Stable formulas in ordered structures
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.
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 definable in the ordered field such that and , and take
We will construct a sequence such that if and only if . We start with any . Take such that , and and continue this way. Such a sequence can be similarly produced if we replace by a decreasing function. On the other hand, it is easy to see that there is no such sequence such that if and only if when is the graph of or the entire set .
For the rest of the paper, let be either the theory of dense linear orderings in (DLO), the theory of divisible ordered abelian groups in (DOAG), or the theory of real closed fields in (RCF), let be a model of with underlying set , let be an -formula, and let denotes the o-minimal dimension of ; see [17] for the basic definition and results. When we say that is stable, we implicitly assume that stability is with respect to the pair . We say that is rectangular if is -equivalent to with and being -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 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 is stable, and . Then there is an -formula which is a disjuntion of rectangular -formulas such that .
Proposition 1.2 is not sufficient for the purpose of classifying stable formulas in as it says nothing when . We say that is order-free if is equivalent over to quantifier-free -formulas which do not contain . Order-free formulas form another natural class of stables formulas. It is possible to have order-free with . 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 -formula is equivalent over 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 is either or and is stable. Then is equivalent over to a disjunction of special stable -formulas.
We expect that a result in this line can be obtained for more general linear orderings in and linearly ordered abelian groups in . 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]](https://cdn.awesomepapers.org/papers/9890872c-d871-458e-a97e-994ff81865c7/curve.jpg)
rCl 0 &= x^4+2x^2y^2+y^4+x^3-xy^2;
0 ⩽ x;
0 ⩽ y;
3x ⩽ y.
Let with be the conjunction of the above equations and inequalities. It is easy to check that is stable; in fact, every formula in two variables defining a one dimensional set is stable by cell decomposition. Note that the equation defines an irreducible algebraic set of dimension . Hence, if is a disjunction of special stable formulas, we can further arrange that each of these special stable formulas is a conjunction of with a rectangular -formula. As is o-minimal, we can arrange that each of these rectangular formulas defines a set of the form where are intervals. However, any set of the aforementioned form containing the point will have to include a part of the curve below the dashed line . Thus, is not a disjunction of special stable -formulas.
In Example 1.4, the obstacle in expressing as a finite disjunction of special stable formulas comes from the singularity at 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 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 and a relation defined by a formula . 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 is . Then an -formula is stable if and only if it is equivalent over to a disjunction of formulas order-free isomorphic to a special stable -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 is a formula with parameters over an Archimedean subfield of . 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 . In a parallel direction, some natural subsequent questions could be the classification of stable formulas in and , multi-ordered fields ([11]) or .
Notations and conventions
Throughout and are in . We adopt the usual monster model gadgets: is a -saturated and strongly -homogeneous with underlying set , we refer to definable with parameters in as simply definable, parameter sets and are assumed to be small with respect to , and other models are assumed to be elementary submodels of . Moreover, we identify a formula with the set it defines in , and work semantically with definable sets instead of formulas.
In the paper, we use the following convention about variables: Throughout and are finite tuples of variables. Let denote the cartesian power of indexed by the variables in . By we denote the tuple extended by the tuple . With an eye on stability, we also use the usual notational convention to emphasize the division of variables. So when we write , we mean and we have a fixed division of variables .
2. Preliminaries
We continue to let be , , or . Let be a model of with underlying set , let be -definable, and let denote the o-minimal dimension of . Recall that is unstable (with respect to the variable division ) if some (every) formula defining is unstable. More explicitly, such is unstable if there is such that if and only if . We call such an unstable witness of . As is complete, classifying stable formulas over is the same as classifying stable sets over .
Toward classifying stable sets, we consider several classes of sets which are obviously stable. We say that 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 without the linear order. A typical example of an order-free definable set is an algebraic set in . We say a function is order-free definable if the graph of is an order-free definable. Abusing notation, the restriction of to a subset of 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 is rectangular if and only if with and definable in . 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 is a special stable set if is defined by a special stable formula (with respect to the division of variables ), equivalently, there exist definable sets and , and an order-free definable set such that .
