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

A walk on the wild side:
Notions of maximality in first-order theories

Michele Bailetti
(Date: September 28, 2024.)
Abstract.

In the classification of complete first-order theories, many dividing lines have been defined in order to understand the complexity and the behavior of some classes of theories. In this paper, using the concept of patterns of consistency and inconsistency, we describe a general framework to study dividing lines and we introduce a notion of maximal complexity by requesting the presence of all the exhibitable patterns of definable sets. Weakening this notion, we define new properties (Positive Maximality and the PMkPM_{k} hierarchy) and prove some results about them. In particular, we show that PMk+1PM_{k+1} theories are not kk-dependent.

1. Introduction

Classification theory is a line of research in model theory, started essentially by Shelah, in which we define and study dividing lines that differentiate classes of theories in order to understand their complexity and behavior. One of the first dividing lines, introduced by Shelah, is stability (See Definition 2.1), defined as a generalization of Morley’s notion of totally transcendental theories. Stable theories can be characterized by the absence of a local combinatorial property on definable sets (Order Property, See Definition 2.1, (1)). As shown by Shelah, the presence of this local property (the theory is said to be unstable in this case) implies bad structural behaviors of the theory. For instance

Theorem (Shelah,[19]).

If TT is unstable and λ>|T|+0\lambda>|T|+\aleph_{0}, then TT has exactly 2λ2^{\lambda} non-isomorphic models of size λ\lambda.

On the other hand, the assumption of stability provides methods and tools to analyze and understand this class of theories. The rich framework provided by stability led researchers to try to extend these methodologies outside stability, defining more and more dividing lines and giving structure to the universe of first-order theories.

Two prototypical examples of unstable theories, dense linear orders without endpoints and the theory of the random graph, led to the definition of two incomparable extensions of the class of stable theories whose intersection characterizes it: NIPNIP and NSOPNSOP (See Definition 2.1 and Fact 2.2). In order to study the existence of saturated elementary extensions of models of first-order theories , Shelah (in [20]) isolated a class of theories, extending stability, called simple theories (omitting the tree property, see Definition 2.1, (4)). Simple theories were then intensely studied by Hrushovski, Chatzidakis, Pillay, B. Kim et al (See, e.g., [8, 10, 13, 12, 2]), showing that, with the assumption of omitting the tree property, many typically stability-theoretic tools behave correctly and are useful in this enlarged framework. A plethora of dividing lines has been defined, along with great developments regarding the classes of theories that are characterized by them, making classification theory a rich and exciting line of research in model theory.

Many of these dividing lines are defined, similarly to stability, by the absence of specific configurations of definable sets. In this paper, abstracting the essential and combinatorial nature of these properties, we introduce the concept of patterns of consistency and inconsistency (See Definition 3.1) in order to give a general framework for talking about dividing lines. This concept was first defined by Shelah in [22] (Definition 5.17) with a different notation and more recently appeared in [14] (Definition 6.2) in a fashion similar to ours.

Patterns essentially describe abstract configurations of finite collections of sets by asserting conditions about the emptiness (or not) of their intersections. A formula exhibits a pattern in a theory if there are parameters in a model of the theory such that the sets defined by the formula and the parameters satisfy the conditions described by the pattern (See Definition 3.2). In this sense, we can capture the combinatorial properties that define dividing lines by exhibition of patterns. In Section 3, we define all the notions relative to patterns and describe this general framework for dividing lines.

On the wild side of every previously considered dividing line, two notions of maximal complexity for a complete first-order theory have been defined: Cooper’s property 11 (See [7]) and “an extreme condition of the form of unstability” called Straight Maximality (SMSM) defined by Shelah ([23]). In Section 4, we show that these two notions are equivalent and fit in the framework of exhibition of patterns. Indeed, these two properties are characterized by the existence of a formula exhibiting all the exhibitable (see Definition 3.2) patterns. We kept a name similar to Shelah’s for this property (Straight up Maximality).

One of the main approaches in classification theory, as we described above, is to extend nice properties and tools on the tame side of some dividing line to a wider class of theories. Here we reverse the approach and start from the worst (in some sense) combinatorial property that a complete first-order theory can have and study how theories behave when we weaken this bad property. Using this methodology, in Section 5 we define weakenings of SMSM by requesting the presence of a formula exhibiting a smaller collection of patterns, guided by the examples of previously defined properties of formulas. For instance, both OPOP and IPIP do not require any inconsistency conditions. We thus define Consistency Maximality (CMCM, see Definition 5.10) by requiring the existence of a formula realizing all the patterns without inconsistency conditions. We show that CMCM is equivalent to IPIP (even at the level of formulas), giving a different perspective to this well-known property. Other dividing lines, such as TPTP, TP1TP_{1} and TP2TP_{2} (See Definition 2.1, (4),(5),(6)), can be characterized by realizing patterns in which the negation of the formula is not involved. Using this observation, we define a collection of patterns (positive patterns, see Definition 5.1) and a weaker version of SMSM, called Positive Maximality (PMPM, see Definition 5.2) by requesting the existence of a formula exhibiting all the positive patterns. In particular, all the properties that can be characterized by realization of positive patterns are implied by PMPM. Since our characterization of SOPSOP uses patterns that are not positive, a natural question is the following:

Question 1.1.

Is there a positive maximal NSOPNSOP theory?

In Section 5, we answer positively to the above question by finding an example of a NSOP4NSOP_{4} (See Definition 2.1,(747_{4})) positive maximal theory (See Example 5.7). In particular, this shows that neither SOPSOP nor SOPnSOP_{n} for n4n\geq 4 (See Definition 2.1,(7n7_{n})) can be characterized through positive patterns. We point out that, in [9] (Lemma 7.3), Kaplan, Ramsey and Simon proved a characterization of SOP3SOP_{3} which can be obtained by exhibition of positive patterns.

In Section 6, by putting restraints on the size of the inconsistency conditions in positive patterns, we define a collection of properties (PMk)1<k<ω(PM_{k})_{1<k<\omega} (See Definition 6.1), weaker than PMPM but still strong enough to imply all the classical positive dividing lines. We show that, at the level of theories, PMk+1PM_{k+1} implies PMkPM_{k} (See Proposition 6.5) and we give NSOP4NSOP_{4}, PMkPM_{k} and NPMk+1NPM_{k+1} examples showing that the above implication is strict (See Example 6.13). We show that PMkPM_{k} is equivalent to the realization (in some sense) of every finite kk-hypergraph and can be witnessed by a (generic ordered kk-hypergraph)-indexed generalized indiscernible. The kk-independence property (IPkIP_{k}) is a higher-arity version of IPIP introduced by Shelah. In [3], Chernikov, Palacin and Takeuchi showed the equivalence of NIPkNIP_{k} with the collapse of the above kind of generalized indiscernibles. This led the way for showing the following result.

Theorem (See 6.11).

For every 0<k<ω0<k<\omega, if TT is PMk+1PM_{k+1}, then TT has IPkIP_{k}.

The notions of maximality introduced in this paper as essentially given by binary properties. The above result shows, in particular, that the binary notion of positive maximality implies all the kk-ary versions of IPIP (at each step of the way). We also point out that there are examples of simple IPkIP_{k} theories, and thus PMk+1PM_{k+1} is not equivalent to IPkIP_{k}.

Many questions arise about these notions of maximality, the behavior of classes of theories defined by them and their relations with previously defined dividing lines. For instance, which of the standard examples of SOP3SOP_{3} but NSOPNSOP theories are PMPM (or PMkPM_{k} for some 1<k<ω1<k<\omega)? Is there a SOP3SOP_{3} but NSOPNSOP and NPM2NPM_{2} theory? If not, we would have that NSOPNPM2=NSOP3NSOP\cap NPM_{2}=NSOP_{3}; this would imply the collapse of the SOPnSOP_{n} hierarchy inside NTP2NTP_{2}.

2. Preliminaries

Unless otherwise stated, in what follows TT is a complete first-order \mathcal{L}-theory, 𝒰T\mathcal{U}\models T is a monster model. Whenever we state the existence of some tuples, we mean tuples of elements from 𝒰\mathcal{U} and every set is a small (of cardinality less then the saturation of 𝒰\mathcal{U}) subset of 𝒰\mathcal{U}. Frequently we work with partitioned formulas φ(x;y)\varphi(x;y) in which the free variables are divided into object variables and parameters variables (the variables xx and yy respectively). Moreover, whenever we write x,y,z,x,y,z,... for variables we mean tuples of variables (not necessarily singletons) and, as a notation, we write 𝒰x,𝒰y,𝒰z,\mathcal{U}^{x},\mathcal{U}^{y},\mathcal{U}^{z},... for the set of tuples of elements of 𝒰\mathcal{U} that can be plugged in for x,y,z,x,y,z,... respectively.

2.1. Dividing lines

We recall the definition of some of the properties defining dividing lines in the universe of first-order theories. Most of these properties were introduced by Shelah in [18] and [21]. For the definitions and a bird’s-eye view of the universe we refer to [4].

Definition 2.1.

A partitioned formula φ(x;y)\varphi(x;y) has

  1. (1)

    the order property (OPOP) if there are tuples (aibi)i<ω(a_{i}b_{i})_{i<\omega} such that

    φ(ai;bj)i<j\models\varphi(a_{i};b_{j})\iff i<j
  2. (2)

    the independence property (IPIP) if there are tuples (ai)i<ω(a_{i})_{i<\omega} and (bI)Iω(b_{I})_{I\subseteq\omega} such that

    φ(ai;bI)iI\models\varphi(a_{i};b_{I})\iff i\in I
  3. (3)

    the strict-order property (SOPSOP) if there are tuples (bi)i<ω(b_{i})_{i<\omega} such that

    x(φ(x;bj)¬φ(x;bi))i<j\models\exists x(\varphi(x;b_{j})\land\neg\varphi(x;b_{i}))\iff i<j
  4. (4)

    the kk-tree property (kk-TPTP) if there exist tuples (bη)ηω<ω(b_{\eta})_{\eta\in\omega^{<\omega}} such that

    • paths are consistent: for all fωωf\in\omega^{\omega}, we have that {φ(x;bf|i):i<ω}\{\varphi(x;b_{f|i})\ :\ i<\omega\} is consistent, and

    • levels are kk-inconsistent: for all ηω<ω\eta\in\omega^{<\omega}, {φ(x;bηi):i<ω}\{\varphi(x;b_{\eta\frown i})\ :\ i<\omega\} is kk-inconsistent (i.e. every subset of size kk is inconsistent).

    It has the tree property (TPTP) if it has kk-TPTP for some 1<k<ω1<k<\omega

  5. (5)

    the tree property of the first kind (TP1TP_{1}) if there are tuples (bη)ηω<ω(b_{\eta})_{\eta\in\omega^{<\omega}} such that

    • paths are consistent: for all fωωf\in\omega^{\omega}, we have that {φ(x;bf|i):i<ω}\{\varphi(x;b_{f|i})\ :\ i<\omega\} is consistent, and

    • incomparable elements are inconsistent: for all incomparable η,μω<ω\eta,\mu\in\omega^{<\omega}, we have that {φ(x;bη),φ(x;bμ)}\{\varphi(x;b_{\eta}),\varphi(x;b_{\mu})\} is inconsistent.

  6. (6)

    the kk-tree property of the second kind (kk-TP2TP_{2}) if there are tuples (bij)i,j<ω(b_{i}^{j})_{i,j<\omega} such that

    • paths are consistent: for every fωωf\in\omega^{\omega}, we have that {φ(x;bif(i)):i<ω}\{\varphi(x;b_{i}^{f(i)})\ :\ i<\omega\} is consistent, and

    • rows are kk-inconsistent: for each i<ωi<\omega, {φ(x;bij):j<ω}\{\varphi(x;b_{i}^{j})\ :\ j<\omega\} is kk-inconsistent.

    It has the tree property of the second kind (TP2TP_{2}) if it has kk-TP2TP_{2} for some 1<k<ω1<k<\omega.

  7. (7n7_{n})

    nn-strong order property (SOPn)(SOP_{n}) for some n3n\geq 3, if |x|=|y||x|=|y| and there are (ai)i<ω(a_{i})_{i<\omega} such that for all i<ji<j, we have φ(ai;aj)\varphi(a_{i};a_{j}) and the set

    {φ(x1;x2),φ(x2;x3),,φ(xn1,xn),φ(xn,x1)}\{\varphi(x_{1};x_{2}),\varphi(x_{2};x_{3}),...,\varphi(x_{n-1},x_{n}),\varphi(x_{n},x_{1})\}

    is inconsistent.

A complete first-order theory TT is

  • stable if no formula has OPOP in TT. It is said to be unstable otherwise.

  • simple if no formula has TPTP in TT.

  • NpNp if no formula has pp in TT, where p{IP,SOP,TP1,TP2,SOPn}p\in\{IP,SOP,TP_{1},TP_{2},SOP_{n}\}; otherwise, we say that TT has pp. For instance, TT is NIPNIP if no formula has IPIP in TT; if there is a formula with IPIP in TT, we say that TT has IPIP.

The reader can find a diagram of known relations and implications between these properties below.

Fact 2.2.

The following relations hold for properties of complete first-order theories.

  • Stable = NIPNSOPNIP\cap NSOP.

  • Simple = NTP1NTP2NTP_{1}\cap NTP_{2}.

OPOPIPIPTPTPTP2TP_{2}TP1TP_{1}SOPnSOP_{n}SOPn+1SOP_{n+1}...SOPSOP
Proof.

For the proofs of the above we refer to [1], [5], [11] and [18]. ∎

2.2. Generalized indiscernibles

We review the basic definitions and notions around generalized indiscernibles, introduced by Shelah in [18] and developed by Scow in [17, 16].

Definition 2.3.

Fix a language \mathcal{L}^{\prime}, \mathcal{L}^{\prime}-structures II and JJ and tuples bi=f(i)b_{i}=f(i), cj=g(j)c_{j}=g(j) of some fixed length \ell for some f:I𝒰f\colon I\to\mathcal{U}^{\ell} and g:J𝒰g\colon J\to\mathcal{U}^{\ell}.

  1. (1)

    We say that (bi)iI(b_{i})_{i\in I} is an II-indexed generalized indiscernible (or just II-indiscernible) if for all n<ωn<\omega and for all i¯=i0,,in1\bar{i}=i_{0},...,i_{n-1} and j¯=j0,,jn1\bar{j}=j_{0},...,j_{n-1} from II, we have

    qftp(i¯)=qftp(j¯)tp(bi¯)=tp(bj¯)qftp_{\mathcal{L}^{\prime}}(\bar{i})=qftp_{\mathcal{L}^{\prime}}(\bar{j})\Rightarrow tp_{\mathcal{L}}(b_{\bar{i}})=tp_{\mathcal{L}}(b_{\bar{j}})
  2. (2)

    We say that (bi)iI(b_{i})_{i\in I} is based on (cj)jJ(c_{j})_{j\in J} if for any finite set of formulas Σ\Sigma, for any n<ωn<\omega and any tuple i¯=i0in1\bar{i}=i_{0}...i_{n-1} from II there is a tuple j¯=j0jn1\bar{j}=j_{0}...j_{n-1} from JJ with qftp(i¯)=qftp(j¯)qftp_{\mathcal{L}^{\prime}}(\bar{i})=qftp_{\mathcal{L}^{\prime}}(\bar{j}), such that

    tpΣ(bi¯)=tpΣ(cj¯)tp_{\Sigma}(b_{\bar{i}})=tp_{\Sigma}(c_{\bar{j}})
  3. (3)

    We say that II has the modeling property if given any (ci)iI(c_{i})_{i\in I} there exists an II-indiscernible (bi)iI(b_{i})_{i\in I} based on (ci)iI(c_{i})_{i\in I}.

Fact 2.4 (Stretching indiscernibles, see [17]).

Let II be an \mathcal{L}^{\prime}-structure, (bi)iI(b_{i})_{i\in I} II-indiscernible. Then, for any \mathcal{L}^{\prime}-structure JJ with age(J)age(I)age(J)\subseteq age(I), we can find JJ-indiscernible (cj)jJ(c_{j})_{j\in J} based on (bi)iI(b_{i})_{i\in I}.

