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

\UseRawInputEncoding

Stable formulas in ordered structures

Daniel Max Hoffmann Instytut Matematyki
Uniwersytet Warszawski
Warszawa
Poland
and
  Department of Mathematics
University of Notre Dame
Notre Dame
IN
USA
[email protected] https://sites.google.com/site/danielmaxhoffmann/home
Chieu-Minh Tran Department of Mathematics
University of Notre Dame
Notre Dame
IN
USA
[email protected] https://faculty.math.illinois.edu/ mctran2/
 and  Jinhe Ye Institut de Mathématiques de Jussieu-Paris Rive Gauche [email protected] https://sites.google.com/view/vincentye
Abstract.

We classify the stable formulas in the theory of Dense Linear Orders without endpoints, the stable formulas in the theory of Divisible Abelian Groups, and the stable formulas without parameters in the theory of Real Closed Fields. The third result, unexpectedly, requires the Hironaka’s theorem on resolution of singularities.

2010 Mathematics Subject Classification. Primary 03C64; Secondary 03C45, 03C10
Key words and phrases. real closed fields, ordered structures, stable formulas
SDG. The first author is supported by the Polish Natonal Agency for Academic Exchange and the National Science Centre (Narodowe Centrum Nauki, Poland) grants no. 2016/21/N/ST1/01465, and 2015/19/B/ST1/01150.

1. Introduction

In recent years, we have seen rapid development of the neostability program which aims to extend the ideas of stability to other settings. Efforts have been made toward investigating weaker notions (NIP, simplicity, NSOP1, NTP2, etc), considering the stable components (stably dominated types, stables formulas, etc) in unstable theories, or a mix and match between these themes; see [16] for the relevant definitions (e.g. Section 8.2 for the stability related notions). In this paper, we are interested in stable formulas—also called “stable relations”—in unstable theories, in other words, the local stability of these theories. This is an old direction which nevertheless continues to hold relevance with recent applications in combinatorics ([5], [4], [3], [6], [14]). Stable formulas is related to thorn-forking [7] and is the subject of stable forking conjecture for simple theories ([12], [13]). Surprisingly, not much attention have been paid to the down-to-earth problem of classifying stable formulas in frequently seen examples of unstable theory. Our goal here is to fill this gap for the most obvious unstable structures, those that involves an ordering.

We know that ordering gives us unstability. The example below provides us with a slightly more general situation where we have unstability, namely, the formula defines a “large set” with a “slope”. It also points out why “large” and having a “slope” is necessary.

Example 1.1.

Consider a strictly increasing function f:[0,1][0,1]f:[0,1]_{\operatorname{\mathbb{R}}}\to[0,1]_{\operatorname{\mathbb{R}}} definable in the ordered field (;+,×)(\operatorname{\mathbb{R}};+,\times) such that f(0)=0f(0)=0 and f(1)=1f(1)=1, and take

D:={(a,b)[0,1]2|b<f(a)}.D:=\{(a,b)\in[0,1]_{\operatorname{\mathbb{R}}}^{2}\;|\;b<f(a)\}.

We will construct a sequence (ai,bi)i<ω(a_{i},b_{i})_{i<\omega} such that (ai,bj)D(a_{i},b_{j})\in D if and only if iji\leqslant j. We start with any (a1,b1)D(a_{1},b_{1})\in D. Take 0<a2<a10<a_{2}<a_{1} such that f(a2)<b1f(a_{2})<b_{1}, and 0<b2<f(a2)0<b_{2}<f(a_{2}) and continue this way. Such a sequence can be similarly produced if we replace ff by a decreasing function. On the other hand, it is easy to see that there is no such sequence (ai,bj)(a_{i},b_{j}) such that (ai,bj)D(a_{i},b_{j})\in D if and only if iji\leqslant j when DD is the graph of ff or the entire set [0,1]2[0,1]_{\operatorname{\mathbb{R}}}^{2}.

For the rest of the paper, let TT be either the theory of dense linear orderings in L={<}L=\{<\} (DLO), the theory of divisible ordered abelian groups in L={+,0,<}L=\{+,0,<\} (DOAG), or the theory of real closed fields in L={+,,,0,1,<}L=\{+,-,\cdot,0,1,<\} (RCF), let \mathscr{M} be a model of TT with underlying set MM, let φ(x;y)\varphi(x;y) be an L(M)L(M)-formula, and let dim\dim denotes the o-minimal dimension of TT; see [17] for the basic definition and results. When we say that φ(x;y)\varphi(x;y) is stable, we implicitly assume that stability is with respect to the pair (x;y)(x;y). We say that φ(x;y)\varphi(x;y) is rectangular if φ(x;y)\varphi(x;y) is TT-equivalent to ψ(x)θ(y)\psi(x)\wedge\theta(y) with ψ(x)\psi(x) and θ(y)\theta(y) being L(M)L(M)-formulas. It is easy to see that rectangular formulas are stable, and so are their boolean combinations. Propostion 1.2, which combines the later Propostion 3.6 and Proposition 3.8, tell us that Example 1.1 essentially points us in the right direction. Note that the condition dimφ()=|x|+|y|\dim\varphi(\mathscr{M})=|x|+|y| is the precise version of what we meant by “large”, and the notion of rectangular formula makes precise the idea of “having no slope”.

Proposition 1.2.

Suppose φ(x;y)\varphi(x;y) is stable, and dimφ()=|x|+|y|\dim\varphi(\mathscr{M})=|x|+|y| . Then there is an L(M)L(M)-formula φ(x;y)\varphi^{\prime}(x;y) which is a disjuntion of rectangular L(M)L(M)-formulas such that dim(φ()φ())<|x|+|y|\dim\big{(}\varphi(\mathscr{M})\triangle\varphi^{\prime}(\mathscr{M})\big{)}<|x|+|y|.

Proposition 1.2 is not sufficient for the purpose of classifying stable formulas in TT as it says nothing when dimφ()<|x|+|y|\dim\varphi(\mathscr{M})<|x|+|y|. We say that φ(x;y)\varphi(x;y) is order-free if φ(x;y)\varphi(x;y) is equivalent over TT to quantifier-free L(M)L(M)-formulas which do not contain <<. Order-free formulas form another natural class of stables formulas. It is possible to have φ(x;y)\varphi(x;y) order-free with dimφ()<|x|+|y|\dim\varphi(\mathscr{M})<|x|+|y|. If we consider a conjunction of a rectangular formula and an order-free formula, we still obtain a stable formula. Formulas of this form are said to be special stable. A natural guess would be that every stable L(M)L(M)-formula is equivalent over TT to a finite union of special stable formulas. In Theorem 4.1, we show this is the case when the theory under consideration is either the theory of Dense Linear Orders without Endpoints or the theory of Divisible Ordered Abelian Groups:

Theorem 1.3.

Suppose TT is either DLO\mathrm{DLO} or DOAG\mathrm{DOAG} and φ(x;y)\varphi(x;y) is stable. Then φ(x;y)\varphi(x;y) is equivalent over TT to a disjunction of special stable L(M)L(M)-formulas.

We expect that a result in this line can be obtained for more general linear orderings in L={<}L=\{<\} and linearly ordered abelian groups in L={<,+,0}L=\{<,+,0\}. However, we do not address this question in this paper.

Now, let us move to theory of Real Closed Fields and start with the following example, where we can see that the above description breaks down.

Example 1.4.

In the next figure, the part of the curve on the left above the dashed line is defined by the system of equations and inequalities on the right:

[Uncaptioned image]
{IEEEeqnarray*}

rCl 0 &= x^4+2x^2y^2+y^4+x^3-xy^2;

0 ⩽ x;

0 ⩽ y;

3

x ⩽ y.

Let φ(x;y)\varphi(x;y) with |x|=|y|=1|x|=|y|=1 be the conjunction of the above equations and inequalities. It is easy to check that φ(x;y)\varphi(x;y) is stable; in fact, every formula in two variables defining a one dimensional set is stable by cell decomposition. Note that the equation x4+2x2y2+y4+x3xy2x^{4}+2x^{2}y^{2}+y^{4}+x^{3}-xy^{2} defines an irreducible algebraic set of dimension 11. Hence, if φ(x;y)\varphi(x;y) is a disjunction of special stable formulas, we can further arrange that each of these special stable formulas is a conjunction of x4+2x2y2+y4+x3xy2x^{4}+2x^{2}y^{2}+y^{4}+x^{3}-xy^{2} with a rectangular L(M)L(M)-formula. As RCF\mathrm{RCF} is o-minimal, we can arrange that each of these rectangular formulas defines a set of the form I×JI\times J where I,JI,J\subseteq\operatorname{\mathbb{R}} are intervals. However, any set of the aforementioned form I×JI\times J containing the point (0,0)(0,0) will have to include a part of the curve below the dashed line y=3xy=\sqrt{3}x. Thus, φ(x;y)\varphi(x;y) is not a disjunction of special stable L(M)L(M)-formulas.

