Splitting differential equations using Galois theory
Abstract.
This article is interested in pullbacks under the logarithmic derivative of algebraic ordinary differential equations. In particular, assuming the solution set of an equation is internal to the constants, we would like to determine when its pullback is itself internal to the constants. To do so, we develop, using model-theoretic Galois theory and differential algebra, a connection between internality of the pullback and the splitting of a short exact sequence of algebraic Galois groups. We then use algebraic group theory to obtain internality and non-internality results.
Key words and phrases:
geometric stability, differentially closed fields, internality to the constants, logarithmic derivative2020 Mathematics Subject Classification:
03C45, 03C98, 12H05, 12L121. Introduction
Differential algebra and model theory have a long history of fruitful interaction, starting with the work of Robinson [22], where differentially closed fields were introduced. Both disciplines inform each other: the theory of differentially closed fields is a rich source of examples in model theory, and model theoretic techniques can provide useful tools for the study of algebraic differential equations. A reason for this is -stability of the theory of differentially closed fields of characteristic zero, which brings all the machinery of finite rank stability theory into the picture.
This article is concerned with one such tool, the semi-minimal analysis, and the development of Galois theoretic methods to control it, with an eye towards differential-algebraic applications. Recall that the semi-minimal analysis of a type is, roughly, a sequence of -definable maps such that , along with each fiber of an , is semi-minimal, meaning internal to some rank one type. That this can be done in gives a powerful way to decompose algebraic differential equations.
To make this analysis useful, we need to understand semi-minimal types. For algebraic differential equations, this is done via Zilber’s dichotomy for : a semi-minimal type is either internal to a locally modular minimal type, and in particular has a rudimentary geometric structure, or is internal to the constants. We will be interested in the second possibility, of which we now give a geometrical account.
A finite dimensional type , over some algebraically closed differential field , is always interdefinable with the generic type of an algebraic vector field . Internality to the constants means that is, after possibly taking extensions to a differential field , birationally equivalent to the trivial vector field . Equivalently, after a base change, there are algebraically independent rational functions on that are constant for the induced derivation (but are not elements of ). These functions are called first integrals in the literature. Note that if we only ask for the existence of one such first integral, we obtain the notion of non-orthogonality to the constants. See [17] for more on this perspective.
Even more concretely, the solution set of a differential equation (E) is internal to the constants if, roughly, there are finitely many solutions of (E) such that any other solution can be given as a rational function of the and finitely many constants. This is sometimes called a superposition principle in the literature (see [11] for example), and generalizes the fact that solution sets of linear differential equations are vectors spaces over the constant field.
A first step to understand semi-minimal analyses is to study a fibration with internal fibers, i.e. a definable map with all its fibers internal to a fixed definable set . In the case of , we will always take to be the field of constants. A starting question would be: suppose that is also -internal, when is internal to ? This was one of the motivations of the articles [5], [7] and [10]. In [5], a necessary and sufficient condition is given in the form of uniform internality: there are parameters (or a field extension in the concrete case of ) witnessing internality of all the fibers at once.
As a guiding example, fix some differential field , some and consider:
as well as the following system of differential equations:
On non-zero , there is a logarithmic derivative map which takes a solution of this system to a solution of (1). It is well-known (and easy to prove) that any fiber of is internal to the constants, and thus this fits into our general setup. We want to understand (1) by seeing it as fibered over . In particular, we want to answer:
Question 1.
From the point of view of vector fields, we know that each fiber of has a first integral, and equation (1) has first integrals by assumption. What the question is asking is whether from these, we can obtain first integrals to equation (1). There is no reason for this to be true in general, and many counterexamples are given in [10]. As explained before, this also has direct impact on our ability to recover solutions of equation (1) from the data of finitely many fixed solutions.
Note that because arbitrary field extensions are allowed, there is a priori no effective way to check if a given equation is internal to the constants. This is a general phenomenon in model-theoretic differential algebra: many useful concepts involve taking arbitrary field extensions, and controlling which extension one needs to look at has been an important theme. See for example recent work in [1] and [2] on the degree of non-minimality, which led to the development of effective tools for the study of algebraic differential equations.
Here, we will achieve some effectivity by considering a sufficient condition for internality, introduced in [10] and which we will call splitting. Going back to the general case of a fibration , we say that splits if there is in -definable bijection between and the Morley product , for some -internal type , and the diagram
commutes. If we replace the bijection by a finite-to-finite correspondence, we obtain almost splitting. Note that there is no need to take extra parameters anymore. In the particular case of differential equations, it removes the need to take an arbitrary field extension. One of the pay-offs for understanding splitting is criteria for internality of algebraic differential equations, for example [10, Theorem B]. We will likewise produce such criteria here.
One of the main tools of [5] and [10] is that, under certain conditions, uniform internality and splitting are equivalent. Then, one can use algebraic characterizations of splitting to obtain strong restrictions on when a fibration can be uniformly internal. This is also at the heart of our methods.
Question 1 has been asked for in [10], but we remove that restriction here. One issue with picking (and working over possibly non-constant parameters) is that we loose a convenient criteria due to Rosenlicht [23] that tells us exactly when equation (1) is internal to the constants. To the authors’ knowledge, no such criteria is known without these assumptions.
To answer Question 1, we proceed in two steps, each of independent interest:
-
(a)
develop general model-theoretic tools to determine sufficient conditions for a definable fibration to split,
-
(b)
give some concrete algebraic characterization of splitting for the logarithmic derivative.
For step (a), we use model-theoretic Galois theory, which makes sense in any -stable theory. Given a fixed definable set , any -internal type comes equipped with a definable group of transformations, called its binding group. In the case of algebraic differential equations, this corresponds to the Galois group associated to Kolchin’s strongly normal extensions [12], themselves a generalization of the Galois groups of Picard-Vessiot extensions. We obtain the following general criteria:
Theorem A.
Let be an algebraically closed differential subfield of a monster model and let be a -internal, weakly -orthogonal type with binding group definably isomorphic to the -points of a nilpotent algebraic group. Let be the generic type of the pullback of under the logarithmic derivative. Then is almost -internal if and only if almost splits.
In [10], Jin and Moosa prove a similar theorem for generic type of an equation of the form with , internal to the constants and with binding group , or . Theorem A is therefore a generalization of their result to arbitrary high dimension, with the caveat that we now need the binding group to be linear nilpotent (thus excluding the case). The result should also hold for linear solvable binding groups, we reserve this question for future work.
To prove Theorem A, we make heavy use of binding groups. Consider a fibration , and assume that is -internal. This gives rise to a definable group morphism from the binding group of to the binding group of . Using the type-definable groupoids of [7], we connect splitting of the fibration to definable splitting of the induced morphism between binding groups, thus obtaining necessary and sufficient conditions for splitting of .
In the case of differential equations, binding groups over the constants are always definably isomorphic to the constant points of algebraic groups. In particular, we obtain Theorem A by using the very constrained structure of nilpotent linear algebraic groups.
Regarding step (b), to be more concrete, Equation 1 splits if, roughly, it can be transformed, using a differentially birational map, into an equation of the form:
for some . In [10] and [5], concrete characterizations of splitting were given if is a field of constants. Let be the generic type of the equation , splitting was characterized in the following cases:
It follows from [5, Corollary 5.5] that the two criteria are in fact identical when . This raises the possibility of a general characterization, free of assumptions on and . We achieve that goal under a mild additional assumption:
Theorem B.
Let be an algebraically closed differential field, a rational function and the generic type of . Let , and suppose that . Then is almost split if and only if there are a non-zero , some and some integer such that:
where is the unique derivation on such that and for all .
This theorem can be considered a simultaneous generalization of the criteria given by [5, Corollary 5.5] and [10, Theorem B], under the extra assumption that .
Together, Theorem A and Theorem B allow us to prove that some pullbacks under the logarithmic derivative are not internal: we first show they must be split using Theorem A, and then use the concrete algebraic characterization of splitting given by Theorem B to prove that splitting is impossible. As an example of application, we prove that if generates a Picard-Vessiot extension, then is never -internal, provided that the binding group of is nilpotent.
Finally, our methods provide examples of internal types with a non-split fibration . More precisely, using non-split semi-abelian varieties, as well as a bit of algebraic group cohomology, we also answer a question of Jin and Moosa in [10] by showing the existence of a -internal type such that the map does not split, with the binding group of isomorphic to an elliptic curve:
Theorem C.