In her thesis ([17]), Scow showed a profound connection between Ramsey classes (See, e.g., [17], Definition 3.2.6) and the modeling property for generalized indiscernibles.

Fact 2.5.

Let II be a structure in a finite relational language and 𝒦=age(I)\mathcal{K}=age(I). Then 𝒦\mathcal{K} is a Ramsey class if and only if II has the modeling property.

Fact 2.6.

Let k2k\geq 2. Applying the “main theorem” in [15], we have that the class of finite ordered kk-hypergraphs is a Ramsey class. By Fact 2.5, the generic ordered kk-hypergraph has the modeling property.

3. Patterns of consistency and inconsistency

In this section, we want to describe a general framework for talking about combinatorial properties defining dividing lines. We start with a visual representation of some of the properties of formulas defined in Definition 2.1.

The order property (OP)(OP) for a formula φ(x;y)\varphi(x;y) can be represented as follows: there are (bi)i<ω(b_{i})_{i<\omega} from 𝒰y\mathcal{U}^{y} such that, for each i<ωi<\omega we have

b0b_{0}b1b_{1}bib_{i}bi+1b_{i+1}x\exists x

where the red dashed lines represent ¬φ\neg\varphi and the green lines represent φ\varphi. Similarly, the independence property (IP)(IP) can be visualized as follows: there are (bi)i<ω(b_{i})_{i<\omega} from 𝒰y\mathcal{U}^{y} such that for every Xω{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}X}\subseteq\omega

b0b_{0}b1b_{1}b2b_{2}b3b_{3}b4b_{4}b5b_{5}x\exists x

A new element of complexity is added for describing the strict-order property (SOP)(SOP): there is (bi)i<ω(b_{i})_{i<\omega} from 𝒰y\mathcal{U}^{y} such that for all i<ωi<\omega we have

b0b_{0}bib_{i}bi+1b_{i+1}x\exists x¬x\neg\exists x

We see that the these properties are realized by satisfying consistency (x\exists x) or inconsistency (¬x\neg\exists x) requests on positive (green line i.e. φ\varphi) or negative (red dashed line i.e. ¬φ\neg\varphi) instances of the formula φ(x;y)\varphi(x;y) on subsets of the set of parameters. We are now ready to define patterns as pairs (𝒞,)(\mathcal{C},\mathcal{I}) where 𝒞\mathcal{C} and \mathcal{I} will be sets of conditions of the form (X+,X)(X^{+},X^{-}); the conceptual correspondence should be

𝒞x\displaystyle\mathcal{C}\longleftrightarrow\exists x X+φ\displaystyle X^{+}\longleftrightarrow\varphi
¬x\displaystyle\mathcal{I}\longleftrightarrow\neg\exists x X¬φ\displaystyle X^{-}\longleftrightarrow\neg\varphi
Definition 3.1.

Fix n<ωn<\omega. An nn-pattern (of consistency and inconsistency) is a pair (𝒞,)(\mathcal{C},\mathcal{I}) where 𝒞,𝒫(n)×𝒫(n){(,)}\mathcal{C},\mathcal{I}\subseteq\mathcal{P}(n)\times\mathcal{P}(n)\setminus\{(\emptyset,\emptyset)\}.

A pair (X+,X)(X^{+},X^{-}) in 𝒞\mathcal{C} (resp. in \mathcal{I}) is called a consistency (resp. inconsistency) condition, or just condition, of the pattern (𝒞,)(\mathcal{C},\mathcal{I}).

Definition 3.2.

Let φ(x;y)\varphi(x;y) be a partitioned formula, let n<ωn<\omega and let (𝒞,)(\mathcal{C},\mathcal{I}) be a nn-pattern. We say that φ(x;y)\varphi(x;y) exhibits (or realizes) the pattern (𝒞,)(\mathcal{C},\mathcal{I}) (in TT) if there is B=(bi)i<nB=(b_{i})_{i<n} from 𝒰y\mathcal{U}^{y} such that

  • (i)

    For all 𝒜=(A+,A)𝒞\mathcal{A}=(A^{+},A^{-})\in\mathcal{C} the partial φ\varphi-type

    p𝒜B(x)={φ(x;bi):iA+}{¬φ(x;bj):jA}p_{\mathcal{A}}^{B}(x)=\{\varphi(x;b_{i})\ :\ i\in A^{+}\}\cup\{\neg\varphi(x;b_{j})\ :\ j\in A^{-}\}

    is consistent, and

  • (ii)

    for all 𝒵=(Z+,Z)\mathcal{Z}=(Z^{+},Z^{-})\in\mathcal{I}, the partial φ\varphi-type

    p𝒵B(x)={φ(x;bi):iZ+}{¬φ(x;bj):jZ}p_{\mathcal{Z}}^{B}(x)=\{\varphi(x;b_{i})\ :\ i\in Z^{+}\}\cup\{\neg\varphi(x;b_{j})\ :\ j\in Z^{-}\}

    is inconsistent.

In the next proposition we show how to characterize some of the dividing lines using exhibition of patterns.

Proposition 3.3.

Let φ(x;y)\varphi(x;y) be a partitioned formula. Then

  1. (1)

    φ(x;y)\varphi(x;y) has OPOP if and only if, for all n<ωn<\omega, the formula φ(x;y)\varphi(x;y) exhibits the nn-pattern (𝒞n,n)(\mathcal{C}_{n},\mathcal{I}_{n}) where

    𝒞n={({i,i+1,,n1},{0,,i1}):i<n}\mathcal{C}_{n}=\{(\{i,i+1,...,n-1\},\{0,...,i-1\})\ :\ i<n\}
    n=\mathcal{I}_{n}=\emptyset
  2. (2)

    φ(x;y)\varphi(x;y) has IPIP if and only if, for all n<ωn<\omega, the formula φ(x;y)\varphi(x;y) exhibits the nn-pattern (𝒞n,n)(\mathcal{C}_{n},\mathcal{I}_{n}) where

    𝒞n={(X,nX):Xn}\mathcal{C}_{n}=\{(X,n\setminus X)\ :\ X\subseteq n\}
    n=\mathcal{I}_{n}=\emptyset
  3. (3)

    φ(x;y)\varphi(x;y) has SOPSOP if and only if, for all n<ωn<\omega, the formula φ(x;y)\varphi(x;y) exhibits the nn-pattern (𝒞n,n)(\mathcal{C}_{n},\mathcal{I}_{n}) where

    𝒞n={({i+1},{i}):i<n1}\mathcal{C}_{n}=\{(\{i+1\},\{i\})\ :\ i<n-1\}
    n={({i},{i+1}):i<n1}\mathcal{I}_{n}=\{(\{i\},\{i+1\})\ :\ i<n-1\}
  4. (4)

    For 1<k<ω1<k<\omega, φ(x;y)\varphi(x;y) has kk-TPTP if and only if, for all n<ωn<\omega, the formula φ(x;y)\varphi(x;y) exhibits the n<nn^{<n}-pattern (𝒞n,n)(\mathcal{C}_{n},\mathcal{I}_{n}) where

    𝒞n={( pathi,):inn}\mathcal{C}_{n}=\{(\text{ path}_{i},\emptyset)\ :\ i\in n^{n}\}
    n={({i0,,ik1},):i0,,ik1 children of the same node}\mathcal{I}_{n}=\{(\{i_{0},...,i_{k-1}\},\emptyset)\ :\ i_{0},...,i_{k-1}\text{ children of the same node}\}

    where {pathi:inn}\{path_{i}\ :\ i\in n^{n}\} is the collection of all the paths in the tree n<nn^{<n}.

  5. (5)

    φ(x;y)\varphi(x;y) has TP1TP_{1} if and only if, for all n<ωn<\omega, the formula φ(x;y)\varphi(x;y) exhibits the n<nn^{<n}-pattern (𝒞n,n)(\mathcal{C}_{n},\mathcal{I}_{n}) where

    𝒞n={( pathi,):inn}\mathcal{C}_{n}=\{(\text{ path}_{i},\emptyset)\ :\ i\in n^{n}\}
    n={({i,j},):i,j are incomparable elements in the tree n<n}\mathcal{I}_{n}=\{(\{i,j\},\emptyset)\ :\ i,j\text{ are incomparable elements in the tree }n^{<n}\}

    with a similar notation as above.

  6. (6)

    For 1<k<ω1<k<\omega, φ(x;y)\varphi(x;y) has kk-TP2TP_{2} if and only if, for all n<ωn<\omega, the formula φ(x;y)\varphi(x;y) exhibits the n2n^{2}-pattern (𝒞n,n)(\mathcal{C}_{n},\mathcal{I}_{n}) where

    𝒞n={( pathi,):inn}\mathcal{C}_{n}=\{(\text{ path}_{i},\emptyset)\ :\ i\in n^{n}\}
    n={({i0,,ik1},):i0,,ik1 are on the same row }\mathcal{I}_{n}=\{(\{i_{0},...,i_{k-1}\},\emptyset)\ :\ i_{0},...,i_{k-1}\text{ are on the same row }\}

    Here, by a path in n2n^{2}, we mean a function f:nnf\colon n\to n i.e. a choice of one element per row in the array n2n^{2}.

Proof.

The characterization is clear. We get the usual definitions of the properties using compactness. ∎

The above proposition shows that exhibition of patterns gives a general framework for talking about dividing lines in first-order theories. There are some exceptions. For instance, it is not clear whether the SOPnSOP_{n} hierarchy, for n4n\geq 4, can be characterized using pattern exhibition. There exists a characterization of SOP3SOP_{3} ([9], Lemma 7.3) that can be described through patterns.

Definition 3.4.

We say that an nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) is exhibitable if there is a theory TT and a formula φ(x;y)\varphi(x;y) such that φ\varphi exhibits (𝒞,)(\mathcal{C},\mathcal{I}) in TT.

There are “syntactical reasons” that make a pattern not exhibitable; with the next definition we want to exclude those reasons from the picture.

Definition 3.5.

Fix n<ωn<\omega. A reasonable nn-pattern (of consistency and inconsistency) is an nn-pattern such that

  • (i)

    For all (Z+,Z)(Z^{+},Z^{-})\in\mathcal{I} and for all (Y+,Y)𝒞(Y^{+},Y^{-})\in\mathcal{C}, we have that ZϵYϵZ^{\epsilon}\not\subseteq Y^{\epsilon} for some ϵ{+,}\epsilon\in\{+,-\}.

  • (ii)

    For all (A+,A)𝒞(A^{+},A^{-})\in\mathcal{C}, A+A=A^{+}\cap A^{-}=\emptyset.

  • (iii)

    For all (Z+,Z)(Z^{+},Z^{-})\in\mathcal{I}, Z+Z=Z^{+}\cap Z^{-}=\emptyset.

Remark 3.6.

In the above definition, a failure of condition (iii) doesn’t really make a pattern not exhibitable; however, an inconsistency condition (Z+,Z)(Z^{+},Z^{-}) for which Z+ZZ^{+}\cap Z^{-}\neq\emptyset doesn’t add any meaningful request to the pattern and thus we avoid it. We also want to point out that “reasonability” does not imply “exhibitability”. For example, a pattern in which \mathcal{I} contains both ({0},)(\{0\},\emptyset) and (,{0})(\emptyset,\{0\}) is not exhibitable.

Even assuming reasonability, deciding whether a pattern is exhibitable or not is a complex problem. More precisely, let PSATPSAT be the following decision problem: given n<ωn<\omega and a reasonable nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}), is (𝒞,)(\mathcal{C},\mathcal{I}) exhibitable?

Proposition 3.7.

The decision problem PSATPSAT is NPNP-complete.

The proof of the above will appear in the author’s PhD thesis.

Definition 3.8.

Fix n<ωn<\omega. A nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) is called condition complete, or just complete, if each one of its conditions is of the form (X,nX)(X,n\setminus X).

Remark 3.9.

Every complete pattern (𝒞,)(\mathcal{C},\mathcal{I}) with 𝒞=\mathcal{C}\cap\mathcal{I}=\emptyset is reasonable.

Remark 3.10.

As a justification for the above nomenclature, consider that, given any nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) and a formula φ(x;y)\varphi(x;y) exhibiting it with witnesses B=(bi)i<nB=(b_{i})_{i<n}, each condition (X+,X)(X^{+},X^{-}) is associated with a φ\varphi-type over BB:

pX+,XB(x)={φ(x;bi):iX+}{¬φ(x;bj):jX}p_{X^{+},X^{-}}^{B}(x)=\{\varphi(x;b_{i})\ :\ i\in X^{+}\}\cup\{\neg\varphi(x;b_{j})\ :\ j\in X^{-}\}

that is asked to be consistent or inconsistent based on where the condition lies (in 𝒞\mathcal{C} or in \mathcal{I} respectively). Saying that a pattern is complete is saying that those types corresponding to conditions are all complete φ\varphi-types over BB.

Definition 3.11.

A nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) is called fully complete if it is complete, 𝒞\mathcal{C}\neq\emptyset and every condition of the form (X,nX)(X,n\setminus X) is either in 𝒞\mathcal{C} or in \mathcal{I} but not both.

Proposition 3.12.

Every fully complete pattern is exhibitable.

Proof.

We actually show something stronger: there is a theory TT and a formula φ\varphi such that, for every n<ωn<\omega and fully complete nn-pattern, φ\varphi exhibits it.

Let TT be the theory of infinite atomic Boolean algebras in the language ={,,()c,0,1}\mathcal{L}=\{\cup,\cap,(-)^{c},0,1\}, work in a model \mathcal{B}, let ψ(x;y)=At(x)xy\psi(x;y)=At(x)\land x\leq y and define

φ(x;yw0w1)=(ψ(x;y)w0=w1)(¬ψ(x;y)w0w1)\varphi(x;yw_{0}w_{1})=(\psi(x;y)\land w_{0}=w_{1})\lor(\neg\psi(x;y)\land w_{0}\neq w_{1})

where At(x)At(x) is the formula saying that xx is an atom. Fix n<ωn<\omega and a fully complete nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}). Enumerate 𝒞={(Aj,nAj):j<k}\mathcal{C}=\{(A_{j},n\setminus A_{j})\ :\ j<k\}. Since the pattern is fully complete, 𝒞\mathcal{C}\neq\emptyset, so k1k\geq 1. Moreover, let a0,,ak1a_{0},...,a_{k-1} be kk distinct atoms of \mathcal{B}.

First, assume that (,n)(\emptyset,n)\notin\mathcal{I}. Let cc any element of \mathcal{B} and, for each i<ni<n, define

bi={aj:j<k,iAj}b_{i}=\bigcup\{a_{j}\ :\ j<k,i\in A_{j}\}

and ci0=ci1=cc_{i}^{0}=c_{i}^{1}=c. We claim that (bici0ci1)i<n(b_{i}c_{i}^{0}c_{i}^{1})_{i<n} witness the exhibition of (𝒞,)(\mathcal{C},\mathcal{I}) by φ\varphi. Indeed, let j<kj<k and consider the condition (Aj,nAj)𝒞(A_{j},n\setminus A_{j})\in\mathcal{C}. We have that iAji\in A_{j} if and only if ajbia_{j}\leq b_{i} if and only if ajAt(x)xbia_{j}\models At(x)\land x\leq b_{i} if and only if ajφ(x;bici0ci1)a_{j}\models\varphi(x;b_{i}c_{i}^{0}c_{i}^{1}). This holds for every j<kj<k and thus any consistency condition is satisfied. Let (Z,nZ)(Z,n\setminus Z)\in\mathcal{I}. By assumption ZZ\neq\emptyset. Assume, toward contradiction, that there is a𝒰xa\in\mathcal{U}^{x}, such that φ(a;bici0ci1)\models\varphi(a;b_{i}c_{i}^{0}c_{i}^{1}) if and only if iZi\in Z. Since ZZ\neq\emptyset, there is at least one ii such that aa is an atom below bib_{i}. Thus there exists j<kj<k such that a=aja=a_{j}. But now iZi\in Z if and only if φ(aj;bici0ci1)\models\varphi(a_{j};b_{i}c_{i}^{0}c_{i}^{1}) if and only if ajbia_{j}\leq b_{i} if and only if iAji\in A_{j}. Thus Z=AjZ=A_{j} and (Z,nZ)𝒞(Z,n\setminus Z)\in\mathcal{C}. Contradiction. Thus, in the case that (,n)(\emptyset,n)\notin\mathcal{I}, we are done.