In Example 1.4, the obstacle in expressing φ(x;y)\varphi(x;y) as a finite disjunction of special stable formulas comes from the singularity at (0,0)(0,0) of the curve in the picture. This brings us to the idea of using blowing up, or more precisely, resolution of singularities. It turns out that this is essentially the only obstruction when the stable formulas considered are parameter-free, and every such formula is equivalent over RCF\operatorname{RCF} to disjunctions of special stable formulas up to a certain kind of isomorphism. To be more precise, in Section 2.3, we will define the notion of order-free isomorphism between a relation defined by a formula φ(x;y)\varphi(x;y) and a relation defined by a formula φ(x;y)\varphi^{\prime}(x^{\prime};y^{\prime}). This notion generalizes birational equivalence with a catch, namely, the division of variables must be respected. We obtain in Section 5 our main result in this paper:

Theorem 1.5.

Suppose TT is RCF\operatorname{RCF}. Then an LL-formula φ(x;y)\varphi(x;y) is stable if and only if it is equivalent over TT to a disjunction of formulas order-free isomorphic to a special stable LL-formula.

One could ask whether the statement of the above theorem also holds for a stable formula with parameters, and we think the answer should be yes. Our current proof for Theorem 1.5, in fact, goes through for the more general case when φ(x;y)\varphi(x;y) is a formula with parameters over an Archimedean subfield of \mathscr{M}. However, the proof involves topological compactness. It is unclear if this technique transfers to the most general case when the infinite/non-Archimedean parameters occur in φ(x;y)\varphi(x;y). In a parallel direction, some natural subsequent questions could be the classification of stable formulas in exp\operatorname{\mathbb{R}}_{\mathrm{exp}} and an\operatorname{\mathbb{R}}_{\mathrm{an}}, multi-ordered fields ([11]) or p\mathbb{Q}_{p}.

Notations and conventions

Throughout mm and nn are in ={0,1,}\operatorname{\mathbb{N}}=\{0,1,\ldots\}. We adopt the usual monster model gadgets: 𝓜\boldsymbol{\mathscr{M}} is a κ\kappa-saturated and strongly κ\kappa-homogeneous with underlying set 𝑴\boldsymbol{M}, we refer to definable with parameters in 𝑴\boldsymbol{M} as simply definable, parameter sets AA and BB are assumed to be small with respect to 𝓜\boldsymbol{\mathscr{M}}, and other models are assumed to be elementary submodels of 𝓜\boldsymbol{\mathscr{M}}. Moreover, we identify a formula φ\varphi with the set DD it defines in 𝓜\boldsymbol{\mathscr{M}}, and work semantically with definable sets instead of formulas.

In the paper, we use the following convention about variables: Throughout xx and yy are finite tuples of variables. Let 𝑴x\boldsymbol{M}^{x} denote the cartesian power of 𝑴\boldsymbol{M} indexed by the variables in xx. By (x,y)(x,y) we denote the tuple xx extended by the tuple yy. With an eye on stability, we also use the usual notational convention (x;y)(x;y) to emphasize the division of variables. So when we write D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)}, we mean D𝑴(x,y)D\subseteq\boldsymbol{M}^{(x,y)} and we have a fixed division of variables (x;y)(x;y).

2. Preliminaries

We continue to let TT be DLO=Th(,<)\text{DLO}=\operatorname{Th}(\operatorname{\mathbb{R}},<), DOAG=Th(,+,0,<)\text{DOAG}=\operatorname{Th}(\operatorname{\mathbb{R}},+,0,<), or RCF=Th(,+,,,0,1,<)\mathrm{RCF}=\text{Th}(\operatorname{\mathbb{R}},+,-,\cdot,0,1,<). Let 𝓜\boldsymbol{\mathscr{M}} be a model of TT with underlying set 𝑴\boldsymbol{M}, let D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)} be L(𝑴)L(\boldsymbol{M})-definable, and let dim\dim denote the o-minimal dimension of TT. Recall that D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)} is unstable (with respect to the variable division (x;y)(x;y)) if some (every) formula defining DD is unstable. More explicitly, such DD is unstable if there is (ai,bj)i<ω,j<ω(a_{i},b_{j})_{i<\omega,j<\omega} such that (ai,bj)D(a_{i},b_{j})\in D if and only if iji\leqslant j. We call such (ai,bj)i<ω,j<ω(a_{i},b_{j})_{i<\omega,j<\omega} an unstable witness of DD. As TT is complete, classifying stable formulas over TT is the same as classifying stable sets over 𝓜\boldsymbol{\mathscr{M}}.

Toward classifying stable sets, we consider several classes of sets which are obviously stable. We say that DD is order-free definable if it is defined by an order-free formula as in the introduction. This property is equivalent to being quantifier-free definable in the reduct of 𝓜\boldsymbol{\mathscr{M}} without the linear order. A typical example of an order-free definable set is an algebraic set in 𝓜RCF\boldsymbol{\mathscr{M}}\models\mathrm{RCF}. We say a function f:XYf:X\to Y is order-free definable if the graph of ff is an order-free definable. Abusing notation, the restriction of ff to a subset ZZ of XX is also called order-free definable.

A definable set is rectangular if it is defined by a rectangular formula as in the introduction. It is easy to see that DD is rectangular if and only if D=X×YD=X\times Y with X𝑴xX\subseteq\boldsymbol{M}^{x} and Y𝑴yY\subseteq\boldsymbol{M}^{y} definable in 𝓜\boldsymbol{\mathscr{M}}. Clearly, order-free definable sets and rectangular sets are stable. A special kind of stable sets which generalizes both order-free definable sets and rectangular set is given in the following definition.

Definition 2.1.

We say that D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)} is a special stable set if DD is defined by a special stable formula (with respect to the division of variables (x;y)(x;y)), equivalently, there exist definable sets X𝑴xX\subseteq\boldsymbol{M}^{x} and Y𝑴yY\subseteq\boldsymbol{M}^{y}, and an order-free definable set Z𝑴(x;y)Z\subseteq\boldsymbol{M}^{(x;y)} such that D=Z(X×Y)D=Z\cap(X\times Y).

The following easy fact will be later used in the proof of Theorem 5.2.

Lemma 2.2.

Every Boolean combination of special stable sets is a finite union of special stable sets.

Proof.

The intersection of special stable sets is again a special stable set, so it suffices to show that the complement of a special stable set is a finite union of special stable sets. Consider a special set D=Z(X×Y)D=Z\cap(X\times Y) where X𝑴xX\subseteq\boldsymbol{M}^{x} and Y𝑴yY\subseteq\boldsymbol{M}^{y} are definable, and Z𝑴(x;y)Z\subseteq\boldsymbol{M}^{(x;y)} is order-free definable. Note that

Dc=Zc(X×Y)c=(ZcX×Y)(X×Y)c,D^{c}=Z^{c}\cup(X\times Y)^{c}=(Z^{c}\cap X\times Y)\cup(X\times Y)^{c},

(X×Y)c=(Xc×Yc)(Xc×Y)(X×Yc)(X\times Y)^{c}=(X^{c}\times Y^{c})\cup(X^{c}\times Y)\cup(X\times Y^{c}), and ZcZ^{c} is order-free definable. The desired conclusion follows. ∎

It is well-known that the stability of formulas is preserved under taking Boolean combinations [15, Lemma 2.1], so Boolean combinations of special stable sets are stable. As it was noted in Example 1.4 from the introduction, not all stable sets are Boolean combinations of special stable sets. We introduce the following notion to remedy this situation.

Definition 2.3.