There exists an algebraically closed field , a type such that does not almost-split, and the type is strongly minimal and has a binding group isomorphic to the constant points of an elliptic curve .
We now describe the results of the present work in some details. In the preliminary Section 2, we first give some results on binding groups. Most of these are well-known, but we single out Lemma 2.10 as being of particular interest, and, to the authors’ knowledge, new. Given any two -internal types , it shows that their binding groups have a common definable quotient, which encodes the interactions between the two types over the definable set . Next, we give a definable version of the Jordan decomposition of a linear algebraic group, mimicking some of the results of [6] regarding definable Chevalley and Rosenlicht decompositions. Using this decomposition, we prove that some short exact sequences of definable groups always definably split, which will later yield our automatic splitting result.
Section 3 consists of general model-theoretic results on splitting. We always consider a definable fibration with -internal fibers, for some fixed definable set . In Subsection 3.1, we define (almost) splitting of to be, roughly speaking, interdefinability (or interalgebraicity) between and , for some -internal type over the same parameters. We connect it with uniform internality, as was done [5]. In Subsection 3.2, we introduce the main tool of our work: the connection between splitting of a fibration and splitting of a short exact sequence of definable groups. To do so, we make use of the type-definable groupoid associated to a fibration that was introduced in [7]. In particular, we use the notion of retractability of the groupoid introduced in [3], also reminiscent of descent arguments, see for example [20, Proposition 4.2]. Roughly speaking, it means that we can uniformly and definably pick a family of maps between pairs of fibers of , in a manner compatible with composition. We recall and expand on some results of that article, in particular connecting splitting of and retractability, but under the very restrictive the condition that is fundamental. The main new result in that subsection is Lemma 3.16, which gives a sufficient condition to expand this when is not fundamental: no subgroup of the binding group of should map definably and with non-trivial image to the binding group of a fiber.
We remark that all the techniques developed in Sections 2 and 3 are applicable to any -stable theory. The specific example of the theory of compact complex manifolds comes to mind as another potential area for applications. This would require a good understanding of definable Galois theory in , which has recently been investigated, see [5] and [6] for example.
In Section 4, we focus on the case of the logarithmic derivative. In subsection 4.1, we prove various results around splitting of the logarithmic derivative. We connect it with another notion of splitting, which we call product-splitting, introduced by Jin in his thesis [9] and further studied by Jin and Moosa in [10]. In [9, Conjecture 5.4], Jin conjectures that almost splitting and product splitting are equivalent in the specific case where the image of the fibration is internal. We prove that splitting and product-splitting are equivalent under some various additional assumptions, making progress on Jin’s conjecture. We then prove an algebraic characterization of product-splitting, which leads us to prove Theorem B (Corollary 4.10).
In Subsection 4.2, we consider any fibration , where is internal, and the binding group of nilpotent. In that case, we obtain automatic splitting by Theorem A (Theorem 4.11). As an example of application, we prove that the logarithmic derivative pullback of the generic type of a Picard-Vessiot extension with nilpotent binding group is never almost internal to the constants.
Finally, Subsection 4.3 is dedicated to proving Theorem C (Corollary 4.21): we construct an example of a non-split fibration by the logarithmic derivative, answering a question of Jin and Moosa in [10, bottom of page 5].
Acknowledgements. The authors are thankful to Rahim Moosa for multiple conversations on the subject of this article.
2. Preliminaries
2.1. Internality and binding groups
In this article, we work, unless specified otherwise, in a sufficiently saturated model of some totally transcendental theory , eliminating imaginaries. We will make use, without mentioning it, of the fact that in this context, type-definable groups are always definable (see [21, Corollary 5.19]). Definable will always mean definable over some parameters, and we will always specify the parameters when relevant. When is a type, we denote the set of realizations of in , and similarly for realizations of definable sets. We will often conflate definable sets and types with their set of realizations in . For example, we may say that that a definable group acts on a type to mean that it acts on . For this preliminary subsection, we fix an algebraically closed set of parameters .
We will assume familiarity with geometric stability theory, for which a good reference is [19].
Our results will make heavy use of the binding group of an internal type. Most of the literature uses binding groups over a fixed definable set, but we will sometimes need to work over a family of types. We recall the definition of (almost) internality in that case:
Definition 2.1.
Let be a family of partial types. A type is (almost) internal to if there are some , some , some such that for all the type extends a type in , and:
-
•
(resp. )
-
•
.
In particular, if is any partial type, the type is (almost) internal if and only if it is (almost) internal in the sense of the previous definition.
By a realization of , we mean any tuple realizing some partial type in . We denote the set of realizations of by , or simply if the context is clear enough. Internality of is then equivalent to simply stating that there are parameters such that .
In practice, the minimal assumption for the theory to go through is that the family is -invariant: any automorphism of fixing fixes the set of realizations of pointwise. This happens, for example, if every partial type in is over , which we assume for the rest of this subsection.
A very useful fact is that if is -internal, then the parameters witnessing it can be chosen equal to , where . We call the a fundamental system of solutions of .
Let be the group of automorphisms of fixing pointwise. A consequence of the previous discussion is that, if is -internal, the restriction of any to is entirely determined by , where the form a fundamental system.
In fact, we can consider the group of permutations of coming from the restriction of an element of , i.e. the set of maps:
Fundamental systems of solutions can be used to prove the following classical theorem (see [19, Theorem 7.4.8] for a proof):
Fact 2.2.
Let be a family of partial types over , and a complete type, internal to . Then is isomorphic to an -definable group, and its natural action on is -definable.
We will also denote this definable group . Again, if the family is reduced to a single partial type, this coincides with the usual binding group.
We will also need the notion of (weak) orthogonality:
Definition 2.3.
The type is
-
•
weakly orthogonal to if any realization is independent, over , from any tuple of realizations of ,
-
•
orthogonal to if for any , any realisation is independent, over , from any tuple of realizations of .
It is well-known that if is orthogonal to , then it is not almost -internal (unless it is algebraic). On the other hand, a non-algebraic type can be both -internal and weakly -orthogonal, and we have the well known fact:
Fact 2.4.
A -internal type is weakly -orthogonal if and only if acts transitively on .
For the rest of this subsection, we fix an -definable set . We will consider -internal types.
Note that any non-forking extension of a -internal type is -internal, and we record another well-known fact:
Fact 2.5.
Let be -internal. Then for any , the binding group is a definable subgroup of .
Proof.
Pick a Morley sequence of realizations of . It is also a Morley sequence in , and we can pick large enough so that it is a fundamental system for both and . It is then easy to check that the map given by sending any to the (unique) such that is definable and injective. ∎
Recall that a relatively definable map from a type-definable set to a type-definable set is a relatively definable set that is the graph of a function from to . In the rest of this article, we forgo the adjective relative. If are type-definable over and the map is defined over , then any gives another map with its graph being . The following easy and well-known observation will be useful:
Observation 2.6 (In any theory).
Let be two -type definable sets, and let be a definable map. Then for any , we have that .
Proof.
We have:
and thus for any , we see that . Applying this identity to , we get:
∎
In the rest of this work, we will often forget about which set one needs to restrict to. This will allow us to write which is a slight, but harmless, abuse of notation.
We will often apply this result to definable maps coming from binding groups, and we take the opportunity to point out a subtlety. We constantly view in two ways: as a definable bijection on , and as an element of the -definable group . Let us, for now, denote the first as and the second as . What Fact 2.2 tells us is that there is an -definable group action such that is exactly as a bijection on . If and , then:
In other words, the actions by on , viewed as an -definable group or as a group of definable bijections of , coincide. In the rest of the article, we will identify them and constantly go back-and-forth between the two.
We record the following well-known consequence of Observation 2.6:
Corollary 2.7.
Let be a -internal type. Any -definable subgroup is normal.
Proof.
Let be such a subgroup, and consider and . Then , where is any extension of to . As fixes and is -definable, this implies that . ∎
Given two types such that is -internal, we can also form the binding group . Then:
Lemma 2.8.
The group is -definably isomorphic to an -definable normal subgroup of .
Proof.
The group is -definable, and acts definably on . Viewing and as groups of definable maps acting on , we see that there is an inclusion .