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 where and are definable, and is order-free definable. Note that
, and 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 , , and are the projection maps, and and are defined similarly. We say that a map is an order-free morphism if there exist
-
(1)
a finite covering of by order-free definable sets,
-
(2)
a finite covering of by order-free definable sets,
-
(3)
a family of order-free definable maps ,
-
(4)
a family of order-free definable maps
such that for each we have that , and that . If, moreover, the map is a bijection for each , then we call an order-free isomorphism. If there exists an order-free isomorphism between and , we say that is order-free isomorphic to .
Because of the preservation of stability under Boolean combinations, one can deduce the following easy fact: if the definable sets are stable and is definable, then is stable if and only if is stable for each . This easy fact leads to the following, more important, observation:
Lemma 2.4.
Suppose the definable sets and are order-free isomorphic. If is stable, then is stable.
Proof.
Let , , and , for , be as in Definition 2.3. Suppose that is unstable. From the fact preceding the lemma, we see that is unstable for some . Let be an unstable witness of . By Ramsey’s theorem [16, Theorem 5.1.5] there exists an infinite set such that for all we have by the fact that is order-free. Without loss of generality, we assume that .
Consider , , where . Because for all , we have that
If then and thus . If , then . As is injective over , we have that . Therefore, is an unstable witness of . ∎
3. Large stable sets
Our strategy of classifying stable sets relies on an induction on . The goal of this section is to obtain a description of stable sets with . 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 is definably connected by codimension if for all definable with , the set is definable connected in the o-minimal structure . Before moving forward, let us recall a few basic facts from o-minimal structures, see [17] for details on this subject.
Suppose is a definable subset of . For , the local dimension at of is defined as
where the “” 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 is an order-free definable subset of with nonempty . As usual, the closure of is the smallest closed subsets in containing , the interior of in is the largest open subset of that contains, and the boundary is defined as . The essence of in is the set
The essential interior of in is the set
The essential boundary of in is the set
Note that . We will omit “” if for some natural number .
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 be the open unit ball in with the equator, i.e.,
and . In this case, is the unit closed ball. And is the unit sphere and is the open unit ball. In particular, has dimension , which is in accordance with Lemma 3.5.
We collect some well known facts about o-minimal theories.
Fact 3.2.
-
(1)
Every definable set can be definably triangulated in .
-
(2)
Local dimension is definable in .
-
(3)
The closure/interior/boundary/essence/essential boundary/essential interior are definable in .
-
(4)
For a definable subset , 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 is topologically connected after removing a subset of Hausdorff codimension .
The next lemma gives us a model-theoretic counterpart of Fact 3.3.
Lemma 3.4.
Every definably connected open subset of is definably connected by codimension .
Proof.
Let be a definably connected open subset of . Suppose to the contrary. Then we can obtain a definable with codimension in and a clopen subset of . Then we can obtain -formulas , , and such that there is with defining , defining , and defining . Using the fact that is complete and dimensions are definable, we get , , and in 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 are subsets of with , are definable and is order-free and definably connected by codimension . Then either
Moreover, if and only if .
Proof.
Note that DLO and DOAG are reducts of , 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 . Hence, it suffices to consider the case where is . We will first show that if , then . Note that is a subset of the boundary of in the ambient space , so . Obtain a triangulation of such that is the union of and is the union of . It suffices to show that there is an open simplex in and an open simplex in sharing a common -face, as points on this common -face are elements of . Obtaining , , and from , , and by taking only the simplices of dimension . Set , , and be the union of , , and . Let be in , and let be in . As is definably connected by codimension and semialgebraic, is connected. Hence, there is a semialgebraic path with and . Let . Then lies on a simplex in which is a common face of a simplex in and a simplex in . This common face is an element in and must have dimension as contains no element of dimension .
Suppose . Let be an element of . Then we have and . Hence, . The same argument as in the preceding paragraph can be carried out replacing by and by , where is any small open ball containing . This gives us the stronger conclusion . If , then . If , then the argument of the preceding paragraph shows that . So we have obtained all the desired conclusions. ∎
We now prove the main result of this section.
Proposition 3.6.
Suppose is , and are definable, is stable with , , and . Then there exists a finite family of semialgebraic sets and such that and for each , and
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 and , we can arrange that and are open subsets as subsets of and . By decomposing and into definably connected components, we may further assume that and are definably connected. Hence, is a definably connected open subset of . By Lemma 3.4, is definably connected by codimension 2. Set . In the special case where is empty—by the last statement in Lemma 3.5—one can simply choose the family to consist of the single pair and get . We will reduce the general situation to this special case.