Suppose that D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)}, D𝑴(x;y)D^{\prime}\subseteq\boldsymbol{M}^{(x^{\prime};y^{\prime})}, πx:𝑴(x,y)𝑴x\pi_{x}:\boldsymbol{M}^{(x,y)}\to\boldsymbol{M}^{x} and πy:𝑴(x,y)𝑴y\pi_{y}:\boldsymbol{M}^{(x,y)}\to\boldsymbol{M}^{y} are the projection maps, and πx\pi_{x^{\prime}} and πy\pi_{y^{\prime}} are defined similarly. We say that a map f:DDf:D^{\prime}\to D is an order-free morphism if there exist

  1. (1)

    a finite covering (Ui)iI(U_{i})_{i\in I} of DD by order-free definable sets,

  2. (2)

    a finite covering (Ui)iI(U^{\prime}_{i})_{i\in I} of DD^{\prime} by order-free definable sets,

  3. (3)

    a family of order-free definable maps (fxi:πx(Ui)πx(Ui))iI\big{(}f^{i}_{x}:\pi_{x^{\prime}}(U^{\prime}_{i})\to\pi_{x}(U_{i})\big{)}_{i\in I},

  4. (4)

    a family of order-free definable maps (fyi:πy(Ui)πy(Ui))iI\big{(}f^{i}_{y}:\pi_{y^{\prime}}(U^{\prime}_{i})\to\pi_{y}(U_{i})\big{)}_{i\in I}

such that for each iIi\in I we have that fxi×fyi(Ui)=Uif^{i}_{x}\times f^{i}_{y}(U^{\prime}_{i})=U_{i}, fxi×fyi(UiD)=UiDf^{i}_{x}\times f^{i}_{y}(U^{\prime}_{i}\cap D^{\prime})=U_{i}\cap D and that f|UiD=(fxi×fyi)|UiDf|_{U^{\prime}_{i}\cap D^{\prime}}=(f^{i}_{x}\times f^{i}_{y})|_{U^{\prime}_{i}\cap D^{\prime}}. If, moreover, the map fxi×fyi:UiUif^{i}_{x}\times f^{i}_{y}:U^{\prime}_{i}\to U_{i} is a bijection for each iIi\in I, then we call ff an order-free isomorphism. If there exists an order-free isomorphism between DD and DD^{\prime}, we say that DD is order-free isomorphic to DD^{\prime}.

Because of the preservation of stability under Boolean combinations, one can deduce the following easy fact: if the definable sets D1,,Dn𝑴(x;y)D_{1},\ldots,D_{n}\subseteq\boldsymbol{M}^{(x;y)} are stable and DD1DnD\subseteq D_{1}\cup\ldots\cup D_{n} is definable, then DD is stable if and only if DDiD\cap D_{i} is stable for each ini\leqslant n. This easy fact leads to the following, more important, observation:

Lemma 2.4.

Suppose the definable sets D𝐌(x;y)D\subseteq\boldsymbol{M}^{(x;y)} and D𝐌(x;y)D^{\prime}\subseteq\boldsymbol{M}^{(x^{\prime};y^{\prime})} are order-free isomorphic. If DD is stable, then DD^{\prime} is stable.

Proof.

Let UiU^{\prime}_{i}, UiU_{i}, fxif^{i}_{x} and fyif^{i}_{y}, for iIi\in I, be as in Definition 2.3. Suppose that DD^{\prime} is unstable. From the fact preceding the lemma, we see that DUkD^{\prime}\cap U^{\prime}_{k} is unstable for some kIk\in I. Let (ai,bj)i,j<ω(a^{\prime}_{i},b^{\prime}_{j})_{i,j<\omega} be an unstable witness of DUkD^{\prime}\cap U^{\prime}_{k}. By Ramsey’s theorem [16, Theorem 5.1.5] there exists an infinite set NωN\subseteq\omega such that for all i,jNi,j\in N we have (ai,bj)Uk(a^{\prime}_{i},b^{\prime}_{j})\in U^{\prime}_{k} by the fact that UkU_{k} is order-free. Without loss of generality, we assume that N=ωN=\omega.

Consider ai:=fxk(ai)a_{i}:=f^{k}_{x}(a^{\prime}_{i}), bj:=fyk(bj)b_{j}:=f^{k}_{y}(b^{\prime}_{j}), where i,j<ωi,j<\omega. Because (ai,bj)Uk(a^{\prime}_{i},b^{\prime}_{j})\in U^{\prime}_{k} for all i,j<ωi,j<\omega, we have that

(ai,bj)=fxk×fyk(ai,bj)Uk for all i,j<ω.(a_{i},b_{j})=f^{k}_{x}\times f^{k}_{y}(a^{\prime}_{i},b^{\prime}_{j})\in U_{k}\text{ for all }i,j<\omega.

If iji\leqslant j then (ai,bj)DUk(a^{\prime}_{i},b^{\prime}_{j})\in D^{\prime}\cap U^{\prime}_{k} and thus (ai,bj)DUk(a_{i},b_{j})\in D\cap U_{k}. If i>ji>j, then (ai,bj)DUk(a^{\prime}_{i},b^{\prime}_{j})\not\in D^{\prime}\cap U^{\prime}_{k}. As fxk×fykf^{k}_{x}\times f^{k}_{y} is injective over UkU^{\prime}_{k}, we have that (ai,bj)DUk(a_{i},b_{j})\not\in D\cap U_{k}. Therefore, (ai,bj)i,j<ω(a_{i},b_{j})_{i,j<\omega} is an unstable witness of DD. ∎

3. Large stable sets

Our strategy of classifying stable sets DD relies on an induction on dim(D)\dim(D). The goal of this section is to obtain a description of stable sets D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)} with dim(D)=|x|+|y|\dim(D)=|x|+|y|. See Proposition 3.6 and 3.8 for such a description. Eventually, they will enable us to carry out the induction.

Next, we will obtain a description for large stable sets up to sets of smaller dimension. A definable set EE is definably connected by codimension 22 if for all definable EEE^{\prime}\subseteq E with dim(EE)dimE2\dim(E\setminus E^{\prime})\leq\dim E-2, the set EE^{\prime} is definable connected in the o-minimal structure 𝓜\boldsymbol{\mathscr{M}}. Before moving forward, let us recall a few basic facts from o-minimal structures, see [17] for details on this subject.

Suppose XX is a definable subset of 𝑴x\boldsymbol{M}^{x}. For a𝑴xa\in\boldsymbol{M}^{x}, the local dimension at aa of XX is defined as

dima(X)=inf{dim(UX)U is a definable open subset of 𝑴x},\dim_{a}(X)=\inf\{\dim(U\cap X)\mid U\text{ is a definable open subset of }\boldsymbol{M}^{x}\},

where the “dim(UX)\dim(U\cap X)” refers to the o-minimal dimension and “open” refers to being open in the topology generated by open boxes defined by the ordering (here and in the rest of this paper an open box is just the Cartesian product of open intervals - similarly for a closed box). Suppose EE is an order-free definable subset of 𝑴x\boldsymbol{M}^{x} with nonempty XEX\subseteq E. As usual, the closure clE(X)\mathrm{cl}_{E}(X) of XX is the smallest closed subsets in EE containing XX, the interior intE(X)\mathrm{int}_{E}(X) of XX in EE is the largest open subset of EE that XX contains, and the boundary bdE(X)\mathrm{bd}_{E}(X)is defined as clE(X)intE(X)\mathrm{cl}_{E}(X)\setminus\mathrm{int}_{E}(X). The essence of XX in EE is the set

essE(X)={aEdima(X)=dimX}.\operatorname{ess}_{E}(X)=\{a\in E\mid\dim_{a}(X)=\dim X\}.

The essential interior of XX in EE is the set

intessE(X)={aessE(X)dima(EX)<dim(X)}.\operatorname{\operatorname{int}^{\operatorname{ess}}}_{E}(X)=\{a\in\operatorname{ess}_{E}(X)\mid\dim_{a}(E\setminus X)<\dim(X)\}.

The essential boundary of XX in EE is the set

bdessE(X)={aessE(X)dima(EX)dimX}.\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X)=\{a\in\ \operatorname{ess}_{E}(X)\mid\dim_{a}(E\setminus X)\geq\dim X\}.

Note that ess(X)=intessE(X)bdessE(X)\operatorname{ess}(X)=\operatorname{\operatorname{int}^{\operatorname{ess}}}_{E}(X)\,\cup\,\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X). We will omit “EE” if E=𝑴nE=\boldsymbol{M}^{n} for some natural number nn.

The following example illustrates the above notions. This was pointed out to us by the anonymous referee leading to a correction of a mistake in the earlier proof of Lemma 3.5, where a wrong definition of essence was used.

Example 3.1.

Let XX be the open unit ball in 3\operatorname{\mathbb{R}}^{3} with the equator, i.e.,

X={(x,y,z):x2+y2+z2<1}{(x,y,z):z=0x2+y2=1}X=\{(x,y,z):x^{2}+y^{2}+z^{2}<1\}\cup\{(x,y,z):z=0~{}x^{2}+y^{2}=1\}

and E=3E=\operatorname{\mathbb{R}}^{3}. In this case, essE(X)\operatorname{ess}_{E}(X) is the unit closed ball. And bdessE(X)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X) is the unit sphere and intessE(X)\operatorname{\operatorname{int}^{\operatorname{ess}}}_{E}(X) is the open unit ball. In particular, bdessE(X)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X) has dimension 2=312=3-1, which is in accordance with Lemma 3.5.