On the definable groups side, fix some Morley sequence in long enough to be a fundamental system for over both and . Then we define a map sending any to the unique element of having the same action on the (and thus on ). This map is definable.
We will make use of the following definable version of Goursat’s lemma (see [15, Chapter I, Exercise 5]), the proof of which is left to the reader:
Lemma 2.9.
Let and be -definable groups, and an -definable subgroup of . Consider the natural projections for , and assume that the maps are surjective. Let and , which we can identify with -definable normal subgroups of and . Then and are -definably isomorphic.
Here is a consequence regarding the interaction between binding groups:
Lemma 2.10.
Let with and being -internal. Then the quotients and are -definably isomorphic.
Proof.
Consider the subgroup of consisting of pairs having a common extension to an automorphism in . Let us first show it is -definable.
We start by showing it is type-definable over extra parameters. If we fix two fundamental systems of solutions and of and , then if and only if , by stable embeddedness of . Still by stable embeddedness, we know that for any tuple , we have . Thus, if we consider the set of -definable functions with domain containing and with image in , we have that if and only if for all , which is indeed an -type definable condition.
By -stability, it must then be an -definable subgroup. We show that it is -definable by showing it is fixed by any . Let . By Observation 2.6, we know that . If we let be a common extension of and , then the automorphism also fixes pointwise, and extends and . Thus .
Let be the projection maps from to and , respectively. By Lemma 2.9, all we need to show is that:
-
(a)
the restriction of the to are surjective,
-
(b)
-
(c)
.
where we made the identification of with subgroups of and .
For (a), we notice that any extends to some , which restricts to some , and thus , which implies that is surjective. We use the same argument for .
Statements (b) and (c) are similar, and we only prove (b). Let Then and have a common extension , showing . For the reverse inclusion, let , this exactly means that . ∎
Consider the condition , or equivalently, that any extends to fixing pointwise. An immediate consequence is that this is symmetrical: any element of also extends to an automorphism of fixing pointwise. A sufficient condition for this to happen is for the two binding groups to not have any isomorphic definable quotients:
Corollary 2.11.
If there is no -definable normal subgroups and such that and are -definably isomorphic, then . In particular, if is weakly orthogonal to , then is weakly orthogonal to .
Proof.
By Lemma 2.10, we obtain that the quotients and are -definably isomorphic. By assumption, this implies that . As is weakly orthogonal to , we know that acts transitively on , and thus so does , which precisely means that is weakly orthogonal to . ∎
Opposed to this condition is isogeny: we say two groups and are (definably) isogenous if there are finite subgroups and such that and are (definably) isomorphic. Note that if and are abelian varieties, this coincides with the usual algebraic geometry definition of isogenous. This will be connected to interalgebraicity of types:
Definition 2.12.
Let , we say that:
-
•
and are interdefinable if for any , there is such that . Note that by compactness, this implies that there is an -definable bijection between and .
-
•
and are interalgebraic if for any , there is such that .
We obtain the following well-known corollary:
Corollary 2.13.
Let be two -internal types, and suppose that they are interalgebraic over . Then and are -definably isogenous.
Proof.
By Lemma 2.10, it is enough to show that the groups and are finite. Pick a fundamental system for . Since and are interalgebraic over , there are realizations of such that for all . Therefore there are only finitely many possibilities for with , for all . Thus there are only finitely many possibilities for the action of on , implying that is finite. Similarly we show that is finite. ∎
If instead the types are interdefinable, the binding groups are isomorphic.
2.2. Definable Jordan decomposition
In this subsection, we take inspiration from the definable Chevalley and Rosenlicht decompositions of [6, Subsection 2.3] and produce a definable Jordan decomposition for groups isomorphic to an algebraic group. For this subsection only, we work with some arbitrary theory eliminating imaginaries (we do not call it to avoid confusion with tori, which we will denote ), some sufficiently saturated , and some small algebraically closed sets of parameters . Let be an -definable purely stably embedded algebraically closed field.
We will extend the Jordan decomposition for linear solvable algebraic groups to groups definably isomorphic to the -points of a linear solvable algebraic group. In particular, we prove the existence of a definable unipotent radical, as well as some basic properties. Note that our methods would in fact allow us to define the definable unipotent radical of any group definably isomorphic to the -points of a linear algebraic group.
Recall that a torus is an algebraic group that is isomorphic to , for some , and a d-group is a subgroup isomorphic to a definable subgroup of for some . If a torus is defined over some algebraically closed field , then this isomorphism is defined over (see [4, 34.3]). We call a group definably isomorphic to the -points of a torus a definable torus, and use a similar terminology for definable unipotent, nilpotent, solvable linear algebraic groups and d-groups.
In this subsection, as well as the rest of this article, we will use, without mentioning it, the fact that for any pure algebraically closed field of characteristic zero, definable groups and algebraic groups coincide, and that any definable map between algebraic groups is a morphism of algebraic groups.
We have the following, which is a consequence of the Lie-Kolchin Theorem (see [4, Chapter 19] for a proof):
Fact 2.14.
Let be a connected solvable linear algebraic group. There is a split short exact sequence of algebraic groups:
In particular , and moreover, it is the unique maximal unipotent subgroup of and the group is a torus.
We prove that there is a definable Jordan decomposition:
Lemma 2.15.
Let be a solvable -definable group, and be a definable isomorphism, where is a connected linear algebraic group. Then there is a -definable normal subgroup which is the unique maximal subgroup of definably isomorphic to the -points of a unipotent group. We thus obtain an exact sequence:
and is definably isomorphic to the points of a torus.
The proof is similar to the proof of [6, Fact 2.8]:
Proof.
By Fact 2.14 we have a definable, definably split, short exact sequence . Let This is a definable subgroup of which is definably isomorphic to the points of the unipotent group .
To show that is unique and maximal, let be another definable subgroup, and let be a definable isomorphism such that , the -points of an unipotent linear algebraic group . Since is purely stably embedded, the map is a morphism of algebraic groups, and thus is an unipotent algebraic subgroup of . By Fact 2.14, it must be contained in , and therefore .
In fact is -definable since any -conjugate of will also be a maximal definable subgroup of which is definably isomorphic to the points of a unipotent algebraic group. By uniqueness of , any such -conjugate is equal to , hence is -definable.
Finally, the algebraic group is a torus, and is definably isomorphic to . ∎
We call this subgroup the definable unipotent radical of . We show that, just as is the case for linear algebraic groups, it is preserved by definable maps:
Lemma 2.16.
Let be a map between solvable -definable groups and let and be definable isomorphisms, where and are connected linear algebraic groups. Then
Proof.
We obtain, by stable embeddedness of , a -definable morphism of linear algebraic groups , and we must have, by [4, Section 15.3] that . Applying , we obtain . By construction of the definable unipotent radical, this gives us . ∎
If the group is nilpotent, we obtain more:
Lemma 2.17.
Let be a nilpotent -definable group, and be a definable isomorphism, where is a connected linear algebraic group. Then there are -definable normal subgroups and which are the unique maximal subgroup of definably isomorphic to the -points of a unipotent group, respectively a torus. Moreover is -definably isomorphic to .
Proof.
Here is a nilpotent connected linear algebraic group, and thus by [4, Proposition 19.2] splits into a product for some unipotent and torus . These groups are the maximal unipotent subgroup and torus of . Using this, we can prove, as was done in Lemma 2.15, that their preimages under are normal -definable subgroups of , and it then follows that is -definably isomorphic to . ∎
In our application, we will obtain a short exact sequence that goes the other way: its kernel is a torus, and its image nilpotent. Having both of these forces splitting as a direct product:
Lemma 2.18.
Let be a -definable connected group, and a definable isomorphism, where is an algebraic group. Suppose that there is a short exact sequence:
where is a definable d-group, and is definably nilpotent linear. Then is nilpotent and is linear. If moreover is a torus and the maximal tori of and are -definably isomorphic to the -points of tori, then the short exact sequence is -definably split and .
Proof.
As an extension of a solvable group by an abelian group, the group is solvable, and is linear as an extension of a linear group by a linear group. By [4, Proposition 19.4], as is a definable d-group, this implies that , and as is normal, is must be central. Therefore is nilpotent because it is a central extension of a nilpotent group.