Suppose is non-empty. Lemma 3.5 then gives us . Now, decompose 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 be the set of points in the components with dimension . Hence,
Suppose is a point in . Let be the tangent space at of . Then, also has dimension . Let be the tangent space of at , and let be the tangent space of at . Since are open in and , is an isomorphic copy of the vector space over the underlying field of , is an isomorphic copy of the vector space over the underlying field , and is a hyperplane in .
We will show that either where is a subset of with or where is a subset of with . Suppose it is neither of the above, then the projection maps from to and are surjective. We can then obtain a line in , such that the projection of onto and has dimension . This translates to the existence of a curve in such that with the projection of on and the projections of on , we have and is homeomorphic to the situation in Example 1.1. More precisely, we obtain a continuous and increasing function with and , together with definable homemorphisms and such that the image of the graph of under is ,
and
Then maps the witness of unstability from Example 1.1 to an unstable witness in , which is a contradiction.
Now, let range over , and set
It is easy to check that and . Take a cell decomposition of such that is contained in the union of the cells of dimension , and a cell decomposition of such that is contained in the union of the cells of dimension . Let be the collection of products of cells of dimension in the cell decomposition of and cells of dimension in the cell decomposition of . It is easy to see that Moreover, as and for all , so . Therefore,
By general o-minimality knowledge, Thus, has dimension 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 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 a topology coarser than the Euclidean topology. Let denote the collection of sets defined by lattice combinations (i.e. positive boolean combinations) of sets defined by subsets of defined by equations. As is complete, we can arrange that is a reduct of a model of . Recall that the Zariski topology in RCF is noetherian, so for any decreasing sequence of elements in , there is such that . Hence, is the collection of closed sets of certain noetherian topology which we call the linear topology on . As usual, an element of is irreducible if cannot be written as a nontrivial union of two elements of . It is then a standard fact about noetherian topology that every element of can be uniquely (up to permutation) decomposed into a finite union of irreducible elements of such that there is no containment between any two distinct elements.
Remark 3.7.
Suppose . If is DLO, an irreducible closed subset of is the solution set of a system where each equation is of either the form or the form with and in and . If is , an irreducible closed subset of is the solution set of a system consisting of equations of the form with integers for and .
With still either DLO or DOAG, we will define suitable versions of tangent spaces and smoothness in this setting. Suppose that is definable. Let be the closure of with respect to the (coarser) linear topology on , and let lists all irreducible components of . The tangent space of at is the smallest irreducible closed set containing . Remark 3.7 tells us that this a reasonable definition. Moreover, we say that is smooth if there is unique such that . 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 ’s are all affine subspaces, hence algebraic varieties. All affine subspaces are in particular smooth varieties. Hence, a point is a smooth point if and only if there is a unique such that ; a point lying on the intersection of two distinct irreducible components is not regular, hence not smooth.
Proposition 3.8.
Suppose is either or , and are definable, is stable with , , and . Then there exists a finite family of definable sets and such that and for each , and
4. Stable formulas in DLO and DOAG
In this very short section, we classify the stable formulas in the case where is either DLO or DOAG.
Theorem 4.1.
Suppose is either or Then is stable if and only if is a finite union of special stable sets.
Proof.
Let be the closure of in the linear topology on . We argue by induction on the dimension of . If , the conclusion is immediate. Since we are allowed to take finite unions in the theorem, we can arrange that is irreducible in the linear topology on . By Remark 3.7, there are subtuples of variables of and of such that the projection map induces an isomorphism between and . From Proposition 3.8, we get a finite set , and and for each such that
Set and where and are the projection maps. Then the dimension of the symmetric difference between and is strictly smaller than . Set to be the order-free closure of . Then for each , is a special stable subset of . Moreover,
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 is because if is an irreducible variety, one cannot choose subtuples of and of such that the projection map as defined in the proof induces an isomorphism.
5. Stable formulas in RCF
Finally, we are heading to the classification of stable formulas when is . 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 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 denote of the -dimensional projective space over corresponding to the tuple of variables . We will identify in the usual way with a Zariski open subset of .
Lemma 5.1.
