is an NjN_{j}-heir for each n<ωn<\omega. Therefore qjq_{j} is strongly NN_{\ell}-bi-invariant. Finally, we have that φ(x,d)\varphi_{\ell}(x,d_{\ell}) Kim-divides over NN_{\ell}. Therefore TT has ATP by 1.7. ∎

Corollary 3.11.

If TT is NATP, then it satisfies generic stationary local character.

Proof.

This follows immediately from Propositions 1.7 and 3.10. ∎

In our opinion the lesson of 3.10 and 3.11 is that it is unlikely that NCTP will be characterized by something like 3.8. It is difficult to imagine how to tune the forbidden configuration so that it will build bi-invariant types but not also strongly bi-invariant types.999Naturally, this would be a non-issue if it does turn out that CTP and ATP are equivalent.

Broadly speaking, Propositions 0.6 and 1.5 would seem to indicate that bi-invariance is expected generically at the level of formulas, whereas 3.10 indicates that strong bi-invariance is expected generically at the level of partial types.

Of course, 3.11 is unsatisfactory in a couple of different ways. First of all, it is not a characterization.

Question 3.12.

If TT satisfies generic stationary local character, does it follow that TT is NATP?

But also the local character characterizations of simplicity, NTP2, and NSOP1 have tighter cardinal bounds than we have stated. Our proof of 3.10 shows that if generic stationary local character fails at λ\lambda, then we can build an instance of ATP from any (22λ)+(2^{2^{\lambda}})^{+}-saturated m del. For all three of the aforementioned conditions, it is known that this is witnessed by any |T|+|T|^{+}-saturated model. The proofs of these tighter bounds rely on more detailed structural understandings of the relevant class of tame theories, and in particular that the tameness condition is characterized by the associated notion of local character.

Question 3.13.

If TT fails to have generic stationary local character, is this witnessed by models that are |T|+|T|^{+}-saturated?

For NTP2 in particular, the original statement of generic local character in [4] did not require the big model to have any degree of saturation but instead required that each tp(d/N)\operatorname{tp}(d/N) extends to a strictly invariant type. We could make a similar statement here, more in the vein of 3.4, and drop the requirement that the big model be saturated but require that each tp(d/N)\operatorname{tp}(d/N) extends to a strongly bi-invariant type. We did not opt to highlight this version of 3.10, however, as the interesting thing about the proposition is the fact that we do not need to assume that some configuration of (strongly) bi-invariant types happens to exist, as we do in 3.4.

Finally, in the case of NSOP1, we have done the relevant results in [8] a bit of a disservice. In [8], Kaplan, Ramsey, and Shelah actually prove that in NSOP1 theories, for any global type p(x)p(x), the set of |T||T|-sized models M𝕆M\prec\mathbb{O} over which p(x)p(x) does not Kim-divide is a club. This is an instance of the kind of dichotomous behavior you expect from a dividing line: Either every type is ‘good’ on a club of small models or there is a type that is ‘bad’ on a club of small models. Again, however, the proof of this relies heavily on a structural understanding of NSOP1 theories, so it is entirely unclear if something like this would generalize to the present context, which leaves an obvious question.

Question 3.14.

Say that a theory has generic club local character if for any global type p(x)p(x) and any M𝕆M\prec\mathbb{O} with |M||T||M|\leq|T|, the set {N𝕆:NM,|N||T|,Ξ(p,M,N)}\{N\prec\mathbb{O}:N\succeq M,~{}|N|\leq|T|,~{}\Xi(p,M,N)\} is a club in [𝕆]|T|[\mathbb{O}]^{|T|}. Is generic stationary local character equivalent to generic club local character?

Appendix A Dual results for NSOP1

In [11, Lem. 2.8], Mutchnik shows that SOP2 is equivalent to the following condition:

  • There is a k<ωk<\omega, a tree (bσ)σω<ω(b_{\sigma})_{\sigma\in\omega^{<\omega}}, and a formula φ(x,y)\varphi(x,y) such that for any path αωω\alpha\in\omega^{\omega}, {φ(x,bαn):n<ω}\{\varphi(x,b_{\alpha\operatorname{\upharpoonright}n}):n<\omega\} is consistent but for any right-comb Cω<ωC\subset\omega^{<\omega}, {φ(x,bσ):σC}\{\varphi(x,b_{\sigma}):\sigma\in C\} is kk-inconsistent.