Now suppose that is a torus. By Lemma 2.17, we know that is -definably isomorphic to the product of its maximal unipotent subgroup and torus, and the same is true for . Consider the following diagram, where all vertical and horizontal short sequences are exact and the vertical sequences are given by Lemma 2.17:
The map has image a subgroup of . Since is a definable torus and is definably unipotent, we can show that , thus is injective. Note that its image must be contained in by Lemma 2.16. Moreover, as is surjective, the Morley rank of (which equals the dimension of the linear algebraic group it is isomorphic to) must be smaller or equal to the Morley rank of of . Thus the map must be bijective.
From the diagram, we see that there is an induced short exact sequence:
By assumption, the maximal tori and are -definably isomorphic to the -points of tori and , which are by stable embeddedness defined over . Using the machinery of characters of -groups (see [4, Section 16.2]), we can prove that the induced map is split. Moreover, because is algebraically closed, by [26, Proposition 3.2.12], it is -definably split. Therefore the induced short exact sequence from to is -definably split, implying that is -definably isomorphic to .
Thus we have obtained that and are -definably isomorphic to are , respectively. We can conclude immediately from this. ∎
We conclude this section with a lemma on decomposition of actions of definably nilpotent linear algebraic groups:
Lemma 2.19.
Let be a definable nilpotent linear algebraic group, where is its unipotent radical and its maximal torus. Consider a definable group action of on some definable set . Then for all , we have .
Proof.
Let and be the projections on and . We first prove:
Claim.
If is a definable subgroup, then .
Proof.
By Lemma 2.9 applied to the group , there is a definable isomorphism between and , where and are normal definable subgroups of , which we can identify with subgroups of and . But the group is definably unipotent and the group is a quotient of a definable torus, so there cannot be any definable isomorphism between them, unless they are both trivial. Therefore and , which is equivalent to . ∎
The action of on is definably isomorphic to its action on , where is a definable subgroup. By the claim, we know that , and thus the action is definably isomorphic to the action of on . In particular for any , we have . ∎
3. Splitting
3.1. Definable fibrations and uniform internality
In this section, we go back to considering a stable theory eliminating imaginaries, and further restrict to being -stable, as to have definable binding groups. We fix an algebraically closed set of parameters , and some -definable set .
If , by an -definable map , what we mean is an -definable map with domain containing , which restricts into a relatively definable map on . The image is then the set of realisations of a type over , which we denote .
The question we will try to answer concerns when a type is interdefinable, or interalgebraic, with a product of types. Of course, given a type , it is always interdefinable with , where is a type with only one realization. To make the question interesting, we ask the type to split along a map, and more precisely a definable fibration:
Definition 3.1.
Let be stationary. An -definable map is a fibration if for any , the type is stationary.
Note that the set of realizations of is exactly the fiber . We introduce a notation for these types:
Notation 3.2.
If and is an -definable map, then for any , we denote .
More specifically, we will mostly be interested in fibrations such that each fiber is almost internal to some fixed definable set . These are called relatively internal fibrations in [5].
Now suppose that the type is almost -internal. When is almost -internal as well? The answer is given by uniform internality, introduced in [5]:
Definition 3.3.
Let be stationary, and a fibration. The fibration is said to be uniformly -internal (resp. uniformly almost -internal) if the fibers are -internal (resp. almost -internal) and there is a tuple such that for some (any) , independent from over , we have that (resp. ).
The following is easy to prove (see [6, Proposition 3.16]):
Fact 3.4.
Let be stationary, and a fibration with almost -internal fibers. The type is almost -internal if and only if is almost -internal and is uniformly almost -internal.
If is interalgebraic with , for some almost -internal , then it is easy to prove that is uniformly almost -internal. In the rest of this article, we will be preoccuppied with the converse: given an uniformly internal map, is it reasonable to expect it to come from a product? This was asked directly in [5] and indirectly in [10]. We start by giving a precise definition of splitting.
Definition 3.5.
Let be an -definable fibration with almost -internal fibers. We say that splits (resp. almost splits) if there is an almost -internal type such that for any , there is such that and are interdefinable (resp. interalgebraic) over , and .
Graphically, splitting means that there is an -definable bijection between and such the the diagram:
commutes. When the fibration is clear from context we may write about (almost) split types. Up to interalgebraicity, almost split and split are the same:
Remark 3.6.
Let be an almost split -definable fibration, and let be the type witnessing it. Then is interalgebraic with . Moreover, the fibration has almost internal fibers, and is split. Thus any almost split type is interalgebraic with a split type.
Using this remark, we can also reduce, up to interalgebraicity, to a split fibration with internality replacing almost internality. Recall the following (see [10, Lemma 3.6] for a proof):
Fact 3.7.
If is an almost -internal type, then there is some such that and is -internal. In particular, any almost -internal type is interalgebraic with a -internal type.
Proposition 3.8.
Let be an almost split -definable fibration, and let be the type witnessing it. Then there is a -internal type such that for any , there is such that and are interalgebraic over .
As mentioned previously, it is straightforward to show that splitting implies uniform internality (see for example, [5, Proposition 3.10]):
Proposition 3.9.
If is an almost split -definable fibration with almost -internal fibers, then it is uniformly almost -internal.
It is natural to ask about the converse, as splitting is, a priori, a much stronger condition. This is one of the themes of [5], which focuses on the case where is orthogonal to . In that case, splitting and uniform internality coincide:
Proposition 3.10 ([5], Proposition 3.14).
Let be a fibration with almost -internal fibers. If is orthogonal to , then is almost split if and only if it is uniformly almost -internal.
Therefore if we are interested in whether or not a uniformly almost -internal splits, we may as well assume that . In this article, we will be interested in the case where is itself -internal.
3.2. Groupoids, splitting and short exact sequences
Recall that definable maps give rise to definable morphisms of binding groups (see for example [10, Lemma 3.1]). More precisely:
Fact 3.11.
If is -internal and is an -definable map, then is also -internal and there is a surjective -definable group homomorphism such that for any and , we have:
Our main tool will be a connection between the splitting of a fibration and the definable splitting of the definable short exact sequence arising from it.
We will use the type-definable groupoid associated to the fibration , introduced by the second author in [7] and [8]. We briefly recall the relevant facts and definitions and direct the reader to the aforementioned works for more details.
Let be an -definable fibration with -internal fibers. For any , we define the set of morphisms to be the set of bijections from to which extend to an automorphism of fixing pointwise. We define the groupoid associated to , denoted , to be the groupoid with set of objects and morphisms the sets . This groupoid acts naturally on in the following way: by construction, any element of is a bijection from to , and we define its (partial) action on to be that bijection. The following is [7, Theorem 1.3]:
Theorem 3.12.
The groupoid is isomorphic to an -type-definable groupoid, and its natural action on is relatively -definable.
We recall the definition of retractability from [3]:
Definition 3.13.
The groupoid is retractable if there is an -definable map such that for all :
-
•
,
-
•
,
Roughly speaking, retractability means that we can pick, uniformly definably, morphisms between the fibers of , in a way that is compatible with composition. Note that it implies, in particular, that is connected: any two of its objects have a morphism between them. By Fact 2.4 this is equivalent to being weakly -orthogonal.
We will use the following:
Lemma 3.14.
Let be an -definable fibration with -internal fibers. The groupoid is retractable if and only if there exists an -definable map such that:
-
•
is weakly orthogonal to ,
-
•
is a definable bijection.
In particular retractability of implies splitting of .
For a proof, going from retractability to splitting is given by[7, Proposition 4.3], and the other direction is given in the proof of [8, Theorem 3.3.5]. We do not include a proof, but give a brief sketch.
Assuming is retractable, we have a family of map . We define an equivalence relation on as if and only if . The compatibility conditions of retractability exactly mean that this is an equivalence relation. The map is the quotient map of . The other direction is similar to the right to left direction of the next lemma.
When is fundamental, the connection between splitting of the short exact sequence and retractability is straightforward. This is proven in [7, Theorem 4.10], but we include a considerably streamlined proof:
Lemma 3.15.
Let be a -internal type and be an -definable fibration. Suppose is weakly -orthogonal and fundamental. Then the short exact sequence
is -definably split if and only if the groupoid is retractable (which implies splits).