Assume now that (,n)(\emptyset,n)\in\mathcal{I}. Fix two distinct elements of the model c,cc,c^{\prime} and for i<ni<n, iA0i\notin A_{0}, define

bi={aj:j<k,iAj} and take ci0=c=ci1b_{i}=\bigcup\{a_{j}\ :\ j<k,i\in A_{j}\}\quad\text{ and take }c_{i}^{0}=c=c_{i}^{1}

For iA0i\in A_{0}, define

bi={aj:j<k,iAj} and take ci0=cc=ci1b_{i}=\bigcup\{a_{j}\ :\ j<k,i\notin A_{j}\}\quad\text{ and take }c_{i}^{0}=c\neq c^{\prime}=c_{i}^{1}

Let B=(bici0ci1)i<nB=(b_{i}c_{i}^{0}c_{i}^{1})_{i<n}. Fix (Aj,nAj)𝒞(A_{j},n\setminus A_{j})\in\mathcal{C} and suppose j0j\neq 0. If inA0i\in n\setminus A_{0}, we have

ajφ(x;bici¯)At(aj)(ajbi)ajbiiAjA0a_{j}\models\varphi(x;b_{i}\bar{c_{i}})\iff At(a_{j})\land(a_{j}\leq b_{i})\iff a_{j}\leq b_{i}\iff i\in A_{j}\setminus A_{0}

If iA0i\in A_{0}, we have

ajφ(x;bici¯)¬At(aj)(ajbi)a_{j}\models\varphi(x;b_{i}\bar{c_{i}})\iff\neg At(a_{j})\lor(a_{j}\not\leq b_{i})
ajbiajbic=(iAjaj)c=iAjajciAj\iff a_{j}\not\leq b_{i}\iff a_{j}\leq b_{i}^{c}=(\bigcup_{i\notin A_{j}}a_{j})^{c}=\bigcap_{i\notin A_{j}}a_{j}^{c}\iff i\in A_{j}

Hence, if (Aj,nAj)(A0,nA0)(A_{j},n\setminus A_{j})\neq(A_{0},n\setminus A_{0}), ajpAj,nAjB(x)a_{j}\models p_{A_{j},n\setminus A_{j}}^{B}(x). Consider now (A0,nA0)(A_{0},n\setminus A_{0}) and fix an element bb that is not an atom. Then, if iA0i\in A_{0}, since b¬At(x)b\models\neg At(x), bφ(x;bicc)b\models\varphi(x;b_{i}cc^{\prime}). If iA0i\notin A_{0}, then, since b⊧̸At(x)b\not\models At(x), b¬φ(x;bicc)b\models\neg\varphi(x;b_{i}cc). Thus, we showed that each consistency condition is satisfied.

Fix now (Z,nZ)(Z,n\setminus Z)\in\mathcal{I}. First assume that Z=Z=\emptyset; we want to show that for any aa\in\mathcal{B}, there is i<ni<n such that aφ(x;bic¯i)a\models\varphi(x;b_{i}\bar{c}_{i}). If aAt()a\notin At(\mathcal{B}), then a¬At(x)a\models\neg At(x) and hence aφ(x;bi,cc)a\models\varphi(x;b_{i},cc^{\prime}) for any iA0i\in A_{0}. If aAt()(aj)0<j<ka\in At(\mathcal{B})\setminus(a_{j})_{0<j<k}, then abia\not\leq b_{i} for iA0i\in A_{0} and hence aφ(x;bicc)a\models\varphi(x;b_{i}cc^{\prime}). If a=aja0a=a_{j}\neq a_{0}, then ajφ(x;bicc)a_{j}\models\varphi(x;b_{i}cc) for iAji\in A_{j} (Note that AjA_{j}\neq\emptyset since (,n)(\emptyset,n)\in\mathcal{I}). Hence p(,n)B(x)p_{(\emptyset,n)}^{B}(x) is inconsistent.

Assume ZZ\neq\emptyset and suppose that there is aa\in\mathcal{B} such that

aφ(x;bic¯i)iZa\models\varphi(x;b_{i}\bar{c}_{i})\iff i\in Z

If aa is not one of the (aj)j<k(a_{j})_{j<k}, then aφ(x;bic¯i)iA0a\models\varphi(x;b_{i}\bar{c}_{i})\iff i\in A_{0} and hence A0=ZA_{0}=Z. Contradiction.

If a=aja=a_{j} for some 0<j<k0<j<k, then a=ajφ(x;bic¯i)iAja=a_{j}\models\varphi(x;b_{i}\bar{c}_{i})\iff i\in A_{j} and hence Aj=ZA_{j}=Z. Contradiction. If a=a0a=a_{0}, then if iA0i\in A_{0} then a0=abia_{0}=a\not\leq b_{i} and hence aφ(x;bicc)a\models\varphi(x;b_{i}cc^{\prime}); on the other hand, if iA0i\notin A_{0}, then a0=abia_{0}=a\not\leq b_{i} and hence a⊧̸φ(x;bicc)a\not\models\varphi(x;b_{i}cc). Hence Z=A0Z=A_{0}. Contradiction. Hence p(Z,nZ)B(x)p_{(Z,n\setminus Z)}^{B}(x) is inconsistent.

In conclusion, for every n<ωn<\omega, φ\varphi exhibits every fully complete nn-pattern. ∎

Remark 3.13.

In the above proof we gave an example of a formula φ\varphi that, in a theory TT, exhibits, for every n<ωn<\omega, every fully complete nn-pattern. In the next proposition, we will show that any exhibitable pattern can be extended (in some sense) to a fully complete pattern. Thus, the above proof gives an example of a theory TT and a formula φ\varphi such that for any n<ωn<\omega and any exhibitable nn-pattern, φ\varphi exhibits it. This property is what, in the next section, we will call Straight up Maximality (similar to Shelah’s nomenclature from [23]); thus the theory of infinite atomic Boolean algebras is an example of a straight up maximal theory.

Proposition 3.14.

Fix n<ωn<\omega and let (𝒞,)(\mathcal{C},\mathcal{I}) any exhibitable nn-pattern. Then, there exists a fully complete nn-pattern (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}) such that, if a formula exhibits (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}), then it exhibits (𝒞,)(\mathcal{C},\mathcal{I}).

Proof.

Since (𝒞,)(\mathcal{C},\mathcal{I}) is exhibitable, there is a theory TT, a formula φ(x;y)\varphi(x;y) and parameters B=(bi)i<nB=(b_{i})_{i<n} such that φ\varphi exhibits (𝒞,)(\mathcal{C},\mathcal{I}) in TT with BB. As a convention for this proof, a complete φ\varphi-type over BB is not necessarily consistent. Let Con(φ,B)Con(\varphi,B) be the set of complete consistent φ\varphi-types over BB and, similarly, let Inc(φ,B)Inc(\varphi,B) be the set of complete inconsistent φ\varphi-types over BB. To each complete φ\varphi-type pp, we can associate a condition (Xp,nXp)(X_{p},n\setminus X_{p}) where Xp={i<n:φ(x;bi)p}X_{p}=\{i<n\ :\ \varphi(x;b_{i})\in p\}. Define

𝒞={(Ap,nAp):pCon(φ,B)}\mathcal{C}^{\prime}=\{(A_{p},n\setminus A_{p})\ :\ p\in Con(\varphi,B)\}
={(Zp,nZp):pInc(φ,B)}\mathcal{I}^{\prime}=\{(Z_{p},n\setminus Z_{p})\ :\ p\in Inc(\varphi,B)\}

Suppose ψ(x;y)\psi(x;y) exhibits (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}) in TT with B=(bi)i<nB^{\prime}=(b_{i}^{\prime})_{i<n}.

Given (A+,A)𝒞(A^{+},A^{-})\in\mathcal{C}, since φ(x;y)\varphi(x;y) exhibits (𝒞,)(\mathcal{C},\mathcal{I}) with BB, we have that p(A+,A)B,φ(x)p_{(A^{+},A^{-})}^{B,\varphi}(x) is consistent and thus it is contained in some pCon(φ,B)p\in Con(\varphi,B). Thus, there is (A,nA)𝒞(A,n\setminus A)\in\mathcal{C}^{\prime} such that (A+,A)(A,nA)(A^{+},A^{-})\subseteq(A,n\setminus A). Since p(A,nA)B,ψ(x)p_{(A,n\setminus A)}^{B^{\prime},\psi}(x) is consistent, it must be that p(A+,A)B,ψ(x)p_{(A^{+},A^{-})}^{B^{\prime},\psi}(x) is consistent too.

Let (Z+,Z)(Z^{+},Z^{-})\in\mathcal{I} and assume, toward contradiction, that p(Z+,Z)B,ψ(x)p_{(Z^{+},Z^{-})}^{B^{\prime},\psi}(x) is realized by some a𝒰xa\in\mathcal{U}^{x}. By the choice of φ\varphi and BB, we have that p(Z+,Z)B,φ(x)p_{(Z^{+},Z^{-})}^{B,\varphi}(x) is inconsistent; thus, for any XnX\subseteq n such that (Z+,Z)(X,nX)(Z^{+},Z^{-})\subseteq(X,n\setminus X), we have that (X,nX)(X,n\setminus X)\in\mathcal{I}^{\prime}. But now tpψ(a/B)tp^{\psi}(a/B^{\prime}) gives us a set XnX\subseteq n such that (Z+,Z)(X,nX)(Z^{+},Z^{-})\subseteq(X,n\setminus X) which is in \mathcal{I}^{\prime} and the corresponding type has a realization. Contradiction. ∎

4. Notions of Maximality

In this section we define a notion of maximal complexity for first-order theories by requesting the exhibition (by a single formula) of all the exhibitable patterns.

Definition 4.1.

A formula φ(x;y)\varphi(x;y) is Straight up Maximal (or just SMSM) if, for every n<ωn<\omega and for every exhibitable nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}), φ(x;y)\varphi(x;y) exhibits (𝒞,)(\mathcal{C},\mathcal{I}). A theory TT is SMSM if there is a formula which is SMSM; otherwise we say that TT is NSMNSM.

Other notions of maximality were previously defined by Shelah ([23]) and Cooper ([7]). We will see how both of them agree with our notion of maximality and we will give some examples of theories which present/omit this property.

4.1. Shelah’s maximality

In [23], Shelah defined “an extreme condition of the form of unstability” called Straight Maximality.

Definition 4.2.

A formula φ(x;y)\varphi(x;y) is Shelah-SMSM (Shelah’s Straight maximal) if for every n<ωn<\omega and for every non-empty 2n\mathscr{F}\subseteq 2^{n} there are b0,,bn1𝒰yb_{0},...,b_{n-1}\in\mathcal{U}^{y} such that: Given f2nf\in 2^{n}, ff\in\mathscr{F} if and only if there is a𝒰xa\in\mathcal{U}^{x} such that for every i<ni<n, φ(a;bi)f(i)=1\models\varphi(a;b_{i})\iff f(i)=1.

We can characterize the above property using patterns.

Proposition 4.3.

A formula φ(x;y)\varphi(x;y) is Shelah-SMSM if and only if φ(x;y)\varphi(x;y) exhibits every fully complete pattern.

Proof.

Assume φ\varphi is Shelah-SMSM, fix n<ωn<\omega and let (𝒞,)(\mathcal{C},\mathcal{I}) a fully complete nn-pattern. Define ={char(X):(X,nX)𝒞}2n\mathscr{F}=\{char(X)\ :\ (X,n\setminus X)\in\mathcal{C}\}\subseteq 2^{n} where char(X):n2char(X)\colon n\to 2 is the characteristic function of XnX\subseteq n. Since 𝒞\mathcal{C}\neq\emptyset (given that the pattern is fully complete), we have that \mathscr{F} is not empty. Let B={b0,,bn1}B=\{b_{0},...,b_{n-1}\} given by Shelah-SMSM. Now, fix a condition (Y+,Y)=(Y,nY)𝒞(Y^{+},Y^{-})=(Y,n\setminus Y)\in\mathcal{C}. This corresponds to a function f:n2f\colon n\to 2, ff\in\mathscr{F}, such that f(i)=1f(i)=1 if and only if iYi\in Y. Hence there is a𝒰xa\in\mathcal{U}^{x} such that φ(a;bi)\models\varphi(a;b_{i}) if and only if f(i)=1f(i)=1 if and only if iXi\in X; thus aa witness the consistency of pY+,YB(x)p_{Y^{+},Y^{-}}^{B}(x). Now, given a condition of inconsistency (Z+,Z)=(Z,nZ)(Z^{+},Z^{-})=(Z,n\setminus Z)\in\mathcal{I}, if there was a𝒰xa\in\mathcal{U}^{x} such that φ(a;bi)\models\varphi(a;b_{i}) if and only if iZi\in Z if and only if char(Z)(i)=1char(Z)(i)=1, we would have that char(Z)char(Z)\in\mathscr{F}; but it’s not. Contradiction. Hence the type pZ+,ZB(x)p_{Z^{+},Z^{-}}^{B}(x) is inconsistent. This shows that φ\varphi exhibits (𝒞,)(\mathcal{C},\mathcal{I}). In conclusion (of the proof of this direction), φ\varphi exhibits every fully complete pattern.

On the other hand, assume that φ\varphi exhibits every fully complete pattern. Given n<ωn<\omega and 2n\emptyset\neq\mathscr{F}\subseteq 2^{n}, define the nn-pattern. (𝒞,)(\mathcal{C},\mathcal{I}) as follows

𝒞={(f1({1}),f1({0})):f}\mathcal{C}=\{(f^{-1}(\{1\}),f^{-1}(\{0\}))\ :\ f\in\mathscr{F}\}
={(f1({1}),f1({0})):f2n}\mathcal{I}=\{(f^{-1}(\{1\}),f^{-1}(\{0\}))\ :\ f\in 2^{n}\setminus\mathscr{F}\}

Clearly this is a fully complete pattern. Along the lines of the proof of the previous part, it’s clear that a witness B={b0,,bn1}B=\{b_{0},...,b_{n-1}\} of the exhibition of (𝒞,)(\mathcal{C},\mathcal{I}) by φ\varphi is also a witness of φ\varphi being Shelah-SMSM. ∎

Proposition 4.4.

A formula is Shelah-SMSM if and only if it is SMSM.

Proof.

Let φ(x;y)\varphi(x;y) be a SMSM formula i.e. it realizes every exhibitable pattern. Since every fully complete pattern is exhibitable, φ\varphi exhibits every fully complete pattern. Hence φ\varphi is Shelah-SMSM by Proposition 4.3.

On the other hand, assume that φ(x;y)\varphi(x;y) is Shelah-SMSM; then it realizes every fully complete pattern. Given an exhibitable pattern (𝒞,)(\mathcal{C},\mathcal{I}), by Proposition 3.14, there is a fully complete pattern (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}) such that, if a formula realizes (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}) then it realizes (𝒞,)(\mathcal{C},\mathcal{I}). Since φ\varphi realizes every fully complete pattern, it realizes every exhibitable pattern. In conclusion, it is SMSM. ∎

4.2. Cooper’s maximality

In [7], Cooper defined a condition called 11, requesting the existence of a formula which defines, in some sense, all the possible Venn diagrams. They then proved that this property implies every other property of this kind. For more details about the background and the maximality property, we refer to [7].

Definition 4.5 (Cooper, [7]).
  • A property of formulas is any sequence ρ=(ρi)i<ω\rho=(\rho_{i})_{i<\omega} of consistent quantifier free (q.f.) formulas of BABA (the theory of Boolean Algebras with the language ={0,1,()c,,}\mathcal{L}=\{0,1,(-)^{c},\cup,\cap\}).

  • If φ(x,y)\varphi(x,y) is a formula of a complete theory TT and ψ(z)\psi(z) is a q.f. formula of BABA, we say that φ(x,y)\varphi(x,y) admits ψ(z)\psi(z) in TT if

    𝒫(x)ψ(φ(,a0),,φ(,a|z|1))\mathcal{P}(\mathcal{M}^{x})\models\psi(\varphi(\mathcal{M},a_{0}),...,\varphi(\mathcal{M},a_{|z|-1}))

    for some T\mathcal{M}\models T and some a0,,a|z|1xa_{0},...,a_{|z|-1}\in\mathcal{M}^{x} i.e. there is some model of TT and parameters such that, in the power set of the set of tuples of the model (of the right length) viewed as a Boolean algebra, the φ\varphi-definable sets given by the parameters satisfy ψ\psi. Otherwise we say that φ(x,y)\varphi(x,y) omits ψ(z)\psi(z) in TT.