We collect some well known facts about o-minimal theories.

Fact 3.2.
  1. (1)

    Every definable set can be definably triangulated in T=RCFT=\operatorname{RCF}.

  2. (2)

    Local dimension is definable in TT.

  3. (3)

    The closure/interior/boundary/essence/essential boundary/essential interior are definable in TT.

  4. (4)

    For a definable subset XnX\subseteq\operatorname{\mathbb{R}}^{n}, the o-minimal dimension agrees with the Hausdorff dimension.

Statement (1) can be found in [17, Theorem 2.9, p.130]. Statement (2) follows easily from [17, Chapter 4], and (3) is a consequence of (2). Item (4) can be proven easily using cell decomposition. In fact, a much more general phenomenon is true; see [8, Corollary 1.6].

The following fact is well-known. A proof can be found, for example, from [10].

Fact 3.3.

Every connected open subset of x\operatorname{\mathbb{R}}^{x} is topologically connected after removing a subset of Hausdorff codimension 22.

The next lemma gives us a model-theoretic counterpart of Fact 3.3.

Lemma 3.4.

Every definably connected open subset of 𝐌x\boldsymbol{M}^{x} is definably connected by codimension 22.

Proof.

Let UU be a definably connected open subset of 𝑴x\boldsymbol{M}^{x}. Suppose to the contrary. Then we can obtain a definable CUC\subseteq U with codimension 22 in UU and a clopen subset WW of UCU\setminus C. Then we can obtain LL-formulas φ(x,y)\varphi(x,y), ψ(x,y)\psi(x,y), and θ(x,y)\theta(x,y) such that there is b𝑴yb\in\boldsymbol{M}^{y} with φ(x,b)\varphi(x,b) defining UU, ψ(x,b)\psi(x,b) defining CC, and θ(x,b)\theta(x,b) defining WW. Using the fact that TT is complete and dimensions are definable, we get UU^{\prime}, CC^{\prime}, and WW^{\prime} in \operatorname{\mathbb{R}} with similar properties. This is a contradiction to the Fact 3.3 and Fact 3.2(4). Hence, we have obtained the desired result. ∎

The next lemma is the key ingredient in proving Proposition 3.6.

Lemma 3.5.

Suppose X,EX,E are subsets of 𝐌x\boldsymbol{M}^{x} with XEX\subseteq E, dimX=dimE=n\dim X=\dim E=n are definable and EE is order-free and definably connected by codimension 22. Then either

dima(bdessE(X))=n1 for all abdessE(X)orbdessE(X)=.\dim_{a}(\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X))=n-1\text{ for all }a\in\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X)\quad\text{or}\quad\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X)=\emptyset.

Moreover, bdessE(X)=\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X)=\emptyset if and only if dim(EX)<n\dim(E\setminus X)<n.

Proof.

Note that DLO and DOAG are reducts of RCF\operatorname{RCF}, and for definable sets in DLO and DOAG, their o-minimal dimension are the same as their o-minimal dimension when viewed as definable sets in RCF\operatorname{RCF}. Hence, it suffices to consider the case where TT is RCF\operatorname{RCF}. We will first show that if dimX=dim(EX)=n\dim X=\dim(E\setminus X)=n, then dim(bdessE(X))=n1\dim(\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X))=n-1. Note that bdessE(X)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X) is a subset of the boundary of XX in the ambient space 𝑴x\boldsymbol{M}^{x}, so dim(bdessE(X))n1\dim(\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X))\leq n-1. Obtain a triangulation Δ\Delta of EE such that XX is the union of Δ1Δ\Delta_{1}\subseteq\Delta and EXE\setminus X is the union of Δ2Δ\Delta_{2}\subseteq\Delta. It suffices to show that there is an open simplex in Δ1\Delta_{1} and an open simplex in Δ2\Delta_{2} sharing a common (n1)(n-1)-face, as points on this common (n1)(n-1)-face are elements of bdessE(X)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X). Obtaining Δ\Delta^{\prime}, Δ1\Delta^{\prime}_{1}, and Δ2\Delta_{2}^{\prime} from Δ\Delta, Δ1\Delta_{1}, and Δ2\Delta_{2} by taking only the simplices of dimension n1\geq n-1. Set EE^{\prime}, E1E^{\prime}_{1}, and E2E^{\prime}_{2} be the union of Δ\Delta, Δ1\Delta^{\prime}_{1}, and Δ2\Delta^{\prime}_{2}. Let a1a_{1} be in E1E^{\prime}_{1}, and let a2a_{2} be in E2E^{\prime}_{2}. As EE is definably connected by codimension 22 and semialgebraic, EE^{\prime} is connected. Hence, there is a semialgebraic path p:[0,1]Ep:[0,1]\to E^{\prime} with p(0)=a1p(0)=a_{1} and p(1)=a2p(1)=a_{2}. Let t0=sup{tp[0,t]E1}t_{0}=\sup\{t\mid p[0,t]\subseteq E^{\prime}_{1}\}. Then t0t_{0} lies on a simplex in Δ\Delta^{\prime} which is a common face of a simplex in Δ1\Delta^{\prime}_{1} and a simplex in Δ2\Delta^{\prime}_{2}. This common face is an element in Δ\Delta^{\prime} and must have dimension n1n-1 as Δ\Delta^{\prime} contains no element of dimension <n1<n-1.

Suppose bdessE(X)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X)\neq\emptyset. Let aa be an element of bdessE(X)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X). Then we have dimXdimaXn\dim X\geq\dim_{a}X\geq n and dim(EX)dima(EX)n\dim(E\setminus X)\geq\dim_{a}(E\setminus X)\geq n. Hence, dimX=dim(EX)=n\dim X=\dim(E\setminus X)=n. The same argument as in the preceding paragraph can be carried out replacing EE by UU and XX by XUX\cap U, where UEU\subseteq E is any small open ball containing aa. This gives us the stronger conclusion dima(bdessE(X))=n1\dim_{a}(\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X))=n-1. If dim(EX)<n\dim(E\setminus X)<n, then bdessE(X)=\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X)=\emptyset. If dim(EX)n\dim(E\setminus X)\geq n, then the argument of the preceding paragraph shows that dim(bdessE(X))=n1\dim(\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{E}(X))=n-1. So we have obtained all the desired conclusions. ∎

We now prove the main result of this section.

Proposition 3.6.

Suppose TT is RCF\operatorname{RCF}, X𝐌xX\subseteq\boldsymbol{M}^{x} and Y𝐌yY\subseteq\boldsymbol{M}^{y} are definable, DX×YD\subseteq X\times Y is stable with dim(X)=|x|\dim(X)=|x|, dim(Y)=|y|\dim(Y)=|y|, and dim(D)=|x|+|y|\dim(D)=|x|+|y|. Then there exists a finite family (Xi,Yi)iI(X_{i},Y_{i})_{i\in I} of semialgebraic sets XiXX_{i}\subseteq X and YiYY_{i}\subseteq Y such that dimXi=|x|\dim X_{i}=|x| and dimYi=|y|\dim Y_{i}=|y| for each iIi\in I, and

dim(DiI(Xi×Yi))<|x|+|y|.\dim\big{(}D\triangle\bigcup\limits_{i\in I}(X_{i}\times Y_{i})\big{)}<|x|+|y|.
Proof.

In this proof, we will use the notion of tangent space and smoothness. This is defined as in [2]. Removing sets of smaller dimension from XX and YY, we can arrange that XX and YY are open subsets as subsets of 𝑴x\boldsymbol{M}^{x} and 𝑴y\boldsymbol{M}^{y}. By decomposing XX and YY into definably connected components, we may further assume that XX and YY are definably connected. Hence, X×YX\times Y is a definably connected open subset of 𝑴|x|+|y|\boldsymbol{M}^{|x|+|y|}. By Lemma 3.4, X×YX\times Y is definably connected by codimension 2. Set B=bdessX×Y(D)B=\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{X\times Y}(D). In the special case where BB is empty—by the last statement in Lemma 3.5—one can simply choose the family (Xi,Yi)iI(X_{i},Y_{i})_{i\in I} to consist of the single pair (X,Y)(X,Y) and get dim(D(X×Y))<|x|+|y|\dim(D\triangle(X\times Y))<|x|+|y|. We will reduce the general situation to this special case.