Moreover in that case, the binding group is -definably isomorphic to , where is the map of Lemma 3.14.
Proof.
Suppose first that the short exact sequence is -definably split. Then there exists an -definable map such that . For any , because is fundamental, there exists a unique such that . It is easy to see that the family of maps witnesses retractability of .
Conversely, suppose that is retractable, and consider the map given by Lemma 3.14. The maps and induce an -definable group morphism . As is interdefinable with , this map is injective.
As is fundamental, the action of is free and transitive. As is weakly orthogonal to , so is the action of , and therefore we must have , and by Lemma 2.10 we also obtain .
It is easy to see that : given in the former group, we can extend each to , and then . Therefore the map is an -definable isomorphism between and . ∎
Note that even when is not fundamental, if there is an -definable section , by Corollary 2.7, the group is an -definable normal subgroup of , and this group is -definably isomorphic to .
There is one important case where the type is fundamental: if its binding group is abelian. Indeed, the action of is faithful, and all stabilisers are conjugate, hence equal. So if some fixes some , then its fixes every realisation of , and thus must be the identity by faithfullness. We will use this to deal with binding groups isomorphic to tori.
The assumption that is fundamental can always be obtained, without changing the short exact sequence, by taking a high enough Morley power of (see [6, Fact 2.4]). However, applying Lemma 3.15 would yield a splitting of a Morley power of , which is not what we want. Note that the right to left direction of the proof only uses that is fundamental to obtain the equality . We could therefore replace the assumption that is fundamental by this and obtain the same implication.
Outside of the abelian case, we cannot expect the type to be fundamental. In particular, we will need something more to deal with unipotent binding groups. The new results of this section give conditions allowing us to bypass this restriction.
Lemma 3.16.
Let be a -internal type and be an -definable fibration. Suppose is weakly -orthogonal and that the short exact sequence is -definably split via some -definable section . If the following holds:
(): for any and , if , then .
then is retractable (and in particular splits).
Before starting with the proof, it is worth pointing out what assumption () means. If fixes , then fixes , the fiber above , as a set. This assumption says that it in fact fixes pointwise.
Proof.
Let and such that . Then and are in . By assumption , so . Therefore, we can define, for any , the map , for any with . This is well-defined by the previous discussion, and it is straightforward to show that it witnesses retractability of . ∎
Note that if is fundamental, then any fixing some must be the identity. Therefore in that case, we always have (). If is not fundamental, we find the following sufficient condition for ():
Corollary 3.17.
Under the same assumptions as Lemma 3.16, if for some (any) , there is no non-trivial (i.e. with image not equal to the trivial subgroup) definable morphism from any definable subgroup of to then assumption () holds. In particular splits.
Proof.
Fix , let be the definable subgroup of elements fixing . Then gives rise to a definable map , which must have trivial image. This is exactly saying that if , then fixes pointwise. ∎
To go from unipotent and tori to nilpotent, we need to remark that () is preserved under direct product. More precisely:
Lemma 3.18.
Let be a -internal type and be an -definable fibration. Suppose is weakly -orthogonal and that the short exact sequence is -definably split via some -definable section . Suppose that is -definably isomorphic to a direct product , that for all we have , and that the induced actions of and on both satisfy (). Then the action of on satisfies ().
Proof.
We can write any as with . Suppose that for some . Then , therefore . By our () assumptions, we obtain that , and therefore . ∎
Let us say a word about the assumptions of these lemmas before continuing. First, the assumption that the short exact sequence is -definably split cannot be weakened: a counterexample is given in Jin’s thesis [9, Example 5.29]. Here is a quick presentation of it: fix a sufficiently saturated model, and its field of constants. If is a differential transcendental, Jin considers the -internal types generic of and . He shows that the short exact sequence obtained is isomorphic to an extension of by some unipotent group, and that the map does not almost-split. Any such extension definably splits, but to define a splitting of the short exact sequence of binding groups, one needs a fundamental system of solutions for .
As for the necessity of (), we did not find a counterexample, but we expect there is one.
Before we end this section, we include a result about a degenerated case of splitting, which will be of use later:
Lemma 3.19.
Let be a -internal type and be an -definable fibration. Suppose that is weakly -orthogonal and:
-
•
the short exact sequence is -definably split via some -definable section and is finite,
-
•
the action of satisfies ().
Then the groupoid is retractable, thus by Lemma 3.14 we get a map such that is interdefinable with . Moreover .
Proof.
By Lemma 3.16, we know that the groupoid is retractable by the map defined by , for some (any) with .
The map is defined to be the quotient map of the equivalence relation given by if and only if , for any . By the previous paragraph, this is equivalent to .
Let , and . Then there is such that . We then have:
which implies that . As is finite, this means that the orbit of under is finite. Hence . ∎
4. Logarithmic-differential pullbacks in differentially closed fields
4.1. Preliminaries on logarithmic-differential pullbacks
Using lemmas 3.15, 3.16 and 3.18, we can use information about the splitting of definable group extensions to obtain information on the splitting of types. In this section, we are interested in the specific case, already investigated in [10] and [5], of types obtained by pullback via the logarithmic derivative, in a differentially closed field of characteristic zero.
For the rest of this section, we work with differentially closed fields of characteristic zero and assume familiarity with model theory of , for which a reference is [16]. Fix some sufficiently saturated with its derivation . Also fix some algebraically closed differential subfield .
We let be its field of constants. It is a pure algebraically closed field, and is stably embedded. In particular, this implies that definable groups in are exactly algebraic groups, and definable maps between them are exactly morphisms of algebraic groups (see [16, Chapter 7, Section 4] for the equivalence of definable groups and definable morphisms with algebraic groups and morphisms).
We will consider the logarithmic derivative map:
which is a surjective morphism from to with kernel .
Definition 4.1.
Let . We define its pullback under the logarithmic derivative to be , the type of any such that and .
It is a known fact that this defines a unique complete type, and that each fiber is a strongly minimal -internal type (see [9, Proposition 5.3]).
In his thesis [9], Jin conjectures (Conjecture 5.4) the following, for such types and :
Conjecture 4.2.
If is almost -internal, the following are equivalent:
-
(1)
is almost -internal
-
(2)
is almost split
-
(3)
there is an integer such that for some , there are such that:
-
•
-
•
-
•
-
•
It is condition (3) that he and Moosa call splitting in their subsequent article [10]. To avoid confusion with our terminology, we will call this condition product-splitting.
In his thesis, Jin proves the implications (Remark 5.5). In their article [10], Jin and Moosa completely answer this question in the case where is the generic type of an equation , where and is -internal. Their answer depends on the following fact, which itself is a consequence, in , of [16, Chapter 3, Corollary 1.6]:
Fact 4.3.
Let be a definable (in ), faithful, transitive group action on a strongly minimal set . Then is definably isomorphic to either:
-
(1)
, or for some elliptic curve over the constants, acting regularly on itself,
-
(2)
, acting on by affine transformations,
-
(3)
acting on by projective transformations.
Suppose that is the generic type of , for some , and is -internal. They prove:
Theorem A of [10].
If the binding group of is not of dimension , then is almost -internal if and only if it product-splits (which implies that it is almost-split).
When the dimension is , they give an example showing that we can have almost internality without product splitting, thus disproving Jin’s conjecture. More precisely, they show:
Theorem 4.4 of [10].
There exists a differential field and a minimal type internal to the constants, with binding group , such that is -internal but does not product-splits.
They do not show, in this case, that this also implies that the logarithmic derivative map does not almost split (in our terminology). More generally, it was left open whether almost splitting and product-splitting were equivalent. Note that since this question does not involve -internality of , we need not assume to be almost -internal.
We were unable to show that splitting and product-splitting are equivalent in general. The proof of Remark 5.5 in Jin’s thesis [9] can be easily adapted to prove that product-splitting implies almost splitting. We show below that the converse holds under various additional assumptions.
Our motivation for proving the equivalence of these two notions is that splitting is a convenient notion for working with binding groups, but product-splitting has a very concrete and useful algebraic characterization (Lemma 4.9). Therefore this equivalence is key for our applications.
The following lemma will take care of the case where the fibers of are not weakly -orthogonal (and will be useful later):
Lemma 4.4.
Let be a type over any algebraically closed differential subfield of . Let . If for some , the fiber is not weakly -orthogonal, then there is an integer , some constant and some such that:
-
•
-
•
.