  • If ρ\rho is a property of formulas, we say that a formula φ(x,y)\varphi(x,y) of a complete theory TT admits ρ\rho in TT if there exists a strictly increasing sequence αωω\alpha\in\omega^{\omega} such that φ(x,y)\varphi(x,y) admits ρα(i)\rho_{\alpha(i)} in TT for every i<ωi<\omega. Otherwise we say that φ(x,y)\varphi(x,y) omits ρ\rho in TT. Moreover we say that TT admits ρ\rho if there exists φ(x,y)\varphi(x,y) of TT such that φ(x,y)\varphi(x,y) admits ρ\rho in TT. Otherwise TT omits ρ\rho.

Cooper’s notion of maximality (property 11) is a property of formulas (as in Definition 4.5) which implies all the other properties of formulas. In order to define 11, Cooper started with an enumeration of all the consistent q.f. formulas of BABA, say {ψi:i<ω}\{\psi_{i}\ :\ i<\omega\}, and defined

ρi=j<iψi\rho_{i}=\bigwedge_{j<i}\psi_{i}

Finally, 11 is defined as the property of formulas (ρi)i<ω(\rho_{i})_{i<\omega}.

The next definition gives us a more explicit equivalent (at the level of theories) version of property 11.

Definition 4.6 (Cooper, [7]).

For each 0<n<ω0<n<\omega we define the property 1(n)1^{(n)} as follows:

A formula φ(x,y)\varphi(x,y) admits 1(n)1^{(n)} in some complete theory TT if for arbitrarily large nm<ωn\leq m<\omega there are (bi)i<m(b_{i})_{i<m} in some model 𝒰y\mathcal{U}^{y} such that

  • (i)

    for any lml\leq m and for any X𝒫(m)X\in\mathcal{P}(m) with |X|=l|X|=l there is bXb_{X} such that

    iXφ(𝒰,bi)=φ(𝒰,bX)\bigcup_{i\in X}\varphi(\mathcal{U},b_{i})=\varphi(\mathcal{U},b_{X})
  • (ii)

    for any Y𝒫(m)Y\in\mathcal{P}(m)

    iYφ(𝒰,bi)= if and only if |Y|>n\bigcap_{i\in Y}\varphi(\mathcal{U},b_{i})=\emptyset\text{ if and only if }|Y|>n

In other words, φ(x,y)\varphi(x,y) admits 1(n)1^{(n)} if and only if for arbitrarily large nm<ωn\leq m<\omega there are mm φ\varphi-definable sets such that the union of any lml\leq m is still φ\varphi-definable but the intersection of any lml\leq m of them is non-empty if and only if lnl\leq n.

Fact 4.7 (Cooper, [7]).

If TT is a complete theory and 0<n<ω0<n<\omega, the following hold:

  • (1)

    If some formula φ(x,y)\varphi(x,y) of TT admits 1(n)1^{(n)} in TT then some formula ψ(x,z)\psi(x,z) of TT admits 1(1)1^{(1)} in TT.

  • (2)

    If some formula ψ(x,z)\psi(x,z) of TT admits 1(1)1^{(1)} in TT, then some formula χ(x,w)\chi(x,w) of TT admits 11 in TT.

Remark 4.8.

The above fact shows that, at the level of theories, admitting the property 1(n)1^{(n)} implies admitting the property 1(1)1^{(1)} which, in turn, implies admitting the property 11. Moreover, since each 1(n)1^{(n)} can be described as a property of formulas and each property of formulas is implied by 11, we have that, at the level of theories, each 1(n)1^{(n)} is a different characterization of 11.

Definition 4.9.

A complete theory TT is Cooper maximal if there is a formula which admits 11 (equivalently 1(n)1^{(n)}) in TT.

We are now ready to prove that Cooper’s maximality coincides with straight up maximality.

Proposition 4.10.

Let TT be a complete theory. TT is Cooper maximal if and only if TT is SMSM.

Proof.

Assume TT is SMSM and let φ(x;y)\varphi(x;y) be the SMSM formula witnessing it.

Claim for each n<ωn<\omega there are (bi)i<n(b_{i})_{i<n} such that

φ(𝒰,bi)φ(𝒰,bj)= for all ij<n\varphi(\mathcal{U},b_{i})\cap\varphi(\mathcal{U},b_{j})=\emptyset\text{ for all }i\neq j<n

for each XnX\subseteq n, there is bXb_{X} such that

φ(𝒰,bX)=iXφ(𝒰,bi)\varphi(\mathcal{U},b_{X})=\bigcup_{i\in X}\varphi(\mathcal{U},b_{i})

Equivalently, φ(x;y)\varphi(x;y) admits 1(1)1^{(1)}.

Proof of the Claim: Since φ(x;y)\varphi(x;y) is SMSM, it realizes every fully complete nn-patterns and these patterns can be described by pairs (C,I)(C,I) where C,I𝒫(n)C,I\subseteq\mathcal{P}(n) with CI=C\cap I=\emptyset. Let n<ωn<\omega and consider the following 2n2^{n}-pattern:

C={{i}:i<n}C=\{\uparrow\{i\}\ :\ i<n\}
I=𝒫(2n)CI=\mathcal{P}(2^{n})\setminus C

where {i}={Xn:iX}\uparrow\{i\}=\{X\subseteq n\ :\ i\in X\}. Clearly IC=I\cap C=\emptyset. Let (bX)Xn(b_{X})_{X\subseteq n} be the parameters given by SMSM with this pattern. First notice that φ(𝒰,b{i})φ(𝒰,b{j})=\varphi(\mathcal{U},b_{\{i\}})\cap\varphi(\mathcal{U},b_{\{j\}})=\emptyset. Indeed, suppose aφ(x,b{i})φ(x,b{j})a\models\varphi(x,b_{\{i\}})\land\varphi(x,b_{\{j\}}). Let A={{k}n:aφ(x,b{k})}A=\{\{k\}\subseteq n\ :\ a\models\varphi(x,b_{\{k\}})\}. Since AIA\in I we have that the φ\varphi-type of aa over (bX)Xn(b_{X})_{X\subseteq n} is inconsistent; contradiction.

Let XnX\subseteq n. We want to show that

φ(𝒰,bX)=iXφ(𝒰,b{i})\varphi(\mathcal{U},b_{X})=\bigcup_{i\in X}\varphi(\mathcal{U},b_{\{i\}})

Let aφ(x,b{i})a\models\varphi(x,b_{\{i\}}) for some iXi\in X. Since 𝒫(n)XI\mathcal{P}(n)\setminus X\in I, it is inconsistent to be in φ(𝒰,b{i})\varphi(\mathcal{U},b_{\{i\}}) but not in φ(𝒰,bX)\varphi(\mathcal{U},b_{X}); hence aφ(𝒰,bX)a\in\varphi(\mathcal{U},b_{X}). If aφ(x,bX)a\models\varphi(x,b_{X}). Since {X}I\{X\}\in I, it is inconsistent to be in φ(𝒰,bX)\varphi(\mathcal{U},b_{X}) but not in all the φ(x,b{i})\varphi(x,b_{\{i\}}), hence there is iXi\in X such that aφ(𝒰,b{i})a\in\varphi(\mathcal{U},b_{\{i\}}). \blacksquare

The claim shows that, if φ(x;y)\varphi(x;y) is SMSM, then it admits 1(1)1^{(1)}. By Fact 4.7, there is a formula admitting 11 and thus TT is Cooper maximal.

Assume now that TT is Cooper maximal and let φ(x;y)\varphi(x;y) admitting 11. Since 11 is the strongest property of formulas, it is enough to prove that SMSM can be described as a property of formulas. Indeed SMSM is a property of formulas described by:

ρm:=n<m(𝒞,) exhibitable n-p((Y+,Y)𝒞(iY+xijY(nxj))0)\rho_{m}:=\bigwedge_{n<m}\bigwedge_{\begin{subarray}{c}(\mathcal{C},\mathcal{I})\\ \text{ exhibitable $n$-p}\end{subarray}}(\bigwedge_{(Y^{+},Y^{-})\in\mathcal{C}}(\bigcap_{i\in Y^{+}}x_{i}\cap\bigcap_{j\in Y^{-}}(n\setminus x_{j}))\neq 0)
(Z+,Z)(iZ+xijZ(nxj))=0)\land\bigwedge_{(Z^{+},Z^{-})\in\mathcal{I}}(\bigcap_{i\in Z^{+}}x_{i}\cap\bigcap_{j\in Z^{-}}(n\setminus x_{j}))=0)

This shows that, if TT is Cooper maximal, then it is SMSM. ∎

Remark 4.11.

We want to point out that the above equivalence is only at the level of theories. Indeed, while a formula admitting 11 is SMSM, if a formula is SMSM, it admits 1(1)1^{(1)} and thus (by Fact 4.7) there is a (possibly different) formula admitting 11.

4.3. Examples of SMSM theories

Example 4.12 (Infinite atomic Boolean algebra).

In Proposition 3.12, we showed that the theory TT of infinite atomic Boolean algebras is SMSM by showing a formula that exhibits every exhibitable pattern. We now prove again this result using Cooper’s characterization of SMSM (See Fact 4.7 and Proposition 4.10): It is enough to find a formula with 1(1)1^{(1)} i.e. a formula φ(x;y)\varphi(x;y) such that, for any n<ωn<\omega there are (bi)i<n(b_{i})_{i<n} such that

  • φ(𝒰;bi)φ(𝒰;bj)=\varphi(\mathcal{U};b_{i})\cap\varphi(\mathcal{U};b_{j})=\emptyset for all i<j<ni<j<n, and

  • for each XnX\subseteq n, there is bXb_{X} such that

    φ(𝒰;bX)=iXφ(𝒰;bi)\varphi(\mathcal{U};b_{X})=\bigcup_{i\in X}\varphi(\mathcal{U};b_{i})

Work in some model \mathcal{B}. Let At(x)At(x) be the formula saying that xx is an atom and consider the formula

φ(x;y)=At(x)xy\varphi(x;y)=At(x)\land x\leq y

Fix n<ωn<\omega and choose (bi)i<n(b_{i})_{i<n} such that At(bi)At(b_{i}) and bibjb_{i}\neq b_{j} for i<j<ni<j<n. Now φ(x;bi)φ(x;bj)\varphi(x;b_{i})\land\varphi(x;b_{j}) is inconsistent for i<j<ni<j<n. Moreover, if XnX\subseteq n, let bX:=iXbib_{X}:=\bigcup_{i\in X}b_{i}; then

φ(;bX)={bi:bibXi<n}=iXφ(;bi)\varphi(\mathcal{B};b_{X})=\bigcup\{b_{i}:b_{i}\leq b_{X}\land i<n\}=\bigcup_{i\in X}\varphi(\mathcal{B};b_{i})

In conclusion φ(x;y)\varphi(x;y) has 1(1)1^{(1)} and thus TT is SMSM.

Non-example 4.13 (Atomless Boolean algebras).

Consider the theory TT of atomless Boolean algebras in the language ={,,¬,0,1}\mathcal{L}=\{\cup,\cap,\neg,0,1\}. In [7], Cooper proved that TT does not admit 11 and thus, by Proposition 4.10, it is not SMSM.

Example 4.14 (Skolem arithmetic).

Skolem arithmetic is the first-order theory of the structure (,)(\mathbb{N},\cdot). First of all, we observe that we can define the number 11, the divisibility relation and the set of prime numbers:

  • One(x)=y(xy=y)One(x)=\forall y(x\cdot y=y).

  • x|y=z(y=xz)x|y=\exists z(y=x\cdot z)

  • P(x)=¬One(x)y(y|x(One(y)y=x))P(x)=\neg One(x)\land\forall y(y|x\to(One(y)\lor y=x))

To show that this theory is SMSM we use Cooper’s characterization of SMSM showing a formula φ(x;y)\varphi(x;y) with 1(1)1^{(1)}. Define φ(x;y)=P(x)x|y\varphi(x;y)=P(x)\land x|y and enumerate the prime numbers P()={pi:i<ω}P(\mathbb{N})=\{p_{i}\ :\ i<\omega\}. Fix n<ωn<\omega and for each i<ni<n let bi=pib_{i}=p_{i}. Now, for i<j<ni<j<n, φ(,bi)φ(,bj)={pi}{pj}=\varphi(\mathbb{N},b_{i})\cap\varphi(\mathbb{N},b_{j})=\{p_{i}\}\cap\{p_{j}\}=\emptyset. Moreover, given XnX\subseteq n, take bX=iXpib_{X}=\prod_{i\in X}p_{i}. We have

φ(,bX)={pi:iX}=iX{pi}=iXφ(,bi)\varphi(\mathbb{N},b_{X})=\{p_{i}\ :\ i\in X\}=\bigcup_{i\in X}\{p_{i}\}=\bigcup_{i\in X}\varphi(\mathbb{N},b_{i})

In conclusion, Skolem arithmetic is SMSM.

A similar argument shows that the first-order theory of (,+,,0,1)(\mathbb{Z},+,\cdot,0,1) is SMSM.

Since all the previous examples of SMSM theories are not ω\omega-categorical, we now construct an ω\omega-categorical SMSM theory.

Example 4.15 (ω\omega-categorical SMSM theory).

Consider the language ={S,B,R}BA\mathcal{L}=\{S,B,R\}\cup\mathcal{L}_{BA} where BA={,,()c,0,1}\mathcal{L}_{BA}=\{\cap,\cup,(-)^{c},0,1\} is the language of Boolean algebras, SS and BB are unary predicates and RR is a binary relation. For each BA\mathcal{L}_{BA} term t(y¯)t(\bar{y}) define an \mathcal{L}-formula ψt(x,y¯)\psi_{t}(x,\bar{y}) as follows:

  • If tt is a variable yy then ψt(x,y)=R(x,y)\psi_{t}(x,y)=R(x,y).

  • if tt is 11 then ψt(x)\psi_{t}(x) is x=xx=x.

  • if t1(y¯)t_{1}(\bar{y}),t2(y¯)t_{2}(\bar{y}^{\prime}) are terms, then ψt1t2(x,y¯y¯)=ψt1(x,y¯)ψt2(x,y¯)\psi_{t_{1}\cap t_{2}}(x,\bar{y}\bar{y}^{\prime})=\psi_{t_{1}}(x,\bar{y})\land\psi_{t_{2}}(x,\bar{y}^{\prime}).

  • If t(y¯)=(t(y¯))ct(\bar{y})=(t^{\prime}(\bar{y}))^{c} then ψt(x,y¯)=¬ψt(x,y¯)\psi_{t}(x,\bar{y})=\neg\psi_{t^{\prime}}(x,\bar{y}).

Since 0 and \cup are definable from the other symbols, we can assume that the term tt does not contain them.

Now, given an atomic ψBA\psi\in\mathcal{L}_{BA}, ψ\psi is of the form t1(y¯)=t2(y¯)t_{1}(\bar{y})=t_{2}(\bar{y}^{\prime}). Define

ψ~(x,y¯y¯)=ψt1(x,y¯)ψt1(x,y¯)\tilde{\psi}(x,\bar{y}\bar{y}^{\prime})=\psi_{t_{1}}(x,\bar{y})\leftrightarrow\psi_{t_{1}}(x,\bar{y}^{\prime})

Consider the following \mathcal{L}-theory T0T_{0}:

  • (A1A_{1})

    SS and BB partition the universe.

  • (A2A_{2})

    RS×BR\subseteq S\times B.

  • (A3A_{3})

    BTBAB\models T_{BA} i.e. BB is a Boolean algebra.