Suppose B=bdessX×Y(D)B=\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{X\times Y}(D) is non-empty. Lemma 3.5 then gives us dim(B)=|x|+|y|1\dim(B)=|x|+|y|-1. Now, decompose BB into a finite disjoint union of definable Euclidean open subsets of smooth varieties in the sense of algebraic geometry (i.e. the dimension of the tangent space at every point is the dimension of the variety; see [2, Definition 3.3.4]). Let BB^{*} be the set of points in the components with dimension |x|+|y|1|x|+|y|-1. Hence,

dimB=dimB=|x|+|y|1.\dim B^{*}=\dim B=|x|+|y|-1.

Suppose (a,b)X×Y(a,b)\in X\times Y is a point in BB^{*}. Let T(a,b)T_{(a,b)} be the tangent space at (a,b)(a,b) of BB^{*}. Then, T(a,b)T_{(a,b)} also has dimension |x|+|y|1|x|+|y|-1. Let TaT_{a} be the tangent space of XX at aa, and let TbT_{b} be the tangent space of YY at bb. Since X,YX,Y are open in 𝑴x\boldsymbol{M}^{x} and 𝑴y\boldsymbol{M}^{y}, TaT_{a} is an isomorphic copy of the vector space 𝑴x\boldsymbol{M}^{x} over the underlying field of 𝓜\boldsymbol{\mathscr{M}}, TbT_{b} is an isomorphic copy of the vector space 𝑴y\boldsymbol{M}^{y} over the underlying field 𝓜\boldsymbol{\mathscr{M}}, and T(a,b)T_{(a,b)} is a hyperplane in Ta×TbT_{a}\times T_{b}.

We will show that either T(a,b)=Sa×TbT_{(a,b)}=S_{a}\times T_{b} where SaS_{a} is a subset of TaT_{a} with dimSa=|x|1\dim S_{a}=|x|-1 or T(a,b)=Ta×SbT_{(a,b)}=T_{a}\times S_{b} where SbS_{b} is a subset of TbT_{b} with dimSb=|y|1\dim S_{b}=|y|-1. Suppose it is neither of the above, then the projection maps from T(a,b)T_{(a,b)} to TaT_{a} and TbT_{b} are surjective. We can then obtain a line L(a,b)L_{(a,b)} in T(a,b)T_{(a,b)}, such that the projection of L(a,b)L_{(a,b)} onto TaT_{a} and TbT_{b} has dimension 11. This translates to the existence of a curve CC in BB^{*} such that with πXC\pi_{X}C the projection of CC on XX and πYC\pi_{Y}C the projections of CC on YY, we have (πXC×πYC)D(\pi_{X}C\times\pi_{Y}C)\cap D and (πXC×πYC)D(\pi_{X}C\times\pi_{Y}C)\setminus D is homeomorphic to the situation in Example 1.1. More precisely, we obtain a continuous and increasing function f:[0,1][0,1]f:[0,1]\to[0,1] with f(0)=0f(0)=0 and f(1)=1f(1)=1, together with definable homemorphisms s:[0,1]πXCs:[0,1]\to\pi_{X}C and t:[0,1]πYCt:[0,1]\to\pi_{Y}C such that the image of the graph of ff under s×ts\times t is CC,

(s×t){(c,d)[0,1]2:f(c)>d}=D(πXC×πYC)(s\times t)\{(c,d)\in[0,1]^{2}:f(c)>d\}=D\cap(\pi_{X}C\times\pi_{Y}C)

and

(s×t){(c,d)[0,1]2:f(c)<d}=((X×Y)D)(πXC×πYC).(s\times t)\{(c,d)\in[0,1]^{2}:f(c)<d\}=((X\times Y)\setminus D)\cap(\pi_{X}C\times\pi_{Y}C).

Then s×ts\times t maps the witness of unstability from Example 1.1 to an unstable witness in DD, which is a contradiction.

Now, let (a,b)(a,b) range over BB^{*}, and set

DX={aπX(T(a,b))Ta}andDY={b|πY(T(a,b))Tb}.D_{X}=\{a\mid\pi_{X}(T_{(a,b)})\neq T_{a}\}\quad\text{and}\quad D_{Y}=\{b|\,\pi_{Y}(T_{(a,b)})\neq T_{b}\}.

It is easy to check that dimDX=|x|1\dim D_{X}=|x|-1 and dimDY=|y|1\dim D_{Y}=|y|-1. Take a cell decomposition of XX such that DXD_{X} is contained in the union of the cells of dimension |x|1\leqslant|x|-1, and a cell decomposition of YY such that DYD_{Y} is contained in the union of the cells of dimension |y|1\leqslant|y|-1. Let (Xi,Yi)iI(X_{i},Y_{i})_{i\in I} be the collection of products of cells of dimension |x||x| in the cell decomposition of XX and cells of dimension |y||y| in the cell decomposition of YY. It is easy to see that bdessXi×Yi(DXi×Yi)B.\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{X_{i}\times Y_{i}}(D\cap X_{i}\times Y_{i})\subseteq B. Moreover, as XiDX=X_{i}\cap D_{X}= and YiDY=Y_{i}\cap D_{Y}= for all iIi\in I, so Xi×YiB=X_{i}\times Y_{i}\cap B^{*}=\emptyset. Therefore,

bdessXi×Yi(DXi×Yi)BB.\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{X_{i}\times Y_{i}}(D\cap X_{i}\times Y_{i})\subseteq B\setminus B^{*}.

By general o-minimality knowledge, dimBB<dimB=|x|+|y|1.\dim B\setminus B^{*}<\dim B=|x|+|y|-1. Thus, bdessXi×Yi(DXi×Yi)\operatorname{\operatorname{bd}^{\operatorname{ess}}}_{X_{i}\times Y_{i}}(D\cap X_{i}\times Y_{i}) has dimension <|x|+|y|1<|x|+|y|-1 by construction, and is therefore empty by Lemma 3.5. We reduced the situation to the special case at the beginning of this lemma. ∎

In the remainder of this section, we will supose that TT is either DLO or DOAG and prove an analogue of Proposition 3.6. In a model of RCF, we have two natural topologies, namely, the Zariski topology and Euclidean topology, with the former coarser than the latter. In the same fashion, we can define on 𝑴x\boldsymbol{M}^{x} a topology coarser than the Euclidean topology. Let 𝒞\mathscr{C} denote the collection of sets defined by lattice combinations (i.e. positive boolean combinations) of sets defined by subsets of 𝑴x\boldsymbol{M}^{x} defined by equations. As TT is complete, we can arrange that 𝓜\boldsymbol{\mathscr{M}} is a reduct of a model of RCF\mathrm{RCF}. Recall that the Zariski topology in RCF is noetherian, so for any decreasing sequence (Cn)n{(C_{n})}_{n\in\operatorname{\mathbb{N}}} of elements in 𝒞\mathscr{C}, there is NN\in\operatorname{\mathbb{N}} such that CN=nCnC_{N}=\bigcap_{n}C_{n}. Hence, 𝒞\mathscr{C} is the collection of closed sets of certain noetherian topology which we call the linear topology on 𝑴x\boldsymbol{M}^{x}. As usual, an element XX of 𝒞\mathscr{C} is irreducible if XX cannot be written as a nontrivial union of two elements of 𝒞\mathscr{C}. It is then a standard fact about noetherian topology that every element of 𝒞\mathscr{C} can be uniquely (up to permutation) decomposed into a finite union of irreducible elements of 𝒞\mathscr{C} such that there is no containment between any two distinct elements.

Remark 3.7.

Suppose x=(x1,,xm)x=(x_{1},\ldots,x_{m}). If TT is DLO, an irreducible closed subset of 𝑴x\boldsymbol{M}^{x} is the solution set of a system where each equation is of either the form xi=xjx_{i}=x_{j} or the form xi=cx_{i}=c with ii and jj in {1,,m}\{1,\ldots,m\} and c𝑴c\in\boldsymbol{M}. If TT is DOAG\mathrm{DOAG}, an irreducible closed subset of 𝑴x\boldsymbol{M}^{x} is the solution set of a system consisting of equations of the form k1x1++kmxm=ck_{1}x_{1}+\cdots+k_{m}x_{m}=c with kik_{i} integers for i{1,,m}i\in\{1,\ldots,m\} and c𝑴c\in\boldsymbol{M}.