Let be an irreducible algebraic set. Then there are finite tuples and and an irreducible algebraic set satisfying the following properties:
-
(1)
there are rational maps and such that is a birational morphism from to .
-
(2)
the Zariski closure of in is smooth.
Proof.
View as a subset of the projective space , and let be the Zariski closure of in . Using Hironaka’s theorem on resolution of singularity [9], we obtain a finite tuple of variables and an irreducible algebraic set such that the projection map from to induces a birational surjection from to . More precisely, the resolution of singularity is done using a sequence of blowups, so we can get the desired ; see [1, Theorem A] for details. Choose a new tuple of variables with the same length as . Copy to to get , or more precisely, let is the image of under the map
Choose and such that can be identified with a closed subset of , and can be identified with a closed subset of via Segre embeddings. Identify as a closed subset of , and let be the image of under this identification. Identify with an affine piece of and with an affine piece of in such a way that with and , we have is dense in . Condition (2) is satisfied as the closure of in is , which is smooth. By choosing suitable affine pieces, we get rational maps and such that induces a birational morphism from to . Let and where and are projection onto and . Then, and satisfy the condition specified in (1). ∎
For the next theorem, we are restricting our attention to a copy of the real numbers living in the monster model . The statement also holds if we replace by an arbitary Archimedean subfield of as any such can be embedded into using an automorphism of .
Theorem 5.2.
A set definable over , is stable if and only if is a finite union of sets each order-free isomorphic over to a special stable set defined over .
Proof.
The backward direction follows from Lemma 2.4 and the well-known fact that Boolean combination preserves stability.
Now, suppose that is a stable set defined over . Let be the set of -points of . By transfer principles, it suffices to show the forward direction of the theorem replacing with the field of real numbers, and with . Moreover, we are not using the saturation of in the proof of the current theorem. Therefore, without loss of generality, we assume that and .
Let be the Zariski closure of in . If has dimension , the conclusion is immediate. By the fact that the stability is preserved under taking Boolean combinations, we can arrange that is irreducible. Using induction on dimension of , we can assume that the statement is already proven for all stable sets with dimension smaller than . This induction hypothesis, in particular, allows us to replace with as in Lemma 5.1, and so to arrange that the Zariski closure of in is smooth.
Let range over the pairs where and are subtuples of and such that , let be the projection map, and let denote the maximal Zariski open subset of such that the restriction of to 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 is smooth,
For a fixed , by the definition of and the inverse function theorem, the map induces a homeomorphism from a neighborhood of in onto its image with respect to the Euclidean topology. Therefore, we can obtain a set containing such that is a cartersian product of open intervals in , and the restriction of to is one-to-one.
We consider first the case where is a subset of . As ranges over the pairs specified earlier, and ranges over , the sets form an open cover of . The intersection is closed, and hence compact. Hence, we can obtain a finite cover of where each is of the form with and as specified earlier.
It is sufficient to show that for each , is a Boolean combination of sets algebraically isomorphic to special stable sets. Fix such that . Note that is the intersection of the stable set with an algebraic set and a rectangular set, so is a stable subset of the special stable set . On the other hand, the restriction to of the projection map is an isomorphism on tangent spaces and one-to-one. As a consequence, is stable in as any unstable witness in the image can be pulled back. Applying Proposition 3.6, we get that is up to a set of dimension smaller than a finite union of sets which are products of definable sets in and definable sets in . Taking the preimage under , we learn that the symmetric difference between and has dimension smaller than . Note that
The two sets and have their union the symmetric difference between and , so they are stable sets with dimension smaller than . By the inductive hypothesis, and are Boolean combinations of sets order-free isomorphic to special stable sets. As a consequence, is also of the desired form.
We reduce the general situation to the special case with . We construe as an affine open subset of . For a transition map from to another affine open subset of , set to be . As is an order-free isomorphism from to , as a subset of the is again stable. Moreover, if , then by our arrangement on , the Zariski closure of is smooth. For , the interiors of form an open cover of . Doing the same for the other affine pieces of combine them together, we get a family of closed subset of whose interiors form an open cover of . As is compact, we obtain a finite subfamily where is of the form on the affine piece with the transition map , and the interiors of the members of form an open cover of . In particular, gives us that
Hence, it suffices to show that for each , is a finite union of special stable sets. When , is a finite union of special stable sets in by the aforementioned special case. When , 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. -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.