In particular product-splits.
Proof.
Let . Non-weak -orthogonality implies that . As , we see that , the field generated by over , so there are and such that . Write and for some (with , and , which we may assume by replacing by ). Taking a derivative, we obtain:
which is a polynomial identity in , with coefficients in . Recall that is chosen so that . Therefore all coefficients of this polynomial are zero. Computing the dominant coefficient gives:
and as and we get:
and thus there is such that:
If , this gives the result by picking and . Otherwise, we see that , for some non-zero . In this case, using euclidian division, we find with and such that . As and , we get . Since and , we also get . We can then repeat the proof using these new polynomials to conclude as before. ∎
We now prove the equivalence of almost splitting and product-splitting, as long as the fibration is not degenerate in a precise sense:
Theorem 4.5.
Let be a type over any algebraically closed differential subfield of . Let . Suppose that . Then the fibration is almost split if and only if it is product-split, i.e. there is an integer such that for some , there are such that:
-
•
-
•
-
•
.
Proof.
As stated previously, that product-splitting implies almost splitting is essentially contained in [9, Remark 5.5].
Suppose that the map is almost split, so there is that is -internal and such that any is interalgebraic with , for some with . We fix such and let , so that .
As the types and are interalgebraic, their binding groups must be isogenous by Corollary 2.13. By Lemma 4.4, we may assume that is weakly orthogonal to and do so for the rest of the proof. Hence the binding group of is definably isomorphic to , and the binding group of must be as well. We want to show that is also definably isomorphic to .
First, remark that is infinite. Suppose, on the contrary, that it is finite. Then , and as is interalgebraic with over , this would imply that , contradicting our assumption.
By Fact 2.5, the binding group is a definable subgroup of . Moreover by Lemma 2.8 the binding group is a normal -definable subgroup of , and, by a proof similar to that of Fact 2.5, is also seen to be a definable subgroup of . Since is infinite, as is connected and strongly minimal, it must be equal to . In particular is normal in .
Note that , and as is algebraically closed, the type is strongly minimal. Moreover, as is weakly -orthogonal, forking calculus along with an automorphism argument yields that is weakly -orthogonal, and in particular acts transitively and faithfully on . As it has a normal definable subgroup isomorphic to , Fact 4.3 implies that is definably isomorphic to .
By [10, Lemma 4.1 (b)], there is such that (and in particular is -internal). Note that , so in particular we also have , as . So and are interalgebraic over . For the rest of the proof we can therefore replace by , and assume that .
Let be the (differential) field generated by over (the two are equal since ). We have that , but also , which implies that is field-algebraic over . Thus there is such that , we take monic and of minimal degree, so . Then we compute:
and thus:
As we picked minimal, we obtain, for all :
and thus, for any :
If there is some , we obtain that . Recall that is, by assumption, weakly orthogonal to . As , so is , which implies , and thus , so . If all are zero, then . In any case, there is such that .
Thus there are such that . Taking derivatives, we obtain
and thus, by equating our two expressions for :
Write and for some (with and ), we compute:
In particular, since , these expressions are given by polynomials in applied to . Therefore the equation (4.1):
is a polynomial in over . Since , all coefficients must be zero. In particular, examining the dominant coefficient, and using that and , we get:
which simplifies into:
and implies that there is such that:
Note that . So setting and gives us what we want. ∎
Over a field of constants, we obtain the equivalence in general :
Corollary 4.6.
Let be a type over any algebraically closed differential field of constants. Let . Then is almost split if and only if there is an integer such that for some , there are such that:
-
•
-
•
-
•
.
Proof.
As before, we only have to prove that almost splitting implies product-splitting. In the previous proof, the only time we used the extra assumption was to show that the binding group of was definably isomorphic to . Keeping the notation of that proof, we now show how to obtain this if is a field of constants.
By Fact 2.5, the binding group is a definable subgroup of . Moreover, we still know that acts transitively and faitfhfully on the strongly minimal type . Over the constants, the possible only binding groups in Fact 4.3 are and an elliptic curve (see the proof of [1, Theorem 3.9]), and therefore the only possible case is being definably isomorphic to . ∎
We also always obtain the equivalence if the base type is orthogonal to the constants:
Corollary 4.7.
Let be a type over any algebraically closed differential field, and assume that is orthogonal to the constants. Let . Then is almost split if and only if there is an integer such that for some , there are such that:
-
•
-
•
-
•
.
Proof.
Again, we keep the notation of the proof of Theorem 4.5. By that theorem, we may assume that , which implies that .
As is orthogonal to the constants, we have that . The former group is, by assumption, finite. Thus is -algebraic. This implies, in particular, that for some (any) , the fiber is not weakly -orthogonal. By Lemma 4.4, we get product-splitting. ∎
The following lemma will be key in applications:
Lemma 4.8.
Let be an algebraically closed field, let be an almost -internal type and be a fibration, with being -internal and weakly -orthogonal. Suppose that for some (any) :
-
•
is -internal and weakly -orthogonal,
-
•
is definably isomorphic to .
Then there exists a type , internal to , and a fibration with . Moreover if almost splits then almost splits.
Proof.
By Fact 3.7, there is a finite-to-one definable map such that is -internal. In particular, for any , it restricts to a finite-to-one map from to . By Lemma 2.10, there is a finite-to-one definable morphism . As the former group is definably isomorphic to , so is the latter.
Fix some . Since is weakly -orthogonal, so is . Moreover, as is weakly orthogonal to , we have that , which is algebraically closed. By [10, Lemma 4.1], there is such that . We let and obtain a fibration with -internal strongly minimal fibers. Finally as , we see that is -internal.
For the moreover part, we assume that almost splits. Then there exists , with a -internal type, such that is interalgebraic, over , with and . We are going to show that witnesses splitting of .
Claim.
We have .
Proof of claim.
We know that , and as , we also have , thus . We obtain:
∎
On the other hand, note that , therefore , thus , which implies that .
By construction , and , so . Note that . Since is definably isomorphic to and is weakly -orthogonal, we see that , which implies that , and combining this with the previous equality we get . As , this implies that and are interalgebraic over , and splits. ∎
Finally, we prove a criteria for product-splitting of generic types of differential equations of arbitrary high order. More specifically, if is some algebraically closed differential subfield of , we will be interested definable sets given by solution sets of equations of the form , with . Note that we can write with without common factors, and that our definable set is then given by solutions of the equation . As the polynomial is irreducible, this definable set has a unique generic type.
Recall that if is any differential field, then the field can be equipped with the derivations with respect to , as well as the derivation , which treats all as constants and equals on . If , then , i.e. the polynomial obtained by differentiating the coefficients of . Using this notation, we can compute, for any tuple and , that . We obtain the following:
Lemma 4.9.
Let be an algebraically closed differential field, a rational function and the generic type of . Let , then is product-split if and only if there are a non-zero , some and some integer such that
Proof.
Suppose that there are such and , consider some and . Also denote . As realizes the generic type of , which is of order , we have that is well-defined and non-zero, as otherwise would satisfy a differential equation of order .
Let and , so that , we need to show that and . The first part is immediate, as and . We also compute:
For the converse, suppose that is product-split. There are , some integer and such that:
-
•
-
•
-
•
.
Note that neither or is zero as . We let and . Again denote . As and , there is a non-zero such that . Applying the logarithmic derivative to the equality , we compute:
Because realizes the generic type of and this equation is of order , this implies:
∎
Under the additional assumption of Theorem 4.5, we get a criteria for almost splitting:
Corollary 4.10.
Let be an algebraically closed differential field, some and the generic type of . Let and assume that . Then is almost split if and only if there are a non-zero , some and some integer such that
4.2. Splitting when the binding group is nilpotent
In this subsection, we prove that for any -internal and weakly -orthogonal type over any algebraically closed field , if has nilpotent binding group, then is almost -internal if and only if splits. This can be seen as a partial generalization of Theorem A of [10].
We will sometimes return to the notation of Subsection 2.2: for example, by a definable nilpotent linear group, we mean a definable group definably isomorphic to the -points of a nilpotent linear algebraic group.
Theorem 4.11.