With TT still either DLO or DOAG, we will define suitable versions of tangent spaces and smoothness in this setting. Suppose that X𝑴xX\subseteq\boldsymbol{M}^{x} is definable. Let XcclX^{\mathrm{ccl}} be the closure of XX with respect to the (coarser) linear topology on 𝑴x\boldsymbol{M}^{x}, and let {Vi}in\{V_{i}\}_{i\leqslant n} lists all irreducible components of XcclX^{\mathrm{ccl}}. The tangent space of XX at aXa\in X is the smallest irreducible closed set containing {Vi|in,aVi}\bigcup\{V_{i}\;|\;i\leqslant n,\,a\in V_{i}\}. Remark 3.7 tells us that this a reasonable definition. Moreover, we say that aXa\in X is smooth if there is unique ViV_{i} such that aVia\in V_{i}. The proof of Proposition 3.6 with this new definition of smoothness and tangent space yields the Proposition 3.8 below. Note that this is in accordance with the notion of smoothness in algebraic geometry; again, see [2, Definition 3.3.4]. More precisely, in the case of DOAG, the ViV_{i}’s are all affine subspaces, hence algebraic varieties. All affine subspaces are in particular smooth varieties. Hence, a point aiVia\in\bigcup_{i}V_{i} is a smooth point if and only if there is a unique ii such that aVia\in V_{i}; a point lying on the intersection of two distinct irreducible components is not regular, hence not smooth.

Proposition 3.8.

Suppose TT is either DLO\mathrm{DLO} or DOAG\mathrm{DOAG}, X𝐌xX\subseteq\boldsymbol{M}^{x} and Y𝐌yY\subseteq\boldsymbol{M}^{y} are definable, DX×YD\subseteq X\times Y is stable with dim(X)=|x|\dim(X)=|x|, dim(Y)=|y|\dim(Y)=|y|, and dim(D)=|x|+|y|\dim(D)=|x|+|y|. Then there exists a finite family (Xi,Yi)iI(X_{i},Y_{i})_{i\in I} of definable sets XiXX_{i}\subseteq X and YiYY_{i}\subseteq Y such that dimXi=|x|\dim X_{i}=|x| and dimYi=|y|\dim Y_{i}=|y| for each iIi\in I, and

dim(DiIXi×Yi)<|x|+|y|.\dim\big{(}D\triangle\bigcup\limits_{i\in I}X_{i}\times Y_{i}\big{)}<|x|+|y|.

4. Stable formulas in DLO and DOAG

In this very short section, we classify the stable formulas in the case where TT is either DLO or DOAG.

Theorem 4.1.

Suppose TT is either DLO\mathrm{DLO} or DOAG\mathrm{DOAG} Then D𝐌(x;y)D\subseteq\boldsymbol{M}^{(x;y)} is stable if and only if DD is a finite union of special stable sets.

Proof.

Let ZZ be the closure of DD in the linear topology on 𝑴(x;y)\boldsymbol{M}^{(x;y)}. We argue by induction on the dimension of DD. If dim(D)=0\dim(D)=0, the conclusion is immediate. Since we are allowed to take finite unions in the theorem, we can arrange that ZZ is irreducible in the linear topology on 𝑴(x;y)\boldsymbol{M}^{(x;y)}. By Remark 3.7, there are subtuples of variables xx^{\prime} of xx and yy^{\prime} of yy such that the projection map πx,y:𝑴(x;y)𝑴(x;y)\pi_{x^{\prime},y^{\prime}}:\boldsymbol{M}^{(x;y)}\to\boldsymbol{M}^{(x^{\prime};y^{\prime})} induces an isomorphism between ZZ and 𝑴(x;y)\boldsymbol{M}^{(x^{\prime};y^{\prime})}. From Proposition 3.8, we get a finite set II, and XiX^{\prime}_{i} and YiY^{\prime}_{i} for each iIi\in I such that

dim(πx,y(D)iIXi×Yi)<dim(Z)=dim(D).\dim(\pi_{x^{\prime},y^{\prime}}(D)\triangle\bigcup_{i\in I}X^{\prime}_{i}\times Y^{\prime}_{i})<\dim(Z)=\dim(D).

Set Xi=πx1(Xi)X_{i}=\pi^{-1}_{x^{\prime}}(X^{\prime}_{i}) and Yi=πy1(Yi)Y_{i}=\pi^{-1}_{y^{\prime}}(Y^{\prime}_{i}) where πx:𝑴x𝑴x\pi_{x^{\prime}}:\boldsymbol{M}^{x}\to\boldsymbol{M}^{x^{\prime}} and πy:𝑴y𝑴y\pi_{y^{\prime}}:\boldsymbol{M}^{y}\to\boldsymbol{M}^{y^{\prime}} are the projection maps. Then the dimension of the symmetric difference between DD and Z(iI(Xi×Yi))Z\cap(\bigcup_{i\in I}(X_{i}\times Y_{i})) is strictly smaller than dim(D)\dim(D). Set WiW_{i} to be the order-free closure of (Z(Xi×Yi))D\big{(}Z\cap(X_{i}\times Y_{i})\big{)}\setminus D. Then for each iIi\in I, (ZWi)(Xi×Yi)(Z\setminus W_{i})\cap(X_{i}\times Y_{i}) is a special stable subset of DD. Moreover,

D(iI((ZWi)(Xi×Yi))D\setminus\left(\bigcup_{i\in I}((Z\setminus W_{i})\cap(X_{i}\times Y_{i})\right)

is stable and of smaller dimension. Hence, we can apply the inductive hypothesis and obtain the desired conclusion. ∎

One can observe that the above proof cannot be carried out when TT is RCF\operatorname{RCF} because if ZZ is an irreducible variety, one cannot choose subtuples xx^{\prime} of xx and yy^{\prime} of yy such that the projection map πx,y\pi_{x^{\prime},y^{\prime}} as defined in the proof induces an isomorphism.

5. Stable formulas in RCF

Finally, we are heading to the classification of stable formulas when TT is RCF\operatorname{RCF}. With caveats, the strategy is the same as the proof of Theorem 4.1, namely, taking projections to get largeness and then use Proposition 3.6. This is quite similar to the proof of Theorem 4.1 except that we need to work harder to arrange for the projection maps to be bijective. To this end, we will need to decompose the original sets into finitely many disjoint pieces. Hence, the need for topological compactness, which we get by passing to projective space through Lemma 5.1 and working with a fixed copy of \operatorname{\mathbb{R}} in the monster model. Another obstruction to finite decomposition comes from singularities, and this can be avoided by the use of Hironaka’s resolution of singularities at the cost of obtaining a classification only up to order-free isomorphisms.

In this section, we will use real algebraic geometry in the classical sense (i.e. zeros of systems of polynomial); see [2, Section 2, 3] for the precise set up. We let x\mathbb{P}^{x} denote of the |x||x|-dimensional projective space over 𝓜\boldsymbol{\mathscr{M}} corresponding to the tuple of variables xx. We will identify 𝑴x\boldsymbol{M}^{x} in the usual way with a Zariski open subset of x\mathbb{P}^{x}.

Lemma 5.1.

Let Z𝐌(x;y)Z\subseteq\boldsymbol{M}^{(x;y)} be an irreducible algebraic set. Then there are finite tuples xx^{\prime} and yy^{\prime} and an irreducible algebraic set Z𝐌(x;y)Z^{\prime}\subseteq\boldsymbol{M}^{(x^{\prime};y^{\prime})} satisfying the following properties:

  1. (1)

    there are rational maps f:𝑴x𝑴xf:\boldsymbol{M}^{x^{\prime}}\to\boldsymbol{M}^{x} and g:𝑴y𝑴yg:\boldsymbol{M}^{y^{\prime}}\to\boldsymbol{M}^{y} such that f×gf\times g is a birational morphism from ZZ^{\prime} to ZZ.

  2. (2)

    the Zariski closure of ZZ^{\prime} in x×y\mathbb{P}^{x^{\prime}}\times\mathbb{P}^{y^{\prime}} is smooth.

Proof.

View ZZ as a subset of the projective space x×y\mathbb{P}^{x}\times\mathbb{P}^{y}, and let ZprZ_{\text{pr}} be the Zariski closure of ZZ in x×y\mathbb{P}^{x}\times\mathbb{P}^{y}. Using Hironaka’s theorem on resolution of singularity [9], we obtain a finite tuple of variables zz and an irreducible algebraic set Wprx×y×zW_{\text{pr}}\subseteq\mathbb{P}^{x}\times\mathbb{P}^{y}\times\mathbb{P}^{z} such that the projection map from x×y×z\mathbb{P}^{x}\times\mathbb{P}^{y}\times\mathbb{P}^{z} to x×y\mathbb{P}^{x}\times\mathbb{P}^{y} induces a birational surjection from WprW_{\text{pr}} to ZprZ_{\text{pr}}. More precisely, the resolution of singularity is done using a sequence of blowups, so we can get the desired zz; see [1, Theorem A] for details. Choose a new tuple zz^{\prime} of variables with the same length as zz. Copy WprW_{\text{pr}} to x×z×y×z\mathbb{P}^{x}\times\mathbb{P}^{z}\times\mathbb{P}^{y}\times\mathbb{P}^{z^{\prime}} to get WprW^{\prime}_{\text{pr}}, or more precisely, let WprW^{\prime}_{\text{pr}} is the image of WprW_{\text{pr}} under the map

x×y×zx×z×y×z,(a,b,c)(a,c,b,c).\mathbb{P}^{x}\times\mathbb{P}^{y}\times\mathbb{P}^{z}\to\mathbb{P}^{x}\times\mathbb{P}^{z}\times\mathbb{P}^{y}\times\mathbb{P}^{z^{\prime}},(a,b,c)\mapsto(a,c,b,c).