  • (A4ψA_{4}^{\psi})

    For any ψ(y¯)BAat\psi(\bar{y})\in\mathcal{L}_{BA}^{at}, we have

    (y¯B)(ψ(y¯)(xS)ψ~(x,y¯))(\forall\bar{y}\in B)(\psi(\bar{y})\rightarrow(\forall x\in S)\tilde{\psi}(x,\bar{y}))

We claim that the class 𝒦:=Mod<ω(T0)\mathcal{K}:=Mod_{<\omega}(T_{0}) (i.e. the class of finite models of T0T_{0}) is a Fraïssé class and the theory of its limit is SMSM. Intuitively, each structure in 𝒦\mathcal{K} can be described as follows: we have a set SS, a Boolean algebra BB and a relation RR between the two sorts. The axiom scheme (A4ψ)ψBAat(A_{4}^{\psi})_{\psi\in\mathcal{L}_{BA}^{at}} makes sure that we have a homomorphism of Boolean algebras

h:B𝒫(S)h\colon B\to\mathcal{P}(S)
bR(S,b)b\mapsto R(S,b)

Thus, we have a realization of a homomorphic image of BB as RR-definable (over BB) subsets of SS.

To show that 𝒦\mathcal{K} is a Fraïssé class, we need to show that 𝒦\mathcal{K} has HPHP,JEPJEP,APAP and it is uniformly locally finite i.e. there exists f:ωωf\colon\omega\to\omega that bounds the size of structures in 𝒦\mathcal{K} with respect to the number of their generators. First of all, since T0T_{0} is universal, we have that 𝒦\mathcal{K} is closed under substructures (i.e. 𝒦\mathcal{K} has HPHP). Moreover, since the structure (,{0,1})(\emptyset,\{0,1\}) is an initial object in 𝒦\mathcal{K} (i.e. it embeds uniquely in every structure in 𝒦\mathcal{K}), we have that JEPJEP is implied by APAP.

Consider the function f:ωωf\colon\omega\to\omega defined as f(n)=n+22nf(n)=n+2^{2^{n}}. Now, given CA𝒦C\leq A\in\mathcal{K} with C=<Cgen>C=<C_{gen}>, Cgen={s0,,sl1,b0,,bk1}C_{gen}=\{s_{0},...,s_{l-1},b_{0},...,b_{k-1}\} and l+k=nl+k=n, it’s easy to see that |C|l+22kf(n)|C|\leq l+2^{2^{k}}\leq f(n). Thus 𝒦\mathcal{K} is uniformly locally finite. It remains to show that 𝒦\mathcal{K} has APAP.

Lemma 4.16.

Given A,A𝒦A,A^{\prime}\in\mathcal{K}, there is an embedding AAA\to A^{\prime} if and only if there is an embedding of Boolean algebras B(A)B(A)B(A)\to B(A^{\prime}) and a surjective homomorphism of Boolean algebras 𝒫(S(A))𝒫(S(A))\mathcal{P}(S(A^{\prime}))\to\mathcal{P}(S(A)) such that the following commutes

B(A){B(A)}𝒫(S(A)){\mathcal{P}(S(A))}B(A){B(A^{\prime})}𝒫(S(A)){\mathcal{P}(S(A^{\prime}))}hA\scriptstyle{h_{A}}emb\scriptstyle{emb}hA\scriptstyle{h_{A^{\prime}}}hom\scriptstyle{hom}

where hAh_{A} and hAh_{A^{\prime}} are the Boolean algebra homomorphisms given by bR(S,b)b\mapsto R(S,b).

Proof.

Suppose we have an embedding AAA\to A^{\prime}. In particular we have a embedding of Boolean algebras F:B(A)B(A)F\colon B(A)\to B(A^{\prime}) and an inclusion S(A)S(A)S(A)\subseteq S(A^{\prime}) such that for every sS(A)s\in S(A), R(s,b)R(s,b) if and only if R(s,F(b))R(s,F(b)). By taking XXS(A)X\mapsto X\cap S(A), we have an surjective homomorphism of Boolean algebras 𝒫(S(A))𝒫(S(A))\mathcal{P}(S(A^{\prime}))\twoheadrightarrow\mathcal{P}(S(A)). Therefore, given bB(A)b\in B(A), we have

hA(F(b))S(A)=R(S(A),F(b))S(A)=R(S(A),b)=hA(b)h_{A^{\prime}}(F(b))\cap S(A)=R(S(A^{\prime}),F(b))\cap S(A)=R(S(A),b)=h_{A}(b)

Conversely, given the above diagram, we have a Boolean algebra embedding B(A)B(A)B(A)\to B(A^{\prime}) and by Stone duality, we get an injection S(A)S(A)S(A)\to S(A^{\prime}). Commutativity of the diagram implies that the relation RR is preserved by this pair of maps and thus we have an embedding of \mathcal{L}-structures AAA\to A^{\prime}. ∎

We now prove that 𝒦\mathcal{K} has APAP. Consider the following amalgamation problem in 𝒦\mathcal{K}:

C{C}A{A}C{C^{\prime}}

By our previous lemma, this translates to the following diagram with commutative squares.

B(C){B(C)}𝒫(S(C)){\mathcal{P}(S(C))}B(A){B(A)}𝒫(S(A)){\mathcal{P}(S(A))}B(C){B(C^{\prime})}𝒫(S(C)){\mathcal{P}(S(C^{\prime}))}

Applying Stone duality, we get the following diagram in the category of finite sets, maintaining commutativity.

S1{S_{1}}S(C){S(C)}S0{S_{0}}S(A){S(A)}S2{S_{2}}S(C){S(C^{\prime})}g1\scriptstyle{g_{1}}f1\scriptstyle{f_{1}}i\scriptstyle{i}i\scriptstyle{i^{\prime}}g2\scriptstyle{g_{2}}f2\scriptstyle{f_{2}}

Performing the pullback and pushout (in the category of finite sets) we get

S1{S_{1}}S(C){S(C)}S0{S_{0}}S1×S0S2{S_{1}\times_{S_{0}}S_{2}}S(C)S(A)S(C){S(C)\sqcup_{S(A)}S(C^{\prime})}S(A){S(A)}S2{S_{2}}S(C){S(C^{\prime})\par}g1\scriptstyle{g_{1}}f1\scriptstyle{f_{1}}j\scriptstyle{j}p1\scriptstyle{p_{1}}p2\scriptstyle{p_{2}}i\scriptstyle{i}i\scriptstyle{i^{\prime}}g2\scriptstyle{g_{2}}f2\scriptstyle{f_{2}}j\scriptstyle{j^{\prime}}

Now, S1×S0S2={(a,b)S1×S2:g1(a)=g2(b)}S_{1}\times_{S_{0}}S_{2}=\{(a,b)\in S_{1}\times S_{2}\ :\ g_{1}(a)=g_{2}(b)\} and the functions pi:S1×S0S2Sip_{i}\colon S_{1}\times_{S_{0}}S_{2}\to S_{i} are the projections. Without loss of generality, we consider i,i,j,ji,i^{\prime},j,j^{\prime} inclusions of sets. We want to define a function

h:S(C)S(A)S(C)S1×S0S2h\colon S(C)\sqcup_{S(A)}S(C^{\prime})\to S_{1}\times_{S_{0}}S_{2}

such that f1=p1hjf_{1}=p_{1}\circ h\circ j and f2=p2hjf_{2}=p_{2}\circ h\circ j^{\prime}. Given xS(A)x\in S(A), define h(x)=(f1(x),f2(x))h(x)=(f_{1}(x),f_{2}(x)); since the outer paths are the same, (f1(x),f2(x))S1×S0S2(f_{1}(x),f_{2}(x))\in S_{1}\times_{S_{0}}S_{2}. Now, given xS(C)x\in S(C), since p1p_{1} is surjective, pick bS2b\in S_{2} such that (f1(x),b)S1×S0S2(f_{1}(x),b)\in S_{1}\times_{S_{0}}S_{2} and define h(x)=(f1(x),b)h(x)=(f_{1}(x),b). Similarly, for xS(C)x\in S(C^{\prime}), since p2p_{2} is surjective, define h(x)=(a,f2(x))h(x)=(a,f_{2}(x)) for some aS1a\in S_{1} such that (a,f2(x))S1×S0S2(a,f_{2}(x))\in S_{1}\times_{S_{0}}S_{2}. It is easy to check that f1=p1hjf_{1}=p_{1}\circ h\circ j and f2=p2hjf_{2}=p_{2}\circ h\circ j^{\prime}.

Dually, we get a Boolean algebra B(D)=𝒫(S1×S0S2)B(D)=\mathcal{P}(S_{1}\times_{S_{0}}S_{2}) that solves the amalgamation problem B(A)B(C)B(A)\to B(C) and B(A)B(C)B(A)\to B(C^{\prime}) and a set S(D)S(D) with a homomorphism B(D)𝒫(S(D))B(D)\to\mathcal{P}(S(D)). Using the lemma and commutativity of squares, we get a solution (B(D),S(D))(B(D),S(D)) to the initial amalgamation problem. In conclusion, 𝒦\mathcal{K} has APAP.

Let TT be the first-order theory of the Fraïssé limit of 𝒦\mathcal{K}. We show that TT is SMSM. By Fact 4.7 and Proposition 4.10, it is enough to show that TT has 1(1)1^{(1)}. Let φ(x;y)=R(x,y)\varphi(x;y)=R(x,y). Fix n<ωn<\omega and consider the finite model AA of T0T_{0} given by B(A)=𝒫(n)B(A)=\mathcal{P}(n) and S(B)=nS(B)=n with R(i,X)R(i,X) if and only if iXi\in X. Now, for each i<ni<n, define bi=b{i}b_{i}=b_{\{i\}}. Clearly, R(A,bi)R(A,bj)=R(n,bi)R(n,bj)={i}{j}=R(A,b_{i})\cap R(A,b_{j})=R(n,b_{i})\cap R(n,b_{j})=\{i\}\cap\{j\}=\emptyset for all i<j<ni<j<n. Moreover, given XnX\subseteq n, we have that

R(A,bX)=R(n,bX)=X=iX{i}=iXR(A,bi)R(A,b_{X})=R(n,b_{X})=X=\bigcup_{i\in X}\{i\}=\bigcup_{i\in X}R(A,b_{i})

Embedding AA into \mathcal{M} (by universality of \mathcal{M}), the above properties are preserved. Thus R(x,y)R(x,y) is SMSM in TT.

We can give an explicit description of the Fraïssé limit as follows: Let =(ω,)\mathcal{M}=(\omega,\mathcal{B}) where \mathcal{B} is the countable atomless Boolean algebra. By Stone duality, we know that there is an isomorphism of Boolean algebras 𝑔CLOP(2ω)\mathcal{B}\xrightarrow{g}CLOP(2^{\omega}). Moreover, let f:ω2ωf\colon\omega\to 2^{\omega} such that the image of ff is a countable dense subset of the Cantor space and define R(x,y)R(x,y) as follows:

R(n,b) iff f(n)g(b)\mathcal{M}\models R(n,b)\text{ iff }f(n)\in g(b)

Note that, since 2ω2^{\omega} is totally disconnected, for every nmn\neq m in ω\omega there exists bnmb_{n}^{m}\in\mathcal{B} such that R(n,b)¬R(m,b)\mathcal{M}\models R(n,b)\land\neg R(m,b). Using this, one can check that the extension properties hold in \mathcal{M} and thus show that it is the desired Fraïssé limit.

5. Weakenings of SMSM

Once established the worst (in some sense) combinatorial binary property that a complete first-order theory can have, we identify subcollections of the set of exhibitable patterns which are enough to describe some of the properties of formulas previously considered. From that, we define other notions of maximality realizing the patterns in those subcollections.

5.1. Positive maximality

The tree properties defining simple, NTP1NTP_{1} and NTP2NTP_{2} theories can be described by patterns in which the negation of the formula is not involved. More precisely, we can characterize tree properties by exhibiting patterns in which, for every condition (X+,X)𝒞(X^{+},X^{-})\in\mathcal{C}\cup\mathcal{I}, the set XX^{-} is empty (see Proposition 3.3). This observation leads the way for the concepts and results described in this subsection.

Definition 5.1.

Fix n<ωn<\omega. A positive nn-pattern is a pattern (𝒞,)(\mathcal{C},\mathcal{I}) such that for every condition (X+,X)𝒞(X^{+},X^{-})\in\mathcal{C}\cup\mathcal{I}, we have that X=X^{-}=\emptyset.

We can now define the notion of positive maximality (PMPM) by requesting the exhibition of all reasonable (See Definition 3.5) positive patterns.

Definition 5.2.

A formula φ(x;y)\varphi(x;y) is positive maximal (PMPM), if for every n<ωn<\omega and every reasonable positive nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}), the formula φ\varphi exhibits (𝒞,)(\mathcal{C},\mathcal{I}).

Remark 5.3.

Notice that we didn’t require (as in SMSM) to realize every exhibitable pattern since every reasonable positive pattern is exhibitable. Let’s show this by giving an example of a PMPM theory. Let T=Th()T=Th(\mathcal{B}) where \mathcal{B} is an infinite Boolean algebra with an atomless element. We claim that the formula φ(x;y)=0xy\varphi(x;y)=0\neq x\leq y is PMPM. The existence of an atomless element implies the existence of a sequence (ai)i<ω(a_{i})_{i<\omega} of non-zero elements below the atomless element such that aiaj=0a_{i}\cap a_{j}=0 for every i<j<ωi<j<\omega. Fix n<ωn<\omega and fix a reasonable positive nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}). If 𝒞=\mathcal{C}=\emptyset, take bi=0b_{i}=0 for every i<ni<n; since φ(x;bi)=\varphi(x;b_{i})=\bot for every ii, we get that every inconsistency condition will be satisfied. Assume 𝒞={(Yj,):j<k}\emptyset\neq\mathcal{C}=\{(Y_{j},\emptyset)\ :\ j<k\}. For each i<ni<n, define

bi=iYjajb_{i}=\bigcup_{i\in Y_{j}}a_{j}

Fix j~<k\tilde{j}<k. Then, if iYj~i\in Y_{\tilde{j}} we have that

0aj~iYjaj=bi0\neq a_{\tilde{j}}\leq\bigcup_{i\in Y_{j}}a_{j}=b_{i}

hence {φ(x;bi):iYj¯}\{\varphi(x;b_{i})\ :\ i\in Y_{\bar{j}}\} is consistent. Fix now (Z,)(Z,\emptyset)\in\mathcal{I} and suppose, toward contradiction, that there is aa\in\mathcal{B} such that for every iZi\in Z, aφ(x;bi)a\models\varphi(x;b_{i}). This means that

0aiZiYjaj=f:Zk,(iZ)(iYf(i))iZaf(i)0\neq a\leq\bigcap_{i\in Z}\bigcup_{i\in Y_{j}}a_{j}=\bigcup_{\begin{subarray}{c}f\colon Z\to k,\\ (\forall i\in Z)(i\in Y_{f(i)})\end{subarray}}\bigcap_{i\in Z}a_{f(i)}

If the above union is 00\in\mathcal{B}, then 0a00\neq a\leq 0; contradiction. If it is not empty, this means that there is f:Zkf\colon Z\to k, with for all iZ,iYf(i)i\in Z,i\in Y_{f(i)}, such that iZaf(i)0\bigcap_{i\in Z}a_{f(i)}\neq 0. Since all the aja_{j} are pairwise disjoint, this can only happen if ff is constant, hence there is a j<kj<k such that for all iZi\in Z, aja_{j} appears in the above union i.e. there is a j<kj<k such that for all iZi\in Z, iYji\in Y_{j}; thus ZYjZ\subseteq Y_{j} contradicting reasonableness. In conclusion, this theory is PMPM (despite being NSMNSM, see Remark 4.13) and hence every reasonable positive nn-pattern is exhibitable.

Remark 5.4.

Since every reasonable positive pattern is, in particular, an exhibitable pattern, if a formula φ(x;y)\varphi(x;y) is SMSM, then φ(x;y)\varphi(x;y) is also PMPM.

We now prove a characterization of PMPM at the level of formulas. This equivalence will give us a concrete subfamily of the collection of all reasonable positive patterns which is enough to exhibit positive maximality.

Proposition 5.5.

A formula φ(x;y)\varphi(x;y) is PMPM if and only if, for every n<ωn<\omega, there is (bX)Xn(b_{X})_{X\subseteq n} such that, for any family of subsets of nn, Z𝒫(𝒫(n))\emptyset\neq Z\in\mathcal{P}(\mathcal{P}(n)), the partial φ\varphi-type

{φ(x;bY):YZ} is consistent if and only if Z\{\varphi(x;b_{Y})\ :\ Y\in Z\}\text{ is consistent if and only if }\bigcap Z\neq\emptyset
Proof.