Let be an algebraically closed differential field and let be a -internal, weakly -orthogonal type with binding group definably isomorphic to the -points of a nilpotent linear algebraic group. Then the following are equivalent:
-
(1)
is almost -internal,
-
(2)
product-splits,
-
(3)
almost splits.
Proof.
The implication , as previously discussed, is essentially given by [9, Remark 5.5], and the implication is immediate. Therefore we only prove . In fact, we will prove and .
Assume that is almost -internal. We will start by showing that this implies almost splitting of . To use binding groups, we need to reduce to the case where is -internal.
First note that if is not weakly -orthogonal, by Lemma 4.4 product-splits. So we now assume that is weakly -orthogonal.
By Lemma 4.8, there is another type , internal to , and a fibration with . This last part implies that the binding group of is also definably isomorphic to the -points of a nilpotent linear algebraic group. Moreover if almost splits, so does . Therefore, replacing by , we may assume that is -internal. Because is weakly -orthogonal and each fiber is weakly -orthogonal, the type is also weakly -orthogonal. By [6, Corollary 2.5], the binding group is connected. We get a short exact sequence:
By [6, Lemma 2.7] is definably isomorphic to , where is Morley sequence in such that is a fundamental system of realizations of , and is a fundamental system of realizations for . As is definably isomorphic to for any , this implies that is definably isomorphic to a subgroup of . We would like to apply Lemma 2.18 to get a definable splitting of this short exact sequence. However need not be connected in general, and thus we cannot get splitting. We will now show how to replace by its connected component by replacing with an interalgebraic type.
The kernel is a definable d-group, and by Lemma 2.18, the binding group is a definable nilpotent linear algebraic group. Moreover, by [4, Theorem 16.2], it is equal to , where is a finite -definable subgroup and is a torus.
The subgroup induces an equivalence relation on by if and only if there is such that . Let be the quotient map. We can consider the quotient type , we have an induced group morphism .
Claim.
.
Proof.
The kernel of is the set of such that for all , there is such that . Note that as is -internal and weakly -orthogonal, it is isolated, and therefore the action of is an action of a definable nilpotent linear algebraic group on a definable set. In particular, for any , by Lemma 2.19, if is in its unipotent radical and is its maximal torus, then implies . As is a subgroup of the maximal torus of and the action on is faithful, this implies that intersects the unipotent radical of trivially. Again by Lemma 2.19 (or rather, its proof), we see that is a subgroup of the maximal torus of . Now, if , there must be some and such that . But the maximal torus is abelian, thus its action on is regular. This implies . ∎
Since , we obtain a definable map by defining . This map induces the short exact sequence:
The group is therefore definably isomorphic to the extension of a nilpotent linear algebraic group by a torus, thus by Lemma 2.18 must be isomorphic to the -points of a nilpotent linear algebraic group. Since is connected, we now can apply the second part of Lemma 2.18.
Note that if we can show that splits, then we automatically obtain almost splitting of , because has finite fibers. This is what we will now prove.
The binding group is definably isomorphic to the extension of a nilpotent linear algebraic group by a torus, thus by Lemma 2.18 must be definably isomorphic to the -point of a nilpotent linear algebraic group.
Moreover, remark that the maximal tori of and are -definably isomorphic to the -points of tori. Indeed, because both groups are definably nilpotent, their maximal tori are central. We can modify the proof of [6, Lemma 2.2] to show that they must then be -definably isomorphic to the -points of tori.
By Lemma 2.18, the short exact sequence of binding groups splits -definably. We know that , where is its maximal torus. By Lemma 2.19, for any , we have . Because is unipotent, by Lemma 3.17, its action on satisfies (). Because is abelian, its action also satisfies (). Therefore by Lemma 3.18 the action of satisfies (). By Lemma 3.16, the map splits, thus almost splits.
We still need to obtain product-splitting. By Theorem 4.5, we may assume that . This also implies that . Using similar ideas as for the proof of Corollary 2.13, we see that this implies that is finite. This group is the kernel of the map . By Lemma 3.19, we obtain that the fibers of are not weakly -orthogonal, and thus neither are the fibers of . But we assumed that they were, a contradiction. ∎
We deduce some internality criteria:
Corollary 4.12.
Let be an algebraically closed differential field, some rational function , and consider , the generic type of . If is -internal, weakly -orthogonal and has a nilpotent binding group, then is almost -internal if and only if there are a non-zero , some and some integer such that
In practice, we can now sometimes reduce the question of internality of a logarithmic differential pullback to some valuation computation. Consider some differential field and the rational function field , recall that it is equipped with the derivations , as well as the derivation that is equal to on , and for all .
We can also equip it with valuations: see as , and equip it with the valuation given, for any , by , and . We have, for all and , that and , from which we deduce that and . We could do this for any , and we denote the valuation obtained.
As an example of application, we look at linear differential equations:
Corollary 4.13.
Let be an algebraically closed differential field, and consider . Let be the generic type of , which is always -internal. If is weakly -orthogonal and has nilpotent binding group, then is not almost -internal.
Proof.
Suppose, for a contradiction, that is almost -internal. By Corollary 4.12, there are a non-zero , some and some integer such that we have the following equality, in :
Let be the valuation in with respect to . The coefficients of must be equal on both sides of the equality. We can write with and . We have:
Now consider the valuation with respect to , the left-hand side has valuation , but the right-hand side has valuation strictly greater than , unless , which is a contradiction, as then , but (note that this also works in the case where and ). ∎
If is a field of constants, then this would apply as long as the generic type of is weakly -orthogonal, as its binding group is always abelian in that case (see the proof of [1, Theorem 3.9] for a justification). Note that this is not always the case: for example, the generic type of is never weakly -orthogonal as long as .
Over non-constant parameters, we could pick some differential transcendental , and let . Then this would apply to the generic type of , which one can easily show is -internal, weakly -orthogonal and with unipotent (but non-abelian) binding group.
4.3. Non-splitting when the binding group is an elliptic curve
We now want to use our methods to exhibit more uniformly internal maps that do not split. In [10, bottom of page 5], Jin and Moosa ask whether there exists a strongly minimal type , with algebraically closed, such that is -internal with binding group an elliptic curve, with the type almost -internal, but the map not almost split. Using Lemma 3.15, we will show that there exist many such types.
To deal with algebraic closure issues, we will need to use the machinery of the functor classifying extensions of commutative algebraic groups. A good reference is [25, Chapter 7], which we will use in the rest of this section. We now recall some of the exposition and facts found in that chapter.
Fix some base algebraically closed field , over which everything will be defined, and a -saturated algebraically closed field (it will be the constants in our application) in which all definable sets and algebraic groups will live. Recall that an extension of algebraic groups is a short exact sequence:
of algebraic groups, where the morphisms are maps of algebraic groups. Two such extensions and are isomorphic if there exists a map making the diagram:
commute. In that case, must be an isomorphism. For any two commutative algebraic groups and , denote the set of isomorphism classes of commutative group extensions. We will now work with commutative groups and switch to additive notation. Here is a summary of the material of [25, VII.1] making into a functor.
Given a morphism , there exists a unique extension and a map making the following diagram:
commute. We denote this extension . If is a morphism, there is a similar construction of . We give details on the construction, as we will need them later. Let:
be an element of . There is a unique and map making the diagram:
commute. Moreover, the group is the subgroup of consisting of pairs such that , and the maps and are the natural projections.
The following summarizes the properties we will need (see [25, Chapter 7, 1.1]):
Fact 4.14.
Let be the category of commutative algebraic groups. The previous construction makes into an additive bifunctor on , contravariant in the first coordinate and covariant in the second. In particular is always an algebraic group, and the maps and are morphisms of algebraic groups.
We will also need to identify the specific algebraic group , where is an elliptic curve. Recall that for any abelian variety , we can form its dual abelian variety , which parametrizes the topologically trivial line bundles on (see [18, II.8]). The dual of an elliptic curve is also an elliptic curve. We then have the following (see [25, Chapter 7, 3.16]):
Fact 4.15.
The group is isomorphic to the dual elliptic curve .
Finally, if is a non-trivial morphism of algebraic groups, we will need information on the kernel of the map . We include the proof for the reader’s (and authors’) comfort, even though it may follow immediately from known homological algebra facts.
Proposition 4.16.
Let be an elliptic curve, and be a non-trivial morphism of algebraic groups. Then has finite kernel.