Choose xx^{\prime} and yy^{\prime} such that x×z\mathbb{P}^{x}\times\mathbb{P}^{z} can be identified with a closed subset of x\mathbb{P}^{x^{\prime}}, and x×z\mathbb{P}^{x}\times\mathbb{P}^{z^{\prime}} can be identified with a closed subset of y\mathbb{P}^{y^{\prime}} via Segre embeddings. Identify x×z×y×z\mathbb{P}^{x}\times\mathbb{P}^{z}\times\mathbb{P}^{y}\times\mathbb{P}^{z^{\prime}} as a closed subset of x×y\mathbb{P}^{x^{\prime}}\times\mathbb{P}^{y^{\prime}}, and let ZprZ^{\prime}_{\text{pr}} be the image of WprW^{\prime}_{\text{pr}} under this identification. Identify 𝑴x\boldsymbol{M}^{x^{\prime}} with an affine piece of x\mathbb{P}^{x^{\prime}} and 𝑴y\boldsymbol{M}^{y^{\prime}} with an affine piece of y\mathbb{P}^{y^{\prime}} in such a way that with Z=Zpr𝑴xZ^{\prime}=Z^{\prime}_{\text{pr}}\cap\boldsymbol{M}^{x^{\prime}} and 𝑴y\boldsymbol{M}^{y^{\prime}}, we have ZZ^{\prime} is dense in ZprZ^{\prime}_{\text{pr}}. Condition (2) is satisfied as the closure of ZZ^{\prime} in x×y\mathbb{P}^{x^{\prime}}\times\mathbb{P}^{y^{\prime}} is ZprZ^{\prime}_{\text{pr}}, which is smooth. By choosing suitable affine pieces, we get rational maps f:xx×zf^{\prime}:\mathbb{P}^{x^{\prime}}\to\mathbb{P}^{x}\times\mathbb{P}^{z} and g:yy×zg^{\prime}:\mathbb{P}^{y^{\prime}}\to\mathbb{P}^{y}\times\mathbb{P}^{z^{\prime}} such that f×gf^{\prime}\times g^{\prime} induces a birational morphism from ZprZ^{\prime}_{\text{pr}} to WprW^{\prime}_{\text{pr}}. Let f=πxff=\pi_{x}\circ f^{\prime} and g=πygg=\pi_{y}\circ g^{\prime} where πx\pi_{x} and πy\pi_{y} are projection onto x\mathbb{P}^{x} and y\mathbb{P}^{y}. Then, ff and gg satisfy the condition specified in (1). ∎

For the next theorem, we are restricting our attention to a copy of the real numbers \mathbb{R} living in the monster model 𝓜\boldsymbol{\mathscr{M}}. The statement also holds if we replace \operatorname{\mathbb{R}} by an arbitary Archimedean subfield KK of 𝓜\boldsymbol{\mathscr{M}} as any such KK can be embedded into \mathbb{R} using an automorphism of 𝓜\boldsymbol{\mathscr{M}}.

Theorem 5.2.

A set D𝐌(x;y)D\subseteq\boldsymbol{M}^{(x;y)} definable over \mathbb{R}, is stable if and only if DD is a finite union of sets each order-free isomorphic over \mathbb{R} to a special stable set defined over \mathbb{R}.

Proof.

The backward direction follows from Lemma 2.4 and the well-known fact that Boolean combination preserves stability.

Now, suppose that D𝑴(x;y)D\subseteq\boldsymbol{M}^{(x;y)} is a stable set defined over \mathbb{R}. Let D()D(\operatorname{\mathbb{R}}) be the set of \operatorname{\mathbb{R}}-points of DD. By transfer principles, it suffices to show the forward direction of the theorem replacing 𝓜\boldsymbol{\mathscr{M}} with the field of real numbers, and DD with D()D(\operatorname{\mathbb{R}}). Moreover, we are not using the saturation of 𝓜\boldsymbol{\mathscr{M}} in the proof of the current theorem. Therefore, without loss of generality, we assume that 𝓜=(;+,×,<)\boldsymbol{\mathscr{M}}=(\operatorname{\mathbb{R}};+,\times,<) and D=D()D=D(\operatorname{\mathbb{R}}).

Let ZZ be the Zariski closure of DD in (x;y)\operatorname{\mathbb{R}}^{(x;y)}. If ZZ has dimension 0, the conclusion is immediate. By the fact that the stability is preserved under taking Boolean combinations, we can arrange that ZZ is irreducible. Using induction on dimension of ZZ, we can assume that the statement is already proven for all stable sets with dimension smaller than dim(Z)\dim(Z). This induction hypothesis, in particular, allows us to replace ZZ with ZZ^{\prime} as in Lemma 5.1, and so to arrange that the Zariski closure of ZZ in x×y\mathbb{P}^{x}\times\mathbb{P}^{y} is smooth.

Let (x;y)(x^{\prime};y^{\prime}) range over the pairs where xx^{\prime} and yy^{\prime} are subtuples of xx and yy such that |x|+|y|=dim(D)|x^{\prime}|+|y^{\prime}|=\dim(D), let πx,y:(x;y)(x;y)\pi_{x^{\prime},y^{\prime}}:\operatorname{\mathbb{R}}^{(x;y)}\to\operatorname{\mathbb{R}}^{(x^{\prime};y^{\prime})} be the projection map, and let U(x,y)U_{(x^{\prime},y^{\prime})} denote the maximal Zariski open subset of ZZ such that the restriction of πx,y\pi_{x^{\prime},y^{\prime}} to U(x,y)U_{(x^{\prime},y^{\prime})} induces an isomorphism on the tangent spaces. Note that this set is indeed Zariski open, since it is defined by some non-vanishing conditions on minors of the Jacobian matrix. Since ZZ is smooth,

Z=(x;y)U(x,y).Z=\bigcup_{(x^{\prime};y^{\prime})}U_{(x^{\prime},y^{\prime})}.

For a fixed (a,b)U(x,y)(a,b)\in U_{(x^{\prime},y^{\prime})}, by the definition of U(x,y)U_{(x^{\prime},y^{\prime})} and the inverse function theorem, the map πx,y\pi_{x^{\prime},y^{\prime}} induces a homeomorphism from a neighborhood of (a,b)(a,b) in ZZ onto its image with respect to the Euclidean topology. Therefore, we can obtain a set Bx,ya,b(x;y)B_{x^{\prime},y^{\prime}}^{a,b}\subseteq\operatorname{\mathbb{R}}^{(x;y)} containing (a,b)(a,b) such that Bx,ya,bB_{x^{\prime},y^{\prime}}^{a,b} is a cartersian product of open intervals in \operatorname{\mathbb{R}}, and the restriction of πx,y\pi_{x^{\prime},y^{\prime}} to U(x,y)Bx,ya,bU_{(x^{\prime},y^{\prime})}\cap B_{x^{\prime},y^{\prime}}^{a,b} is one-to-one.

We consider first the case where DD is a subset of [1,1](x;y)[-1,1]^{(x;y)}. As (x;y)(x^{\prime};y^{\prime}) ranges over the pairs specified earlier, and (a,b)(a,b) ranges over U(x,y)[1,1](x;y)U_{(x^{\prime},y^{\prime})}\cap[-1,1]^{(x;y)}, the sets U(x,y)a,b:=U(x,y)Bx,ya,b[1,1](x;y)U^{a,b}_{(x^{\prime},y^{\prime})}:=U_{(x^{\prime},y^{\prime})}\cap B^{a,b}_{x^{\prime},y^{\prime}}\cap[-1,1]^{(x;y)} form an open cover of Z[1,1](x;y)Z\cap[-1,1]^{(x;y)}. The intersection Z[1,1](x;y)Z\cap[-1,1]^{(x;y)} is closed, and hence compact. Hence, we can obtain a finite cover (Ui)iI(U_{i})_{i\in I} of Z[1,1](x;y)Z\cap[-1,1]^{(x^{\prime};y^{\prime})} where each UiU_{i} is of the form U(x,y)a,bU_{(x^{\prime},y^{\prime})}^{a,b}with (a,b)Z[1,1](x;y)(a,b)\in Z\cap[-1,1]^{(x;y)} and (x;y)(x^{\prime};y^{\prime}) as specified earlier.