Furthermore, by the main result of [11], this is equivalent to SOP1. In [9], Kim and Lee show that the above condition is equivalent to the above condition with a tree indexed by 2<ω2^{<\omega}. Essentially all of the proofs given in this paper are insensitive to this duality, so we can freely conclude several new results for NSOP1 theories. (Some of our results, such as 1.7, are trivial when dualized, however.)

Proposition A.1 (Dual of 1.5 and 3.1 part 1).

If TT has SOP1, then there is a model MM, a formula φ(x,y)\varphi(x,y), an MM-coheir p(y)p(y), and an MM-heir-coheir q(y)q(y) such that pM=qMp\operatorname{\upharpoonright}M=q\operatorname{\upharpoonright}M and φ(x,y)\varphi(x,y) qq-Kim-divides but does not pp-Kim-divide.

We also no longer need the measurable cardinal in the following result, as the hard part has already been done for us in [8].

Proposition A.2 (Dual of 3.5).

Fix a complete theory TT. The following are equivalent.

  1. (1)

    TT has SOP1.

  2. (2)

    There is a model MTM\models T with |M|=κ|M|=\kappa, a type p(x)p(x) over MM, and a club CC of small elementary substructures of MM such that for any NCN\in C, there is a φ(x,c)p(x)\varphi(x,c)\in p(x) with cNc\notin N such that for some NN-bi-invariant p(y)tp(c/N)p(y)\supseteq\operatorname{tp}(c/N), φ(x,c)\varphi(x,c) pp-Kim-divides.

  3. (3)

    The same as 2, but with each p(y)p(y) an NN-heir-coheir.

Proof.

The fact that 1 implies 3 follows from the dual of 3.1 part 2. The fact that 3 implies 2 is obvious. Finally, the fact that 2 implies 1 follows from [8, Thm. 1.1]. ∎

Appendix B Characterization of coheirs in terms of invariant extensions in expansions

Here we will give the proof that ¬\neg2b implies ¬\neg2a in 2.1. Recall the following: BAaCB\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{a}}_{A}C means that acl(AB)acl(AC)=acl(C)\operatorname{acl}(AB)\cap\operatorname{acl}(AC)=\operatorname{acl}(C). a\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{a}} satisfies all of the axioms of a strict independence relation except for possibly base monotonicity (see [1, Sec. 1]101010Although note that there are some errors in this source. See [6] for a full account of the relevant results with correct proofs.). In particular, this means that a\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{a}} satisfies full existence: For any AA, BB, and CC, there is a CACC^{\prime}\equiv_{A}C such that BAaCB\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{a}}_{A}C^{\prime}.

First we will need a lemma.

Lemma B.1.

Let AA be a set of parameters satisfying A=acl(A)A=\operatorname{acl}(A). Let PP, UU, and UU^{\ast} be fresh unary predicate symbols. Let 0={P}\mathcal{L}_{0}=\mathcal{L}\cup\{P\} and let 1=0{U,U}\mathcal{L}_{1}=\mathcal{L}_{0}\cup\{U,U^{\ast}\}. For any 1\mathcal{L}_{1}-formula φ(x¯)\varphi(\bar{x}), let φ(x¯)\varphi^{\ast}(\bar{x}) be the result of replacing each instance of UU with UU^{\ast} and each instance of UU^{\ast} with UU in φ(x¯)\varphi(\bar{x}).

Let κ=|A|+||\kappa=|A|+|\mathcal{L}|. For any κ+\kappa^{+}-saturated, κ+\kappa^{+}-homogeneous111111By κ+\kappa^{+}-homogeneous we mean that for any tuples b¯\bar{b} and c¯\bar{c} realizing the same type with |b¯|=|c¯|κ|\bar{b}|=|\bar{c}|\leq\kappa, there is an automorphism taking b¯\bar{b} to c¯\bar{c}. model MAM\supseteq A, there is an 1\mathcal{L}_{1}-structure (N,PN,UN,UN)(N,P^{N},U^{N},U_{\ast}^{N}) such that

  • (N,PN)0(M,A)(N,P^{N})\succeq_{\mathcal{L}_{0}}(M,A),

  • {PN,UN,UN}\{P^{N},U^{N},U_{\ast}^{N}\} forms a partition of NN, and

  • for any 1(A)\mathcal{L}_{1}(A)-formula φ(x¯)\varphi(\bar{x}), N(x¯P)(φ(x¯)φ(x¯))N\models(\forall\bar{x}\in P)(\varphi(\bar{x})\leftrightarrow\varphi^{\ast}(\bar{x})).