Assume φ(x;y)\varphi(x;y) is PMPM. Fix n<ωn<\omega and consider the following reasonable positive 2n2^{n}-pattern

𝒞={(Z,):Z2n such that Z}\mathcal{C}=\{(Z,\emptyset)\ :\ Z\subseteq 2^{n}\text{ such that }\bigcap Z\neq\emptyset\}
={(Z,):Z2n such that Z=}\mathcal{I}=\{(Z,\emptyset)\ :\ Z\subseteq 2^{n}\text{ such that }\bigcap Z=\emptyset\}

where we identify each Z2nZ\subseteq 2^{n} as a family of subsets of nn. By PMPM, there is (bi)i<2n=(bX)Xn(b_{i})_{i<2^{n}}=(b_{X})_{X\subseteq n} such that, for every (Z,)𝒞(Z,\emptyset)\in\mathcal{C}, i.e. for every Z2nZ\subseteq 2^{n} such that Z\bigcap Z\neq\emptyset, the type {φ(x;bi):iZ}={φ(x;bY):YZ}\{\varphi(x;b_{i})\ :\ i\in Z\}=\{\varphi(x;b_{Y})\ :\ Y\in Z\} is consistent and, for every (Z,)(Z,\emptyset)\in\mathcal{I} i.e. for every Z2nZ\subseteq 2^{n} such that Z=\bigcap Z=\emptyset, the type {φ(x;bY):YZ}\{\varphi(x;b_{Y})\ :\ Y\in Z\} is inconsistent.

Assume now that φ(x;y)\varphi(x;y) has the above property; fix n<ωn<\omega and let (𝒞,)(\mathcal{C},\mathcal{I}) be a reasonable positive nn-pattern. By the above property, let (bX)Xn(b_{X})_{X\subseteq n} as above. We can assume that 𝒞\mathcal{C}\neq\emptyset (otherwise, take bi=bb_{i}=b_{\emptyset} for all i<ni<n). Hence let 𝒞={(Yi,):i<k}\mathcal{C}=\{(Y_{i},\emptyset)\ :\ i<k\}. For any i<ni<n define

bi=b{j<k:iYj}b_{i}=b_{\{j<k\ :\ i\in Y_{j}\}}

Fix (Yj~,)𝒞(Y_{\tilde{j}},\emptyset)\in\mathcal{C}. To show that {φ(x;bi):iYj~}\{\varphi(x;b_{i})\ :\ i\in Y_{\tilde{j}}\} is consistent, by the above property, it is enough to show that

iYj~{j<k:iYj}\bigcap_{i\in Y_{\tilde{j}}}\{j<k\ :\ i\in Y_{j}\}\neq\emptyset

For every iYj~i\in Y_{\tilde{j}}, j~\tilde{j} is one of the j<kj<k such that iYj~i\in Y_{\tilde{j}}, thus j~iYj~{j<k:iYj}\tilde{j}\in\bigcap_{i\in Y_{\tilde{j}}}\{j<k\ :\ i\in Y_{j}\}.

Assume now that (Z,)(Z,\emptyset)\in\mathcal{I} and that, toward contradiction, there is a𝒰xa\in\mathcal{U}^{x} such that for all iZi\in Z, we have that aφ(x;bi)a\models\varphi(x;b_{i}) i.e.

{φ(x;bi):iZ}={φ(x;b{j<k:iYj}):iZ} is consistent.\{\varphi(x;b_{i})\ :\ i\in Z\}=\{\varphi(x;b_{\{j<k\ :\ i\in Y_{j}\}})\ :\ i\in Z\}\text{ is consistent.}

By the above property, this is equivalent to

iZ{j<k:iYj}\emptyset\neq\bigcap_{i\in Z}\{j<k\ :\ i\in Y_{j}\}

Take j<kj<k in the above intersection. Then, for every iZi\in Z, iYji\in Y_{j} contradicting reasonability of the pattern. ∎

Remark 5.6.

In [7], Cooper defined a property of formulas (named vp) that closely resembles the above characterization of PMPM. It is easy to check that, using Proposition 5.5, PMPM is equivalent to Cooper’s vp, even at the level of formulas.

As pointed out above, some dividing lines can be defined through realizations of reasonable positive patterns. In particular, TP2TP_{2} is given by positive patterns and hence PMPM implies TP2TP_{2} and a fortiori IPIP and OPOP. As described above, SOPSOP is not given by positive patterns; hence, a natural question is the following: Is there a NSOPNSOP and PMPM theory?

We answer positively to the above question showing an example:

Example 5.7 (A positive maximal NSOP(4)NSOP_{(4)} theory).

Let ={P,O,R,(Ek)0<k<ω}\mathcal{L}=\{P,O,R,(E_{k})_{0<k<\omega}\} where ar(P)=ar(O)=1ar(P)=ar(O)=1, ar(R)=2ar(R)=2 and for any k<ωk<\omega, ar(Ek)=kar(E_{k})=k.
Let T0T_{0} be the following \mathcal{L}-theory:

  • (A1)(A_{1})

    P,OP,O partition the universe.

  • (A2k)(A_{2}^{k})

    EkOkE_{k}\subseteq O^{k}

  • (A3)(A_{3})

    RP×OR\subseteq P\times O

  • (A4k)(A_{4}^{k})

    y0yk1x(Ek(y¯)¬i<kR(x,yi))\forall y_{0}...y_{k-1}\forall x(E_{k}(\bar{y})\to\neg\bigwedge_{i<k}R(x,y_{i}))

Let 𝒦={A:AT0 and A is finite}\mathcal{K}=\{A\ :\ A\models T_{0}\text{ and $A$ is finite}\}. We claim that 𝒦\mathcal{K} is a Fraïssé class with free amalgamation.

Since T0T_{0} is \forall-axiomatized, the class of models of T0T_{0} (in particular the class of its finite models) is closed under taking substructures. Thus 𝒦\mathcal{K} has HPHP. Since we allow the empty structure to be a model of T0T_{0}, it’s enough to prove APAP to get JEPJEP for free.

Proposition 5.8.

The class 𝒦\mathcal{K} has the amalgamation property.

Proof.

Consider the amalgamation problem (A,Bi,fi)i<2(A,B_{i},f_{i})_{i<2} in 𝒰\mathcal{U} where fi:ABif_{i}\colon A\to B_{i} are embeddings.

We define an \mathcal{L}-structure CC as follows: The universe of CC is partitioned by PC=(PB0PA)PB1P^{C}=(P^{B_{0}}\setminus P^{A})\sqcup P^{B_{1}} and OC=(OB0OA)OB1O^{C}=(O^{B_{0}}\setminus O^{A})\sqcup O^{B_{1}}. The interpretation of RR and EkE_{k} is the induced one (with no other relations holding). In particular there is no relation EkE_{k} holding between elements of OB0O^{B_{0}} and OB1O^{B_{1}} inside CC.
Clearly CC solves the amalgamation problem. We need to show that CT0C\models T_{0}.
Clearly CC satisfies the axioms (and axiom schemes) A1A_{1}, (A2k)k<ω(A_{2}^{k})_{k<\omega} and (A3k)k<ω(A_{3}^{k})_{k<\omega}. It remains to show that, for each k<ωk<\omega, C(A4k)C\models(A_{4}^{k}). So, fix k<ωk<\omega, fix c0ck1OCc_{0}...c_{k-1}\in O^{C} and aPCa\in P^{C}. Moreover, assume that Ek(c¯)E_{k}(\bar{c}) holds. We need to show that a¬i<kR(x;ci)a\models\neg\bigwedge_{i<k}R(x;c_{i}). Since EkE_{k} is induced, we have that either c¯(OB0)k\bar{c}\in(O^{B_{0}})^{k} or c¯(OB1)k\bar{c}\in(O^{B_{1}})^{k}, but not both. Without loss of generality, c¯(OB1)k\bar{c}\in(O^{B_{1}})^{k}. Now, either aPB0PAa\in P^{B_{0}}\setminus P^{A} xor aPB1a\in P^{B_{1}}. In the first case, since c¯PB1\bar{c}\in P^{B_{1}} and RR is the induced one, we have that ¬R(a,ci)\models\neg R(a,c_{i}) for every i<ki<k and hence a¬i<kR(x;ci)a\models\neg\bigwedge_{i<k}R(x;c_{i}). In the second case, since B1T0B_{1}\models T_{0}, a¬i<kR(x;ci)a\models\neg\bigwedge_{i<k}R(x;c_{i}). In conclusion C(A4k)C\models(A_{4}^{k}) and hence 𝒦\mathcal{K} has APAP. ∎

Note that, in the above proof, CC is the free amalgam of B0B_{0} and B1B_{1} over AA. Hence 𝒦\mathcal{K} is closed under free amalgamation (We refer to [6], Definition 3.1 and Example 3.2).

Since 𝒦\mathcal{K} is a Fraïssé class, let \mathcal{M} be its limit and let T=Th()T_{\mathcal{M}}=Th(\mathcal{M}). Since 𝒦\mathcal{K} is closed under free amalgamation, we have that TT_{\mathcal{M}} is NSOP4NSOP_{4} (See [6], Theorem 1.1) and in particular NSOPNSOP.

Proposition 5.9.

TT_{\mathcal{M}} is PMPM.

Proof.

We claim that the formula R(x;y)R(x;y) is PMPM. Let n<ωn<\omega and fix 𝒫=(𝒞,)\mathcal{P}=(\mathcal{C},\mathcal{I}) a reasonable positive nn-pattern, where

𝒞={(Y0,),,(Yl1,)} and ={(Z0,),,(Zh1,)}\mathcal{C}=\{(Y_{0},\emptyset),...,(Y_{l-1},\emptyset)\}\text{ and }\mathcal{I}=\{(Z_{0},\emptyset),...,(Z_{h-1},\emptyset)\}

Define the structure A𝒫A_{\mathcal{P}} as follows: The universe of A𝒫A_{\mathcal{P}} is given by PA𝒫OA𝒫={ai:i<l}{bi:i<n}P^{A_{\mathcal{P}}}\cup O^{A_{\mathcal{P}}}=\{a_{i}\ :\ i<l\}\cup\{b_{i}\ :\ i<n\}. We interpret RR by R(ai;bj)R(a_{i};b_{j}) holds if and only if iYji\in Y_{j} and, for k<ωk<\omega, we define EkE_{k} as follows: Ek(bi0,,bik1)E_{k}(b_{i_{0}},...,b_{i_{k-1}}) holds if and only if k=|Zj|k=|Z_{j}| for some j<hj<h and {i0,,ik1}=Zj\{i_{0},...,i_{k-1}\}=Z_{j}. First of all, by definition of reasonable pattern (see Definition 3.5) there is no (Z,)(Z,\emptyset)\in\mathcal{I} such that (Z,)(Y,)(Z,\emptyset)\subseteq(Y,\emptyset) (coordinatewise) for some (Y,)𝒞(Y,\emptyset)\in\mathcal{C}. This ensures that the witness for {R(x,bi):iY}\{R(x,b_{i})\ :\ i\in Y\} doesn’t conflict with the relation EkE_{k} for k<|Y|k<|Y| and the axiom (A4k)(A_{4}^{k}). Thus, we have that A𝒫T0A_{\mathcal{P}}\models T_{0}. Hence there is an embedding of f:A𝒫f\colon A_{\mathcal{P}}\to\mathcal{M}. Let B={f(bi):i<n}B=\{f(b_{i})\ :\ i<n\}. It’s clear that R(x;y)R(x;y) exhibits the pattern 𝒫\mathcal{P} as witnessed by BB. ∎

As we showed, this example is actually a NSOP4NSOP_{4} and PMPM theory. This means that no SOPnSOP_{n}, n4n\geq 4, can be described by positive patterns. It is still open whether SOPnSOP_{n}, for n4n\geq 4, can be described through (non-positive) patterns (even though, since SMSM implies SOPSOP which in turn implies SOPnSOP_{n}, we have SMSM implies every SOPnSOP_{n}).

Again, we remark that SOP3SOP_{3} was characterized by Kaplan, Ramsey and Simon ([9], Lemma 7.3) in a way that can be described using positive patterns.

5.2. Consistency maximality

While tree properties can be characterized by positive patterns, some properties of formulas, such as OPOP and IPIP, do not need any inconsistency conditions i.e. the patterns (𝒞,)(\mathcal{C},\mathcal{I}) needed to get these properties have =\mathcal{I}=\emptyset (see Proposition 3.3). It is thus natural to define the following property.

Definition 5.10.

We say that a formula φ(x;y)\varphi(x;y) is consistency maximal (CMCM) if, for every n<ωn<\omega and every reasonable (consistency) nn-pattern (𝒞,)(\mathcal{C},\emptyset), φ\varphi exhibits (𝒞,)(\mathcal{C},\emptyset).

Remark 5.11.

Note that we didn’t ask for a CMCM formula to realize every exhibitable pattern of the form (𝒞,)(\mathcal{C},\emptyset) since every such pattern is exhibitable. To see this we give an example of a CMCM formula.

Consider the theory TT of the random graph. Recall that this is a countable graph characterized (up to isomorphism) by the following property:

For any A,BA,B finite disjoint subsets of vertices, there exists a vertex vv such that

E(v,a) for every aA\models E(v,a)\text{ for every $a\in A$}
¬E(v,b) for every bB \models\neg E(v,b)\text{ for every $b\in B$ }

where EE is the edge relation.

Let φ(x;y)=E(x;y)\varphi(x;y)=E(x;y), fix n<ωn<\omega, a reasonable nn-pattern (𝒞,)(\mathcal{C},\emptyset) and nn distinct vertices b0,,bn1b_{0},...,b_{n-1}. Then, for every condition (A+,A)𝒞(A^{+},A^{-})\in\mathcal{C}, by the above property characterizing the random graph, there is vv such that

v{φ(x;bi):iA+}{¬φ(x;bi):iA}v\models\{\varphi(x;b_{i})\ :\ i\in A^{+}\}\cup\{\neg\varphi(x;b_{i})\ :\ i\in A^{-}\}

Hence φ(x;y)\varphi(x;y) exhibits (𝒞,)(\mathcal{C},\emptyset). In conclusion φ(x;y)\varphi(x;y) is CMCM.

We prove that CMCM is a weakening of PMPM.

Proposition 5.12.

If a formula φ(x;y)\varphi(x;y) is PMPM then φ(x;y)\varphi(x;y) is CMCM.

Proof.

Let n<ωn<\omega and let (𝒞,)(\mathcal{C},\emptyset) be a reasonable nn-pattern. Define the positive 2n2n-pattern

𝒞:={(A+(n+A),):(A+,A)𝒞}\mathcal{C}^{\prime}:=\{(A^{+}\cup(n+A^{-}),\emptyset)\ :\ (A^{+},A^{-})\in\mathcal{C}\}
={({i,i+n},):i<n}\mathcal{I}^{\prime}=\{(\{i,i+n\},\emptyset)\ :\ i<n\}

where n+A={n+i:iA}n+A^{-}=\{n+i\ :\ i\in A^{-}\}. Note that, since A+A=A^{+}\cap A^{-}=\emptyset, the positive pattern (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}) is reasonable. Since φ(x;y)\varphi(x;y) is positive maximal, there are b0,,bn1,bn,,b2n1b_{0},...,b_{n-1},b_{n},...,b_{2n-1} witnessing that φ\varphi exhibits (𝒞,)(\mathcal{C}^{\prime},\mathcal{I}^{\prime}). We claim that b0,,bn1b_{0},...,b_{n-1} witness that φ\varphi exhibits (𝒞,)(\mathcal{C},\emptyset). Let (A+,A)𝒞(A^{+},A^{-})\in\mathcal{C}. Then

{φ(x;bi):iA+(n+A)} is consistent.\{\varphi(x;b_{i})\ :\ i\in A^{+}\cup(n+A^{-})\}\text{ is consistent.}

Let aa a realization of it and let iAi\in A^{-}. If, toward contradiction, aφ(x;bi)a\models\varphi(x;b_{i}), then, by the choice of aa, aφ(x;bi)φ(x;bn+i)a\models\varphi(x;b_{i})\land\varphi(x;b_{n+i}) which is inconsistent. Hence, for each iAi\in A^{-}, a¬φ(x;bi)a\models\neg\varphi(x;b_{i}). This shows that

{φ(x;bi):iA+}{¬φ(x;bj):jA} is consistent.\{\varphi(x;b_{i})\ :\ i\in A^{+}\}\cup\{\neg\varphi(x;b_{j})\ :\ j\in A^{-}\}\text{ is consistent.}

Hence, φ\varphi exhibits (𝒞,)(\mathcal{C},\emptyset). In conclusion φ\varphi is CMCM. ∎

We now give an example of a NPMNPM but CMCM theory, showing that the implication in Proposition 5.12 is strict.