It is sufficient to show that for each iIi\in I, DUiD\cap U_{i} is a Boolean combination of sets algebraically isomorphic to special stable sets. Fix iIi\in I such that DUiD\cap U_{i}\neq\emptyset. Note that DUiD\cap U_{i} is the intersection of the stable set DD with an algebraic set and a rectangular set, so DUiD\cap U_{i} is a stable subset of the special stable set UiU_{i}. On the other hand, the restriction to UiU_{i} of the projection map πx,y\pi_{x^{\prime},y^{\prime}} is an isomorphism on tangent spaces and one-to-one. As a consequence, πx,y(DUi)\pi_{x^{\prime},y^{\prime}}(D\cap U_{i}) is stable in (x;y)(x^{\prime};y^{\prime}) as any unstable witness in the image can be pulled back. Applying Proposition 3.6, we get that πx,y(DUi)\pi_{x^{\prime},y^{\prime}}(D\cap U_{i}) is up to a set of dimension smaller than |x|+|y|=dim(D)|x^{\prime}|+|y^{\prime}|=\dim(D) a finite union EE^{\prime} of sets which are products of definable sets in x\operatorname{\mathbb{R}}^{x^{\prime}} and definable sets in y\operatorname{\mathbb{R}}^{y^{\prime}}. Taking the preimage under πx,y\pi_{x^{\prime},y^{\prime}}, we learn that the symmetric difference between DUiD\cap U_{i} and E:=Uiπx,y1(E)E:=U_{i}\cap\pi_{x^{\prime},y^{\prime}}^{-1}(E^{\prime}) has dimension smaller than |x|+|y|=dim(D)|x^{\prime}|+|y^{\prime}|=\dim(D). Note that

DUi=[E((UiD)E)][(UiE)D].D\cap U_{i}=[E\setminus\big{(}(U_{i}\setminus D)\cap E\big{)}]\;\cup\;[(U_{i}\setminus E)\cap D].

The two sets (UiD)E(U_{i}\setminus D)\cap E and (UiE)D(U_{i}\setminus E)\cap D have their union the symmetric difference between DUiD\cap U_{i} and EE, so they are stable sets with dimension smaller than dim(D)\dim(D). By the inductive hypothesis, (UiD)E(U_{i}\setminus D)\cap E and (UiE)D(U_{i}\setminus E)\cap D are Boolean combinations of sets order-free isomorphic to special stable sets. As a consequence, DUiD\cap U_{i} is also of the desired form.

We reduce the general situation to the special case with D[1,1](x;y)D\subseteq[-1,1]^{(x;y)}. We construe (x;y)\operatorname{\mathbb{R}}^{(x;y)} as an affine open subset of x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y}. For a transition map tt from (x;y)\operatorname{\mathbb{R}}^{(x;y)} to another affine open subset t(x;y)\operatorname{\mathbb{R}}_{t}^{(x;y)} of x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y}, set DtD_{t} to be t(Domain(t)D)t(\text{Domain}(t)\cap D). As tt is an order-free isomorphism from (x;y)\operatorname{\mathbb{R}}^{(x;y)} to t(x;y)\operatorname{\mathbb{R}}_{t}^{(x;y)}, DtD_{t} as a subset of the t(x;y)\operatorname{\mathbb{R}}_{t}^{(x;y)} is again stable. Moreover, if dimDt=dimD\dim D_{t}=\dim D, then by our arrangement on ZZ, the Zariski closure of DtD_{t} is smooth. For (a,b)(x;y)(a,b)\in\operatorname{\mathbb{R}}^{(x;y)}, the interiors of (a,b)+[1,1](x,y)(a,b)+[-1,1]^{(x,y)} form an open cover of (x;y)\operatorname{\mathbb{R}}^{(x;y)}. Doing the same for the other affine pieces of x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y} combine them together, we get a family of closed subset of x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y} whose interiors form an open cover of x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y}. As x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y} is compact, we obtain a finite subfamily (Vj)jJ(V_{j})_{j\in J} where VjV_{j} is of the form (aj,bj)+[1,1](x;y)(a_{j},b_{j})+[-1,1]^{(x;y)} on the affine piece tj(x;y)\operatorname{\mathbb{R}}_{t_{j}}^{(x;y)}with the transition map tjt_{j}, and the interiors of the members of (Vj)jJ(V_{j})_{j\in J} form an open cover of x×y\operatorname{\mathbb{P}}^{x}\times\operatorname{\mathbb{P}}^{y}. In particular, gives us that

D=jJtj1(DtjVj).D=\bigcup_{j\in J}t_{j}^{-1}(D_{t_{j}}\cap V_{j}).

Hence, it suffices to show that for each jJj\in J, tj1(DtjVj)t_{j}^{-1}(D_{t_{j}}\cap V_{j}) is a finite union of special stable sets. When dim(Dt)=dim(D)\dim(D_{t})=\dim(D), DtjVjD_{t_{j}}\cap V_{j} is a finite union of special stable sets in tj(x;y)\operatorname{\mathbb{R}}_{t_{j}}^{(x;y)} by the aforementioned special case. When dim(Dt)<dim(D)\dim(D_{t})<\dim(D), the desired conclusion follows from the induction hypothesis. As the transition map induces order-free isomorphism where it is defined, we get the desired conclusion. ∎

Acknowledgements

We would like to thank Lou van den Dries, Anand Pillay and Sergei Starchenko for several helpful discussions. We also thank the anonymous referee for many useful comments, and especially for pointing out a mistake in the earlier version.

References

  • [1] Edward Bierstone and Pierre D. Milman. Resolution of Singularities. available on https://arxiv.org/abs/alg-geom/9709028.
  • [2] Jacek Bochnak, Michel Coste, and Marie-Francoise Roy. Real Algebraic Geometry. Springer-Verlag, 1998.
  • [3] Artem Chernikov and Sergei Starchenko. A note on the Erd s-Hajnal property for stable graphs. Proc. Amer. Math. Soc., 146:785–790, 2018.
  • [4] Artem Chernikov and Sergei Starchenko. Regularity lemma for distal structures. J. Eur. Math. Soc. (JEMS), 20(10):2437–2466, 2018.
  • [5] Artem Chernikov and Sergei Starchenko. Definable Regularity Lemmas for NIP Hypergraphs. The Quarterly Journal of Mathematics, 03 2021. haab011.
  • [6] G. Conant, A. Pillay, and Terry C. Structure and regularity for subsets of groups with finite VC-dimension. available on https://arxiv.org/abs/1802.04246.
  • [7] Clifton Ealy and Alf Onshuus. þ\th-Forking and Stable Forking. Revista de la Academia Colombiana de Ciencias Exactas, FÃsicas y Naturales, 40:683 – 689, 12 2016.
  • [8] Philipp Hieronymi and Chris Miller. Metric dimensions and tameness in expansions of the real field. Trans. Amer. Math. Soc., 373(2):849–874, 2020.
  • [9] Heisuke Hironaka. Resolution of singularities of an algebraic variety over a field of characteristic zero: I. Annals of Mathematics, 79(1):109–203, 1964.
  • [10] Del (https://math.stackexchange.com/users/366527/del). Can a set of hausdorff codimension 2 disconnect a connected open set? Stackexchange. URL:https://math.stackexchange.com/questions/2730111/can-a-set-of-hausdorff-codimension-2-disconnect-a-connected-open-set.
  • [11] William Andrew Johnson. Fun with Fields. PhD thesis, University of California, Berkeley, 2016.
  • [12] Byunghan Kim. Simplicity, and stability in there. The Journal of Symbolic Logic, 66(2):822–836, 2001.
  • [13] Byunghan Kim and Anand Pillay. Around stable forking. Fundamenta Mathematicae, 170(1-2):107–118, 2001.
  • [14] M. Malliaris and S. Shelah. Regularity lemmas for stable graphs. Trans. Amer. Math. Soc., 366:1551–1585, 2014.
  • [15] Anand Pillay. Geometric Stability Theory. Oxford Logic Guides. Clarendon Press, 1996.
  • [16] Katrin Tent and Martin Ziegler. A Course in Model Theory. Lecture Notes in Logic. Cambridge University Press, 2012.
  • [17] Lou van den Dries. Tame topology and o-minimal structures, volume 248 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1998.