Proof.

Fix AA and MM as in the statement of the lemma. Note that since MM is κ+\kappa^{+}-saturated as a model of TT, we have that for any tuples b¯\bar{b} and c¯\bar{c} with tp(b¯/A)=tp(c¯/A)\operatorname{tp}_{\mathcal{L}}(\bar{b}/A)=\operatorname{tp}_{\mathcal{L}}(\bar{c}/A) and |b¯|=|c¯|κ|\bar{b}|=|\bar{c}|\leq\kappa, b¯\bar{b} and c¯\bar{c} realize the same {P}\mathcal{L}\cup\{P\}-type over AA as well (by a back-and-forth argument).

Let T1T_{1} be the theory consisting of the elementary diagram of (M,PM)(M,P^{M}), an axiom asserting that {P,U,U}\{P,U,U^{\ast}\} is a partition of the universe, ¬x(P(x)ψ(x,e))\neg\exists x(P(x)\wedge\psi(x,e)), and (x¯P)(φ(x¯)φ(x¯))(\forall\bar{x}\in P)(\varphi(\bar{x})\leftrightarrow\varphi^{\ast}(\bar{x})) for each 1\mathcal{L}_{1}-formula φ\varphi.

Clearly we just need to show that T1T_{1} is consistent. We will prove this by building an expansion of MM in a forcing extension that is a model of T1T_{1}. By absoluteness, this will imply that each finite subset of T1T_{1} is consistent and so T1T_{1} itself is consistent.

Let \mathbb{P} be a forcing poset whose conditions are pairs of the form (B,C)(B,C), where B,CMAB,C\subseteq M\setminus A, |B|κ|B|\leq\kappa, |C|κ|C|\leq\kappa, and BC=B\cap C=\varnothing. The ordering is given by extension, i.e., (B,C)(B,C)(B^{\prime},C^{\prime})\leq(B,C) if and only if BBB^{\prime}\supseteq B and CCC^{\prime}\supseteq C. Let GG be a generic filter for this poset and consider the forcing extension V[G]V[G]. Let UM={B:(B,C)G}U^{M}=\bigcup\{B:(B,C)\in G\} and UM={C:(B,C)G}U^{M}_{\ast}=\bigcup\{C:(B,C)\in G\}. It is clear that {PM,UM,UM}\{P^{M},U^{M},U_{\ast}^{M}\} forms a partition of MM. Let M1M_{1} be the 1\mathcal{L}_{1}-structure (M,PM,UM,UM)(M,P^{M},U^{M},U_{\ast}^{M}). We just need to show that for each 1\mathcal{L}_{1}-sentence φ\varphi, M1φφM_{1}\models\varphi\leftrightarrow\varphi^{\ast}. Fix an 1\mathcal{L}_{1}-formula φ(x¯)\varphi(\bar{x}) and a tuple a¯A\bar{a}\in A and suppose that M1φ(a¯)M_{1}\models\varphi(\bar{a}). By the truth lemma, there is a forcing condition (B,C)G(B,C)\in G such that (B,C)M1φ(a¯)(B,C)\Vdash\text{``}M_{1}\models\varphi(\bar{a})\text{''}. By full existence, we have that for any condition (B,C)(B,C)(B^{\prime},C^{\prime})\leq(B,C), there is a B′′C′′B^{\prime\prime}C^{\prime\prime} (in the monster in VV) such that BCAB′′C′′BC\equiv_{A}B^{\prime\prime}C^{\prime\prime} and BCAaB′′C′′B^{\prime}C^{\prime}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}^{\!\!\textnormal{a}}_{A}B^{\prime\prime}C^{\prime\prime} (i.e., acl(ABC)acl(AB′′C′′)=acl(A)=A\operatorname{acl}(AB^{\prime}C^{\prime})\cap\operatorname{acl}(AB^{\prime\prime}C^{\prime\prime})=\operatorname{acl}(A)=A). In particular, BCB′′C′′=B^{\prime}C^{\prime}\cap B^{\prime\prime}C^{\prime\prime}=\varnothing and so BC′′CB′′=B^{\prime}C^{\prime\prime}\cap C^{\prime}B^{\prime\prime}=\varnothing. Since MM is κ+\kappa^{+}-saturated in VV, we may assume that B′′C′′MB^{\prime\prime}C^{\prime\prime}\subseteq M. Since the forcing poset is invariant under Aut(M/A)\operatorname{Aut}(M/A), we have that (C′′,B′′)M1φ(a¯)(C^{\prime\prime},B^{\prime\prime})\Vdash\text{``}M_{1}\models\varphi^{\ast}(\bar{a})\text{''} and so (BC′′,CB′′)(B^{\prime}C^{\prime\prime},C^{\prime}B^{\prime\prime}) forces this as well. Since we can do this for any (B,C)(B,C)(B^{\prime},C^{\prime})\leq(B,C), we have that (B,C)M1φ(B,C)\Vdash\text{``}M_{1}\models\varphi^{\ast}\text{''}. Therefore M1φ(a¯)φ(a¯)M_{1}\models\varphi(\bar{a})\leftrightarrow\varphi^{\ast}(\bar{a}). Since this is true for any a¯P\bar{a}\in P, we have that M1(x¯P)(φ(x¯)φ(x¯))M_{1}\models(\forall\bar{x}\in P)(\varphi(\bar{x})\leftrightarrow\varphi^{\ast}(\bar{x})). Since φ(x¯)\varphi(\bar{x}) was arbitrary, M1M_{1} satisfies the required axiom schema.