Example 5.13.

Again, consider the theory TT of the random graph. It is well known that this theory is simple (NTPNTP); moreover we proved that it is CMCM. If TT was PMPM, then, since TPTP can be realized through patterns, TT would have TPTP, contradicting the simplicity of the theory.

We now prove that this weaker notion of maximality (CMCM) is actually equivalent to the Independence Property (IP)(IP). Recall that a formula φ(x;y)\varphi(x;y) has IPIP if, by definition, there are tuples (ai)i<ω(a_{i})_{i<\omega} and (bI)Iω(b_{I})_{I\subseteq\omega} such that

φ(ai;bI)iI\models\varphi(a_{i};b_{I})\iff i\in I

By compactness, this is equivalent to having, for every n<ωn<\omega, a nnth approximation of the above property.

Proposition 5.14.

A formula φ(x;y)\varphi(x;y) is CMCM if and only if it has IPIP

Proof.

We will actually prove that φ(x;y)\varphi(x;y) is CMCM if and only if φopp(y;x)\varphi^{opp}(y;x) has IPIP. On the other hand, it’s well-known that a formula has IPIP if and only if the opposite formula has IPIP (See, e.g., [1]).

Suppose φ(x;y)\varphi(x;y) is CMCM. Fix n<ωn<\omega and consider the pattern

𝒞={(X,nX):X𝒫(n)}\mathcal{C}=\{(X,n\setminus X)\ :\ X\in\mathcal{P}(n)\}
=\mathcal{I}=\emptyset

Since this is a consistency pattern, φ(x,y)\varphi(x,y) exhibits it; let B=(bi)i<nB=(b_{i})_{i<n} be a witness. For any X𝒫(n)=𝒞X\in\mathcal{P}(n)=\mathcal{C}, let aXa_{X} a realization of pX,nXB(x)p_{X,n\setminus X}^{B}(x). Then,

φopp(bi,aX)φ(aX,bi)iX\models\varphi^{opp}(b_{i},a_{X})\iff\models\varphi(a_{X},b_{i})\iff i\in X

Hence, for every n<ωn<\omega, φopp(y;x)\varphi^{opp}(y;x) has a nnth approximation of IPIP. By compactness, φopp(y;x)\varphi^{opp}(y;x) has IPIP.

Assume now that φopp(y;x)\varphi^{opp}(y;x) has IPIP. Take (bi)i<ω(b_{i})_{i<\omega} and (aI)Iω(a_{I})_{I\subseteq\omega} witnessing it. Fix a reasonable nn-pattern (𝒞,)(\mathcal{C},\emptyset). Then, for any condition (A+,A)𝒞(A^{+},A^{-})\in\mathcal{C}, we have that

aA+{φopp(bi;x):iA+}{φopp(bi;x):iA+,i<n}a_{A^{+}}\models\{\varphi^{opp}(b_{i};x)\ :\ i\in A^{+}\}\cup\{\varphi^{opp}(b_{i};x)\ :\ i\notin A^{+},i<n\}

Since AnA+A^{-}\subseteq n\setminus A^{+}, we have

{φopp(bi;x):iA+}{φopp(bi;x):iA+}\{\varphi^{opp}(b_{i};x)\ :\ i\in A^{+}\}\cup\{\varphi^{opp}(b_{i};x)\ :\ i\notin A^{+}\}\supseteq
{φopp(bi;x):iA+}{φopp(bi;x):iA}\{\varphi^{opp}(b_{i};x)\ :\ i\in A^{+}\}\cup\{\varphi^{opp}(b_{i};x)\ :\ i\in A^{-}\}

Which in turn equals to

{φ(x;bi):iA+}{φ(x;bi):iA}\{\varphi(x;b_{i})\ :\ i\in A^{+}\}\cup\{\varphi(x;b_{i})\ :\ i\in A^{-}\}

Hence pA+,AB(x)p_{A^{+},A^{-}}^{B}(x) is consistent and hence φ(x;y)\varphi(x;y) is CMCM. ∎

Remark 5.15.

Given the above equivalence, we have an alternative proof of Proposition 5.12:

Since TP2TP_{2} can be realized by positive patterns, if φ(x;y)\varphi(x;y) is PMPM, then φ(x;y)\varphi(x;y) has TP2TP_{2} which in turn implies IPIP (See Fact 2.2) which, by Proposition 5.14, is equivalent to CMCM.

6. The PMkPM_{k} hierarchy

Natural weakenings of PMPM can be obtained by restricting the size of possible sets of inconsistencies. Some combinatorial properties that can be defined in terms of inconsistency sets of size >2>2, such as kk-TPTP and kk-TP2TP_{2}, turn out to be equivalent, at the level of theories, to their counterparts with inconsistency sets of size 22 (22-TPTP and 22-TP2TP_{2}). In contrast, by weakening PMPM restricting the size of possible sets of inconsistencies, we get a strict hierarchy of properties (PMk)1<k<ω(PM_{k})_{1<k<\omega}.

Definition 6.1.

Let 1<k<ω1<k<\omega. A formula φ(x;y)\varphi(x;y) is PMkPM_{k} if for all n<ωn<\omega and for every reasonable positive nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) with for all (Z+,)(Z^{+},\emptyset)\in\mathcal{I}, |Z+|=k|Z^{+}|=k, φ(x;y)\varphi(x;y) exhibits (𝒞,)(\mathcal{C},\mathcal{I}). We say that a theory TT is PMkPM_{k} if there is a PMkPM_{k} formula. Otherwise we say TT is NPMkNPM_{k}.

Remark 6.2.

We don’t consider PM0PM_{0} and PM1PM_{1} since (almost) every theory is PM0PM_{0} and PM1PM_{1}. For PM0PM_{0}, fix a tuple a𝒰xa\in\mathcal{U}^{x} and take the formula

φ(x;y)=(x=y)\varphi(x;y)=(x=y)

Since we excluded conditions of the form (,)(\emptyset,\emptyset) from the definition of patterns, =\mathcal{I}=\emptyset and thus φ(x;y)\varphi(x;y) is asked to exhibit positive patterns of the form (𝒞,)(\mathcal{C},\emptyset). Fix n<ωn<\omega and an nn-pattern as above. Define bi=ab_{i}=a for all i<ni<n; then, for any condition (A+,)𝒞(A^{+},\emptyset)\in\mathcal{C}

iA+(x=a) is consistent, witnessed by a\bigwedge_{i\in A^{+}}(x=a)\text{ is consistent, witnessed by $a$}

For PM1PM_{1}, fix aba\neq b and consider the formula

φ(x;y)=(y=a)(x=x)\varphi(x;y)=(y=a)\land(x=x)

Fix n<ωn<\omega and a reasonable positive nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) with for all (Z+,)(Z^{+},\emptyset), |Z+|=1|Z^{+}|=1. For all ii such that there is a condition ({i},)(\{i\},\emptyset)\in\mathcal{I}, define bi=bb_{i}=b. Otherwise, define bi=ab_{i}=a. Then, for (A+,)𝒞(A^{+},\emptyset)\in\mathcal{C}, since pattern are reasonable, for all iA+i\in A^{+}, bi=ab_{i}=a and thus

i<n(a=a)(x=x) is consistent.\bigwedge_{i<n}(a=a)\land(x=x)\text{ is consistent.}

For ({i},)(\{i\},\emptyset)\in\mathcal{I}, bi=bab_{i}=b\neq a and thus

φ(x;bi)=(a=b)(x=x) is inconsistent.\varphi(x;b_{i})=(a=b)\land(x=x)\text{ is inconsistent.}

In conclusion, we disregard PM0PM_{0} and PM1PM_{1} since they are not very sensible notions.

We can characterize the PMkPM_{k} properties by “realizations of hypergraphs”:

Proposition 6.3.

φ(x;y)\varphi(x;y) is PMkPM_{k} if and only if for every finite kk-hypergraph (n,E)(n,E), φ\varphi realizes (n,E)(n,E) i.e. there are (bi)i<n(b_{i})_{i<n} such that

  • (i)

    for any i0,,ik1ni_{0},...,i_{k-1}\in n with ¬E(i0,,ik1)\neg E(i_{0},...,i_{k-1}), we have

    {φ(x;bij):j<k} is inconsistent\{\varphi(x;b_{i_{j}})\ :\ j<k\}\text{ is inconsistent}

    and

  • (ii)

    for any clique {i0,,il1}n\{i_{0},...,i_{l-1}\}\subseteq n

    {φ(x;bij):j<l} is consistent.\{\varphi(x;b_{i_{j}})\ :\ j<l\}\text{ is consistent.}
Proof.

Suppose φ(x;y)\varphi(x;y) is PMkPM_{k}. Let (n,E)(n,E) a finite kk-hypergraph. Consider the positive nn-pattern (𝒞,)(\mathcal{C},\mathcal{I}) where

𝒞:={({i0,,il1},):{i0,,il1} is a clique }\mathcal{C}:=\{(\{i_{0},...,i_{l-1}\},\emptyset)\ :\ \{i_{0},...,i_{l-1}\}\text{ is a clique }\}
:={({i0,,ik1},):¬E(i0,,ik1)}\mathcal{I}:=\{(\{i_{0},...,i_{k-1}\},\emptyset)\ :\ \neg E(i_{0},...,i_{k-1})\}

Since no non-hyperedge can be a subset of a clique, this is a reasonable pattern. By PMkPM_{k}, φ\varphi exhibits (𝒞,)(\mathcal{C},\mathcal{I}) and thus there are (bi)i<n(b_{i})_{i<n} witnessing it. It’s easy to see that, with these parameters, φ\varphi realizes (n,E)(n,E).

Assume now that φ(x;y)\varphi(x;y) realizes every kk-hypergraph. Let n<ωn<\omega and let (𝒞,)(\mathcal{C},\mathcal{I}) a positive nn-pattern with for all (Z+,)(Z^{+},\emptyset)\in\mathcal{I}, |Z+|=k|Z^{+}|=k. Consider the graph (n,E)(n,E) where, E(i0,,ik1)E(i_{0},...,i_{k-1}) holds if and only if ({i0,,ik1},)(\{i_{0},...,i_{k-1}\},\emptyset)\notin\mathcal{I}. Since φ(x;y)\varphi(x;y) realizes every finite kk-hypergraph, there are (bi)i<n(b_{i})_{i<n} witnessing it. Now, given (A+,)𝒞(A^{+},\emptyset)\in\mathcal{C}, since the pattern is reasonable, every kk-element subset BB of A+A^{+} is such that (B,)(B,\emptyset)\notin\mathcal{I}, thus A+A^{+} forms a clique in (n,E)(n,E) and therefore

{φ(x;bi):iA+} is consistent.\{\varphi(x;b_{i})\ :\ i\in A^{+}\}\text{ is consistent.}

On the other hand, given ({i0,,ik1},)(\{i_{0},...,i_{k-1}\},\emptyset)\in\mathcal{I}, we have ¬E(i0,,ik1)\neg E(i_{0},...,i_{k-1}) and thus

{φ(x;bij):j<k} is inconsistent.\{\varphi(x;b_{i_{j}})\ :\ j<k\}\text{ is inconsistent.}

Thus φ\varphi exhibits (𝒞,)(\mathcal{C},\mathcal{I}). ∎

Proposition 6.4.

If φ(x;y)\varphi(x;y) is PMkPM_{k}, then φ(x;y)\varphi(x;y) realizes the generic kk-hypergraph. Moreover, the parameters may be chosen to be (generic ordered kk-hypergraph)-indiscernibles.

Proof.

Apply compactness and the modeling property (which ordered kk-hypergraphs have by Fact 2.6). ∎

Proposition 6.5.

For every 1<k<ω1<k<\omega, if φ(x;y)\varphi(x;y) is PMk+1PM_{k+1}, then ψ(x;y0yk)=i<k+1φ(x;yi)\psi(x;y_{0}...y_{k})=\bigwedge_{i<k+1}\varphi(x;y_{i}) is PMkPM_{k}.

Proof.

Assume φ(x;y)\varphi(x;y) has PMk+1PM_{k+1}. Consider the formula

ψ(x;y0yk)=i<k+1φ(x;yi)\psi(x;y_{0}...y_{k})=\bigwedge_{i<k+1}\varphi(x;y_{i})

We want to show that ψ\psi is PMkPM_{k}. Let n<ωn<\omega and let (n,E)(n,E) be a finite kk-hypergraph. Define a (k+1)(k+1)-hypergraph (V~,E~)(\tilde{V},\tilde{E}) as follows:

  • (i)

    for each i<ni<n, take k+1k+1 new vertices Ki=vi0,,vikK_{i}=v_{i}^{0},...,v_{i}^{k} with E~(vi0,,vik)\tilde{E}(v_{i}^{0},...,v_{i}^{k}).

  • (ii)

    for each hyperedge i0,,ik1i_{0},...,i_{k-1} in (n,E)(n,E), make Ki0Kik1K_{i_{0}}...K_{i_{k-1}} into a E~\tilde{E}-clique.

  • (iii)

    No other hyperedge is present.

Now, (V~,E~)(\tilde{V},\tilde{E}) is a (k+1)(k+1)-hypergraph, thus by PMk+1PM_{k+1}, there are (bv)vV~(b_{v})_{v\in\tilde{V}} realizing it. Let i0,,ik1{i_{0},...,i_{k-1}} a non-hyperedge in (n,E)(n,E). This implies that (Kij)j<k(K_{i_{j}})_{j<k} doesn’t form a E~\tilde{E}-clique and hence {φ(x;bv):vj<kKij}\{\varphi(x;b_{v})\ :\ v\in\bigcup_{j<k}K_{i_{j}}\} is inconsistent. Thus

{ψ(x;b¯Kij):j<k} is inconsistent.\{\psi(x;\bar{b}_{K_{i_{j}}})\ :\ j<k\}\text{ is inconsistent.}

If i0,,il1i_{0},...,i_{l-1} is a clique in (n,E)(n,E), then (Kij)j<l(K_{i_{j}})_{j<l} forms a clique in (V~,E~)(\tilde{V},\tilde{E}) and thus {φ(x;bv):vj<lKij}\{\varphi(x;b_{v})\ :\ v\in\bigcup_{j<l}K_{i_{j}}\} is consistent; this shows that

{φ(x;b¯Kj):j<l}\{\varphi(x;\bar{b}_{K_{j}})\ :\ j<l\}

is consistent. In conclusion ψ(x;y0yk)\psi(x;y_{0}...y_{k}) realizes (n,E)(n,E) and hence ψ\psi is PMkPM_{k}. ∎

Definition 6.6.

A theory is

  • (i)

    PMPM_{\infty} if there are formulae (φk(xk;yk))1<k<ω(\varphi_{k}(x_{k};y_{k}))_{1<k<\omega} such that for all 1<k<ω1<k<\omega φk(xk;yk)\varphi_{k}(x_{k};y_{k}) is PMkPM_{k}.

  • (ii)

    UPMUPM_{\infty} (Uniformly PMPM_{\infty}) if there is a formula φ(x;y)\varphi(x;y) with PMkPM_{k} for all 1<k<ω1<k<\omega.

Remark 6.7.

Clearly, at the level of theories, for all 2<k<ω2<k<\omega we have

PMUPMPMPMk+1PMkPM2PM\Rightarrow UPM_{\infty}\Rightarrow PM_{\infty}\Rightarrow...\Rightarrow PM_{k+1}\Rightarrow PM_{k}\Rightarrow...\Rightarrow PM_{2}

We will show that the implication PMk+1PMkPM_{k+1}\Rightarrow PM_{k} is strict for every 1<k<ω1<k<\omega.