Finally, since ψ(x,e)\psi(x,e) is not satisfied in A=PMA=P^{M}, we have that M1¬x(P(x)ψ(x,e))M_{1}\models\neg\exists x(P(x)\wedge\psi(x,e)). ∎

Now we can complete the proof of 2.1.

Proof that ¬\neg2b implies ¬\neg2a in 2.1.

Fix a set of parameters A=acl(A)A=\operatorname{acl}(A) and let p(x)p(x) be an AA-invariant type that is not finitely satisfiable in AA. Fix an \mathcal{L}-formula ψ(x,y)\psi(x,y) and a parameter ee such that ψ(x,e)p(x)\psi(x,e)\in p(x) and ψ(x,e)\psi(x,e) is not satisfiable in AA. Let κ=(|A|+||)+\kappa=(|A|+|\mathcal{L}|)^{+}. Let MAM\supseteq A be a κ+\kappa^{+}-saturated, κ+\kappa^{+}-homogeneous model of TT. By κ+\kappa^{+}-saturation we may assume that the parameter ee is in MM. Let PP be a new unary predicate. Expand MM with PP so that PM=AP^{M}=A.

Let NN be the extension of MM guaranteed by B.1. By a standard model-theoretic argument, we may assume (by passing to a sufficiently saturated and homogeneous elementary extension if necessary) that there is an automorphism fAut(N/A)f\in\operatorname{Aut}(N/A) (of NN as an \mathcal{L}-structure) such that f[UN]=UNf[U^{N}]=U_{\ast}^{N}. (Note that ff does not necessary fix all of PNP^{N}.) Now let EE be the equivalence relation on NN with equivalence classes {PN,UN,UN}\{P^{N},U^{N},U_{\ast}^{N}\}. Consider the expansion (N,E)(N,E). Note that ff is still an automorphism of this structure. Note also that (N,E)¬x(x𝐸aψ(x,e))(N,E)\models\neg\exists x(x\mathrel{E}a\wedge\psi(x,e)) for any aAa\in A. Let q(x)q(x) be an Aut((N,E)/A)\operatorname{Aut}((N,E)/A)-invariant extension of pNp\operatorname{\upharpoonright}N. Since ψ(x,e)q(x)\psi(x,e)\in q(x), we must have that q(x)q(x) concentrates on one of the EE-equivalence classes other than the class containing AA, but these classes are not fixed by ff, so we have a contradiction. Therefore no such extension can exist. ∎

One thing to note about this proof is that in the expansion, the set AA is not algebraically closed in T1eqT_{1}^{\mathrm{eq}}. This raises a question.

Question B.2.

Which invariant types p(x)p(x) satisfy the following property?

  • For any model MAM\supseteq A and any expansion MM^{\dagger} of MM, pMp\operatorname{\upharpoonright}M has a completion in S(M)S(M^{\dagger}) that is Autf(M/A)\operatorname{Autf}(M^{\dagger}/A)-invariant.