Question 6.8.

Is there a PMPM_{\infty} but NUPMNUPM_{\infty} theory? Similarly, is there a UPMUPM_{\infty} but NPMNPM theory?

Definition 6.9.

Fix n<ωn<\omega. A relational language \mathcal{L} is nn-ary if, for every RR\in\mathcal{L}, ar(R)nar(R)\leq n. An \mathcal{L}-theory TT is nn-ary if

  • (i)

    \mathcal{L} is nn-ary, and

  • (ii)

    TT has QEQE.

Remark 6.10.

We could have defined a theory to be nn-ary if it is interdefinable with a theory satisfying the conditions in 6.9; on the other hand, all the properties that we are considering here are preserved under interdefinability and thus we don’t actually lose any generality.

In [3], Chernikov, Palacin and Takeuchi studied higher arity versions of IPIP, firstly introduced by Shelah, called IPkIP_{k} (0<k<ω0<k<\omega). In particular, they characterized NIPkNIP_{k} in terms of the collapse of ordered (k+1)(k+1)-hypergraphs indiscernibles to order-indiscernibles. Using this characterization, we get the following result.

Theorem 6.11.

For any 0<k<ω0<k<\omega, PMk+1PM_{k+1} implies IPkIP_{k}.

Proof.

Fix 0<k<ω0<k<\omega. By Theorem 5.4 in [3], it’s enough to exhibit a (generic ordered (k+1)(k+1)-hypergraph)-indiscernible that is not order indiscernible. Suppose φ(x;y)\varphi(x;y) is PMk+1PM_{k+1}. By Ramsey and compactness, we can find (bh)hH(b_{h})_{h\in H} (H,<,E)(H,<,E)-indiscernible witnessing it, where (H,<,E)(H,<,E) is the generic ordered (k+1)(k+1)-hypergraph. Let h0<h1<<hk+1Hh_{0}<h_{1}<...<h_{k+1}\in H such that

E(h0hk)¬E(h1hk+1)\models E(h_{0}...h_{k})\land\neg E(h_{1}...h_{k+1})

By PMk+1PM_{k+1} we have

xi<k+1φ(x;bhi)¬x0<i<k+2φ(x;bhi)\models\exists x\bigwedge_{i<k+1}\varphi(x;b_{h_{i}})\land\neg\exists x\bigwedge_{0<i<k+2}\varphi(x;b_{h_{i}})

and thus, even if the order type of h0<<hkh_{0}<...<h_{k} is the same as the one of h1<<hk+1h_{1}<...<h_{k+1}, we have

tp((bhi)i<k+1)tp((bhi)0<i<k+2tp((b_{h_{i}})_{i<k+1})\neq tp((b_{h_{i}})_{0<i<k+2}

In conclusion (bh)hH(b_{h})_{h\in H} is not order indiscernible and thus some formula has IPkIP_{k}. ∎

Corollary 6.12.

Any kk-ary theory is NPMk+1NPM_{k+1}.

Proof.

From [3], Example 2.2, we have that kk-ary theories are NIPkNIP_{k}. By 6.11, kk-ary theories are NPMk+1NPM_{k+1}. ∎

We now construct examples of theories with PMkPM_{k} but not PMk+1PM_{k+1} for all k2k\geq 2. This will show that the implication PMk+1PMkPM_{k+1}\Rightarrow PM_{k} is strict. We give a uniform recipe to build PMkPM_{k} NPMk+1NPM_{k+1} theories.

Example 6.13 (PMkPM_{k} NPMk+1NPM_{k+1} theory).

Fix 1<k<ω1<k<\omega and consider the language k={O,P,R,Ek}\mathcal{L}_{k}=\{O,P,R,E_{k}\} where P,OP,O are unary predicates, RR is a binary predicate and EkE_{k} is a kk-ary predicate. Consider the k\mathcal{L}_{k}-theory T0T_{0} axiomatized as follows:

  • (A1)(A_{1})

    P,OP,O partition the universe.

  • (A2)(A_{2})

    EkPkE_{k}\subseteq P^{k} and (P,Ek)(P,E_{k}) is a kk-hypergraph.

  • (A3)(A_{3})

    RO×PR\subseteq O\times P

  • (A4)(A_{4})

    v0vk1x(i<kR(x,vi)¬Ek(v¯))\forall v_{0}...v_{k-1}\forall x(\bigwedge_{i<k}R(x,v_{i})\to\neg E_{k}(\bar{v}))

Let 𝒦={AT0:|A|<0}\mathcal{K}=\{A\models T_{0}\ :\ |A|<\aleph_{0}\}. Similarly to the proof of Proposition 5.8, it can be shown that 𝒦\mathcal{K} is a Fraïssé class closed under free-amalgamation. Let k\mathcal{M}_{k} be the Fraïssé limit. Intuitively k\mathcal{M}_{k} has two disjoint sets (OO,PP), (P(k),Ek)(P(\mathcal{M}_{k}),E_{k}) is the generic kk-hypergraph and O(k)O(\mathcal{M}_{k}) is the set of witnesses for the consistency along RR of independent sets and the inconsistency along RR of kk-hyperedges. In particular Tk:=Th(k)T_{k}:=Th(\mathcal{M}_{k}) is PMkPM_{k} (witnessed by ¬R\neg R), ω\omega-categorical, NSOP4NSOP_{4} (See [6], Theorem 1.1) and eliminates quantifiers. Moreover, by Corollary 6.12, TkT_{k} is NPMk+1NPM_{k+1}.

We now consider a well-known theory and see how it relates with this hierarchy:

Example 6.14 (Generic triangle-free graph).

Let ={e}\mathcal{L}=\{e\} where ee is a binary relation and let 𝒦\mathcal{K} be the class of finite triangle-free graphs. It can be shown that 𝒦\mathcal{K} is a Fraïssé class; let \mathcal{H} be its limit and let T=Th()T_{\mathcal{H}}=Th(\mathcal{H}).

Proposition 6.15.

TT_{\mathcal{H}} is PM2PM_{2}.

Proof.

We claim that the formula φ(x;yz)=e(x,y)e(x,z)\varphi(x;yz)=e(x,y)\land e(x,z) is PM2PM_{2}. Let n<ωn<\omega and let (V,E)(V,E) any finite graph an nn vertices. Construct the graph (V~,E~)(\tilde{V},\tilde{E}) as follows: Replace each vertex vVv\in V with two vertices bvcvb_{v}c_{v} with ¬bvE~cv\neg b_{v}\tilde{E}c_{v}. Whenever {v,w}E\{v,w\}\notin E, make a new edge between {bv,cw}E~\{b_{v},c_{w}\}\in\tilde{E} and no other edges.

Claim (V~,E~)(\tilde{V},\tilde{E}) is triangle-free.

Proof of the Claim: As a notation, given a graph and a vertex vv, we write N(v)N(v) for the graph neighborhood of vv i.e. the set of all the vertices in the graph that are connected by an edge to vv. Suppose, toward contradiction, that there is a triangle aE~aE~a′′a~a\tilde{E}a^{\prime}\tilde{E}a^{\prime\prime}\tilde{a}. Since there’s no edge between bvb_{v} and cvc_{v} for every vv, there must be v0,v1,v2v_{0},v_{1},v_{2}, such that the triangle is among (bvicvi)i<3(b_{v_{i}}c_{v_{i}})_{i<3}. Suppose that bv0b_{v_{0}} is in the triangle. But N(bv0)={cv1,cv2}N(b_{v_{0}})=\{c_{v_{1}},c_{v_{2}}\} and ¬cv1E~cv2\neg c_{v_{1}}\tilde{E}c_{v_{2}}, contradiction. Then it must be the case that cv0c_{v_{0}} is in the triangle. But N(cv0)={bv1,bv2}N(c_{v_{0}})=\{b_{v_{1}},b_{v_{2}}\} and ¬bv1E~bv2\neg b_{v_{1}}\tilde{E}b_{v_{2}}. Contradiction. Thus (V~,E~)(\tilde{V},\tilde{E}) is triangle-free. \blacksquare

Since (V~,E~)(\tilde{V},\tilde{E}) is a finite triangle-free graph, it can be embedded in \mathcal{H}. Now, let {v,w}E\{v,w\}\notin E. Then, if there is a vertex aa such that

aφ(x;bvcv)φ(x;bwcw)=(e(x,bv)e(x,cv))((e(x,bw)e(x,cw))a\models\varphi(x;b_{v}c_{v})\land\varphi(x;b_{w}c_{w})=(e(x,b_{v})\land e(x,c_{v}))\land((e(x,b_{w})\land e(x,c_{w}))

Then a,bv,cwa,b_{v},c_{w} form a triangle; contradiction. On the other hand, if v0,,vn1v_{0},...,v_{n-1} is a clique, then (bvicvi)i<n(b_{v_{i}}c_{v_{i}})_{i<n} is an independent set of vertices and hence, by the extension property of \mathcal{H}, there is a vertex aa such that ai<ne(x,bvi)e(x,cvi)a\models\bigwedge_{i<n}e(x,b_{v_{i}})\land e(x,c_{v_{i}}) and thus

{φ(x;bvicvi):i<n} is consistent.\{\varphi(x;b_{v_{i}}c_{v_{i}})\ :\ i<n\}\text{ is consistent.}

In conclusion φ(x;y)\varphi(x;y) is PM2PM_{2}. Thus, by definition, TT_{\mathcal{H}} is PM2PM_{2}. ∎

Moreover, since TT_{\mathcal{H}} is a binary theory, by Corollary 6.12, TT_{\mathcal{H}} is NPM3NPM_{3}.

Acknowledgements

This work is part of the author’s PhD dissertation in progress. The author would like to sincerely thank Alex Kruckman for his guidance, valuable help and fruitful discussions. Furthermore, the exciting environment of the Wesleyan University logic group played an important role in this project, in particular the encouragement and the helpful conversations the author had with Cameron Hill, Rehana Patel and Alex Van Abel.

References

  • Casanovas [2011] Enrique Casanovas. NIP formulas and theories. Model theory Seminar notes, 2011.
  • Chatzidakis and Hrushovski [1999] Zoé Chatzidakis and Ehud Hrushovski. Model theory of difference fields. Trans. Amer. Math. Soc., 351(8):2997–3071, 1999. ISSN 0002-9947,1088-6850. doi: 10.1090/S0002-9947-99-02498-8. URL https://doi.org/10.1090/S0002-9947-99-02498-8.
  • Chernikov et al. [2019] Artem Chernikov, Daniel Palacin, and Kota Takeuchi. On nn-dependence. Notre Dame J. Form. Log., 60(2):195–214, 2019. ISSN 0029-4527,1939-0726. doi: 10.1215/00294527-2019-0002. URL https://doi.org/10.1215/00294527-2019-0002.
  • [4] Gabriel Conant. Map of the universe. URL https://www.forkinganddividing.com/.
  • Conant [2012] Gabriel Conant. Dividing lines in unstable theories. 2012.
  • Conant [2017] Gabriel Conant. An axiomatic approach to free amalgamation. J. Symb. Log., 82(2):648–671, 2017. ISSN 0022-4812,1943-5886. doi: 10.1017/jsl.2016.42. URL https://doi.org/10.1017/jsl.2016.42.
  • Cooper [1982] Glen R. Cooper. On complexity of complete first-order theories. Z. Math. Logik Grundlagen Math., 28(2):93–136, 1982. ISSN 0044-3050. doi: 10.1002/malq.19820280802. URL https://doi.org/10.1002/malq.19820280802.
  • Hrushovski and Pillay [1994] Ehud Hrushovski and Anand Pillay. Groups definable in local fields and pseudo-finite fields. Israel J. Math., 85(1-3):203–262, 1994. ISSN 0021-2172,1565-8511. doi: 10.1007/BF02758643. URL https://doi.org/10.1007/BF02758643.
  • Kaplan et al. [2024] Itay Kaplan, Nicholas Ramsey, and Pierre Simon. Generic stability independence and treeless theories. Forum Math. Sigma, 12:Paper No. e49, 27, 2024. ISSN 2050-5094. doi: 10.1017/fms.2024.35. URL https://doi.org/10.1017/fms.2024.35.
  • Kim [1998] Byunghan Kim. Forking in simple unstable theories. J. London Math. Soc. (2), 57(2):257–267, 1998. ISSN 0024-6107,1469-7750. doi: 10.1112/S0024610798005985. URL https://doi.org/10.1112/S0024610798005985.
  • Kim and Kim [2011] Byunghan Kim and Hyeung-Joon Kim. Notions around tree property 1. Ann. Pure Appl. Logic, 162(9):698–709, 2011. ISSN 0168-0072,1873-2461. doi: 10.1016/j.apal.2011.02.001. URL https://doi.org/10.1016/j.apal.2011.02.001.
  • Kim and Pillay [1997] Byunghan Kim and Anand Pillay. Simple theories. volume 88, pages 149–164. 1997. doi: 10.1016/S0168-0072(97)00019-5. URL https://doi.org/10.1016/S0168-0072(97)00019-5. Joint AILA-KGS Model Theory Meeting (Florence, 1995).
  • Kim and Pillay [1998] Byunghan Kim and Anand Pillay. From stability to simplicity. Bull. Symbolic Logic, 4(1):17–36, 1998. ISSN 1079-8986,1943-5894. doi: 10.2307/421004. URL https://doi.org/10.2307/421004.
  • Kruckman and Ramsey [2024] Alex Kruckman and Nicholas Ramsey. A New Kim’s Lemma. Model Theory, 3(3):825–860, 2024. ISSN 2832-9058,2832-904X. doi: 10.2140/mt.2024.3.825. URL https://doi.org/10.2140/mt.2024.3.825.
  • Nešetřil and Rödl [1983] Jaroslav Nešetřil and Vojtěch Rödl. Ramsey classes of set systems. J. Combin. Theory Ser. A, 34(2):183–201, 1983. ISSN 0097-3165,1096-0899. doi: 10.1016/0097-3165(83)90055-9. URL https://doi.org/10.1016/0097-3165(83)90055-9.
  • Scow [2015] Lynn Scow. Indiscernibles, EM-types, and Ramsey classes of trees. Notre Dame J. Form. Log., 56(3):429–447, 2015. ISSN 0029-4527,1939-0726. doi: 10.1215/00294527-3132797. URL https://doi.org/10.1215/00294527-3132797.
  • Scow [2010] Lynn Cho Scow. Characterization theorems by generalized indiscernibles. ProQuest LLC, Ann Arbor, MI, 2010. ISBN 978-1124-52093-3. URL http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqdiss&rft_dat=xri:pqdiss:3444383. Thesis (Ph.D.)–University of California, Berkeley.
  • Shelah [1990] S. Shelah. Classification theory and the number of nonisomorphic models, volume 92 of Studies in Logic and the Foundations of Mathematics. North-Holland Publishing Co., Amsterdam, second edition, 1990. ISBN 0-444-70260-1.
  • Shelah [1971] Saharon Shelah. The number of non-isomorphic models of an unstable first-order theory. Israel J. Math., 9:473–487, 1971. ISSN 0021-2172. doi: 10.1007/BF02771463. URL https://doi.org/10.1007/BF02771463.
  • Shelah [1980] Saharon Shelah. Simple unstable theories. Ann. Math. Logic, 19(3):177–203, 1980. ISSN 0003-4843. doi: 10.1016/0003-4843(80)90009-1. URL https://doi.org/10.1016/0003-4843(80)90009-1.
  • Shelah [1996] Saharon Shelah. Toward classifying unstable theories. Ann. Pure Appl. Logic, 80(3):229–255, 1996. ISSN 0168-0072,1873-2461. doi: 10.1016/0168-0072(95)00066-6. URL https://doi.org/10.1016/0168-0072(95)00066-6.
  • Shelah [2000] Saharon Shelah. On what I do not understand (and have something to say). I. volume 166, pages 1–82. 2000. doi: 10.4064/fm-166-1-2-1-82. URL https://doi.org/10.4064/fm-166-1-2-1-82. Saharon Shelah’s anniversary issue.
  • Shelah [2012] Saharon Shelah. Model theory. 2012. URL https://arxiv.org/abs/1208.1301.