Appendix C An example of Kim-dividing only along non-extendibly invariant types

Let \mathcal{L} be a language with three sorts: GG (for graph), OO (for orders), and PP (for points). GG has a unary relation UU and a binary relation RR. We have four unary functions, which we will think of as two unary functions with codomain G2G^{2}: fO,gO:OG2\langle f_{O},g_{O}\rangle:O\to G^{2} and fP,gP:PG2\langle f_{P},g_{P}\rangle:P\to G^{2}. Given xx and yy in GG, we’ll write Ox,yO_{x,y} for the fO,gO\langle f_{O},g_{O}\rangle-preimage of x,y\langle x,y\rangle and we’ll write Px,yP_{x,y} for the fP,gP\langle f_{P},g_{P}\rangle-preimage of x,y\langle x,y\rangle. Finally, we have a ternary relation << on O×P2O\times P^{2}, which we’ll write as a parameterized binary relation y<xzy<_{x}z for xOx\in O.

Let Q(x,y,z,w)Q(x,y,z,w) be the formula that says |{x,y,z,w}|=4|\{x,y,z,w\}|=4, {x,y,z,w}U\{x,y,z,w\}\subseteq U, and there is an RR-edge from some element of {x,y}\{x,y\} to some element of {z,w}\{z,w\}.

Let TT be the model companion of the universal theory that says

  • RR is a triangle-free graph relation,

  • U(fO(x))U(f_{O}(x)), U(gO(x))U(g_{O}(x)), U(fP(x))U(f_{P}(x)), and U(gP(x))U(g_{P}(x)) always hold,

  • if Q(x,y,z,w)Q(x,y,z,w), then for any Ox,y\ell\in O_{x,y}, <<_{\ell} is a linear order on Pz,wP_{z,w}, and

  • for any Ox,y\ell\in O_{x,y}, if u<vu<_{\ell}v for some uu and vv, then fP(u),gP(u)=fP(v),gP(v)\langle f_{P}(u),g_{P}(u)\rangle=\langle f_{P}(v),g_{P}(v)\rangle.

It is not hard but also not entirely pleasant to establish that this the finite models of this universal theory form a Fraïssé class with free amalgamation. Furthermore, TT is the theory of its Fraïssé limit and has quantifier elimination.

Fix a model MM of TT. Fix m¬U(M)m\in\neg U(M) and bb and cc outside of MM such that ¬(b𝑅c)\neg(b\mathrel{R}c), the only edge between MM and {b,c}\{b,c\} is c𝑅mc\mathrel{R}m, and U(b)U(b) and U(c)U(c) hold.

Lemma C.1.

The formula x𝑅bx𝑅cx\mathrel{R}b\wedge x\mathrel{R}c Kim-divides over MM.

Proof.

Let p(y,z)p(y,z) be the global MM-invariant type extending tp(bc/M)\operatorname{tp}(bc/M) entailing for all dd in the monster, y𝑅dy\mathrel{R}d if and only if dMd\notin M and d𝑅md\mathrel{R}m, and z𝑅dz\mathrel{R}d if and only if d=md=m. By quantifier elimination, this entails a complete MM-invariant.

If we find bcpMbcb^{\prime}c^{\prime}\models p\operatorname{\upharpoonright}Mbc, then we’ll have b𝑅cb^{\prime}\mathrel{R}c, implying that x𝑅bx𝑅cx𝑅bx𝑅cx\mathrel{R}b\wedge x\mathrel{R}c\wedge x\mathrel{R}b^{\prime}\wedge x\mathrel{R}c^{\prime} is inconsistent. ∎

Lemma C.2.

If x𝑅bx𝑅cx\mathrel{R}b\wedge x\mathrel{R}c Kim-divides with respect to an MM-invariant type p(y,z)tp(bc/M)p(y,z)\supseteq\operatorname{tp}(bc/M), then p(y,z)Q(y,z,b,c)p(y,z)\vdash Q(y,z,b,c).

Proof.

For any MM-invariant type p(y,z)tp(bc/M)p(y,z)\supseteq\operatorname{tp}(bc/M), p(y,z)|{y,z,b,c}|=4p(y,z)\vdash|\{y,z,b,c\}|=4. Furthermore, we necessarily have that p(y,z)U(y)U(z)p(y,z)\vdash U(y)\wedge U(z), so the only thing to check is that if bcpMbcb^{\prime}c^{\prime}\models p\operatorname{\upharpoonright}Mbc, then there is an RR-edge between some element of {b,c}\{b,c\} and some element of {b,c}\{b^{\prime},c^{\prime}\}.

Suppose that this doesn’t happen. Then if (bici)i<ω(b_{i}c_{i})_{i<\omega} is a Morley sequence generated by p(y,z)p(y,z), we’ll have that there are no RR-edges between any pair of elements of {bi,ci:i<ω}\{b_{i},c_{i}:i<\omega\}, implying that {x𝑅bix𝑅ci:i<ω}\{x\mathrel{R}b_{i}\wedge x\mathrel{R}c_{i}:i<\omega\} is consistent, contradicting the fact that this formula Kim-divides along Morley sequences generated by p(y,z)p(y,z). Therefore p(y,z)Q(y,z,b,c)p(y,z)\vdash Q(y,z,b,c). ∎

Let 𝕆\mathbb{O} be the monster model of TT.

Lemma C.3.

Pb,c(𝕆)P_{b,c}(\mathbb{O}) is an MM-indiscernible set.

Proof.

First note that Q(y,z,b,c)Q(y,z,b,c) does not hold for any {y,z}M\{y,z\}\subseteq M. Therefore for any O(M)\ell\in O(M), <<_{\ell} is trivial on Pb,c(𝕆)P_{b,c}(\mathbb{O}). By quantifier elimination, this implies that any two nn-tuples of distinct elements of Pb,c(𝕆)P_{b,c}(\mathbb{O}) have the same type over MM, so Pb,c(𝕆)P_{b,c}(\mathbb{O}) is an MM-indiscernible set. ∎

Let dd be an element of Ob,c(𝕆)O_{b,c}(\mathbb{O}).

Lemma C.4.

If p(y,z,w)tp(bcd/M)p(y,z,w)\supseteq\operatorname{tp}(bcd/M) is an MM-invariant type, then p(y,z,w)¬Q(y,z,b,c)p(y,z,w)\vdash\neg Q(y,z,b,c).

Proof.

Assume for the sake of contradiction that p(y,z,w)Q(y,z,b,c)p(y,z,w)\vdash Q(y,z,b,c). Define a relation <p<^{p} on P(𝕆)P(\mathbb{O}) by e<pfe<^{p}f if and only if p(y,z,w)e<wfp(y,z,w)\vdash e<_{w}f. Note that <p<^{p} is clearly MM-invariant. Since p(y,z,w)Q(y,z,b,c)p(y,z,w)\vdash Q(y,z,b,c), the restriction of <p<^{p} to Pb,c(𝕆)P_{b,c}(\mathbb{O}) needs to be a linear order, but there are no MM-invariant linear orders on Pb,c(𝕆)P_{b,c}(\mathbb{O}), since it is an MM-indiscernible set. ∎

Proposition C.5.

If x𝑅bx𝑅cx\mathrel{R}b\wedge x\mathrel{R}c Kim-divides along an MM-invariant type p(y,z)tp(bc/M)p(y,z)\supseteq\operatorname{tp}(bc/M), then p(y,z)tp(bcd/M)p(y,z)\cup\operatorname{tp}(bcd/M) has no MM-invariant completions.

One thing to note about this example is that it does in fact have ATP (since it interprets the Henson graph). It follows from [9, Thm. 3.10] that C.5 cannot occur in an NATP theory.

References

  • [1] Hans Adler. A geometric introduction to forking and thorn-forking. Journal of Mathematical Logic, 09(01):1–20, June 2009.
  • [2] JinHoo Ahn and Joonhee Kim. SOP1, SOP2, and antichain tree property, 2020.
  • [3] JinHoo Ahn, Joonhee Kim, and Junguk Lee. On the antichain tree property. Journal of Mathematical Logic, 23(02), December 2022.
  • [4] Artem Chernikov. Theories without the tree property of the second kind. Annals of Pure and Applied Logic, 165(2):695–723, February 2014.
  • [5] Artem Chernikov and Itay Kaplan. Forking and dividing in NTP2 theories. The Journal of Symbolic Logic, 77(1):1–20, 2012.
  • [6] Gabriel Conant and James Hanson. Separation for isometric group actions and hyperimaginary independence. Fundamenta Mathematicae, 259(1):97–109, 2022.
  • [7] Itay Kaplan and Nicholas Ramsey. On Kim-independence. Journal of the European Mathematical Society, 22(5):1423–1474, January 2020.
  • [8] Itay Kaplan, Nicholas Ramsey, and Saharon Shelah. Local character of Kim-independence. Proceedings of the American Mathematical Society, 147(4):1719–1732, January 2019.
  • [9] Joonhee Kim and Hyoyoon Lee. Some remarks on Kim-dividing in NATP theories, 2022.
  • [10] Alex Kruckman and Nicholas Ramsey. Kim’s Lemmas and tree properties. ASL Special Session on Model-theoretic Classification Program, II, 2022.
  • [11] Scott Mutchnik. On NSOP2 theories, 2022.