Proof.
The map has finite kernel, denote it . By [25, Chapter 7, 4.23, Theorem 12], the functor is exact on the category of linear algebraic groups, and we thus obtain a short exact sequence:
It is therefore enough to show that is finite, whenever is a finite algebraic group.
To do so, we show that there is a surjective morphism from to , where is the -torsion of . This torsion group is well-known to be finite (see [18, II.4]), so is finite as well.
Let , we have maps and given by sending to . It is well known (see [18, II.4]) that is surjective and has finite kernel. Pick to be a multiple of the order of , so that is the zero map, or in other words, so that for all . We have a short exact sequence:
which is an element of . Any gives rise to a map , and in particular we obtain . To summarize, we have constructed a map:
By [25, Chapter 7, 1.2, Proposition 2], we obtain an exact sequence:
We show that is surjective by proving that the map is the zero map. Let be an element of . We consider the group consisting of pairs such that and obtain the extension as the top row of the commutative diagram:
We need to show that the top short exact sequence is the trivial extension, or equivalently, that it splits (definably and without needing extra parameters).
We show that for any , there is a unique such that , this will define a section. We know that and are surjective, so there are and such that . Then and , giving existence. For uniqueness, suppose that . This implies that , and as , that . Hence , and by choice of , we obtain that . Therefore .
Thus we have obtained a section to , which is immediately seen to be definable without extra parameters. That it is a morphism is left to the reader.
∎
To complete our proof, we will need to exhibit a -internal type with binding group isomorphic to a semiabelian variety. This follows from Kolchin’s solution to the inverse Galois problem for strongly normal extensions (see [13, Theorem 2] and also [14, Proposition 15.1]), as well as some translation to model-theoretic language. We give some details. Our exposition closely follows Marker’s in [16, Chapter 2, Section 9].
Let be differential subfields of , and denote and . We say that is a strongly normal extension if:
-
(1)
is algebraically closed,
-
(2)
is finitely generated,
-
(3)
for any , we have that .
The differential Galois group is the group of differential automorphisms of fixing . The full differential Galois group is the group of differential automorphisms of fixing .
Since is strongly normal, there is some tuple such that . Let . Marker’s proof of [16, Theorem 9.5] (and the discussion following it) shows that is -internal, and that there is an algebraic group such that (resp. ) is isomorphic to (resp. ).
It is easy to see that is definably isomorphic to the binding group . In other words, the binding group of is isomorphic to the -points of the Galois group of the strongly normal extension . By Kolchin’s solution to the inverse Galois problem, any connected algebraic group is the Galois group of some strongly normal extension, over some differential field , which we can assume to be algebraically closed. We can then take the type obtained previously, which is weakly -orthogonal as . We have obtained:
Fact 4.17.
Fix a connected algebraic group defined over . There exists an algebraically closed differential field and some such that is -internal, weakly -orthogonal, and is definably isomorphic to .
Finally, we will use the Galois correspondence for binding groups. We refer the reader to [24, Theorem 2.3] for a proof, as well as the closest account to what we need that we could find. From that theorem, we deduce the following:
Fact 4.18.
Let be an algebraically closed field, and a -internal, fundamental and weakly -orthogonal type. Fix some . If is an -definable normal subgroup of , then there is such that:
-
•
,
-
•
is -internal, weakly -orthogonal and fundamental, and the binding group is definably isomorphic to ,
-
•
is -internal, weakly -orthogonal and fundamental, and the binding group is definably isomorphic to .
If is connected, we also obtain that is stationary (see [19, Chapter 1, Lemma 6.16]).
From this we obtain:
Fact 4.19.
Let be an algebraically closed field, let be a -internal, fundamental and weakly -orthogonal type. If is a connected -definable normal subgroup of , then there is an -definable fibration with -internal fibers, such that and (for any ) are -internal and fundamental, with definably isomorphic to , and giving rise to the following short exact sequence:
We are now equipped to prove the following:
Theorem 4.20.
There exist an algebraically closed field , some type that is -internal, fundamental and weakly -orthogonal, and a definable fibration , such that:
-
•
is definably isomorphic to , where is a non-definably split extension of an elliptic curve by ,
-
•
does not almost split.
-
•
for any , the type is -internal, weakly orthogonal, fundamental, with binding group definably isomorphic to .
Proof.
By Fact 4.15, there is an algebraic group that is a non-definably split extension of an elliptic curve by , i.e. we have a map with kernel . Moreover, we can assume that does not belong to any finite subgroup of the elliptic curve (equivalently is not a torsion point).
By Fact 4.17, there is an algebraically closed field and such that is -internal and weakly -orthogonal, with definably isomorphic to , where is a non-definably split extension of an elliptic curve by . As was observed in [6, Fact 2.4], we may assume that is fundamental by taking a Morley power.
In particular the group has a definable Chevalley decomposition in the sense of [6, Fact 2.8], meaning a subgroup maximal among definable subgroups -definably isomorphic to the -points of a linear algebraic group. This subgroup is called the linear part of . It is easy to show, in this case, that the linear part of is the kernel of the map induced by on (in particular it is definably isomorphic to ). Moreover, the linear part is -definable and normal by [6, Fact 2.8].
By Fact 4.19, there is an -definable fibration giving rise to the short exact sequence:
and is definably isomorphic to .
Suppose, for a contradiction, that does almost split. Then there exist, by Proposition 3.8, some internal type , some and such that is interalgebraic with over , with .
Since both and are fundamental, by [6, Lemma 2.7] is definably isomorphic to , and hence to . By Lemma 2.10 the binding groups and are isogenous, thus is isogenous to . Also is a definable subgroup of by Fact 2.5.
As acts regularly on , we see that . We deduce from this that . As is -internal, this implies that , and as is algebraically closed, the type is strongly minimal.
Fact 4.3 rules out the cases of the additive group and an elliptic curve for . By Lemma 2.10, the groups and are isogenous. The projection map yields a definable surjective morphism . As the former is isogenous to a semi-abelian variety, it cannot have any definable morphism to nor , ruling out the other two cases. Thus is definably isomorphic to , and by connectedness, we get .
So is definably isomorphic to and is definably isomorphic to . Corollary 2.11 implies that , from which we deduce:
As is weakly orthogonal to , Corollary 2.11 also implies that is weakly orthogonal to . Therefore the short exact sequence:
By Lemma 2.10, the groups and are isogenous, let be the -definable group witnessing it. The kernels of the maps to are given by elements fixing (resp. ), and therefore must in particular be contained in the kernels of the maps to , i.e. and . Therefore the two groups are also isogenous, and we have an -definable group and a commutative diagram of -definable maps:
Both and are abelian, and therefore -definably isomorphic to the -points of algebraic groups (see [6, Lemma 2.2] for a proof of that well-known fact). Denote the -points of these algebraic groups by and . They are defined over , are isogenous, and by the same reasoning as in the previous paragraph, we obtain an -definable algebraic group and a commutative diagram:
where everything is defined over .
By Fact 4.14, the maps and give rise to two automorphisms and of . We obtain that .
We know that is a -definably split extension of , therefore it is trivial as an element of , and so is its image , hence is the identity of , and thus the extension belongs to the kernel of . By Proposition 4.16, the map has finite kernel. Therefore the extension , which is isomorphic to the extension , belongs to a finite subgroup of . This contradicts our choice of , as it was assumed that it did not belong to any finite subgroup of . ∎
To answer Jin and Moosa’s question, we must have a pullback by the logarithmic derivative of a strongly minimal type with binding group an elliptic curve.
Corollary 4.21.
There exists an algebraically closed field , a type such that does not almost-split, the type is strongly minimal and has a binding group isomorphic to the constant points of an elliptic curve .
Proof.
Applying Lemma 4.8 to the algebraically closed field and obtained in Theorem 4.20, we obtain a type , internal to , a fibration with . Because does not almost split, the map does not almost split either. To conclude, we just need to show that is strongly minimal and definably isomorphic to the constant points of an elliptic curve.
Because , we have that , therefore by Lemma 2.10 we obtain an -definable surjective map . As is definably isomorphic to , and thus strongly minimal, its kernel is either the whole group, which would imply that is -definable (i.e. ), or finite.
Note that the type cannot be algebraic, as if it was, then would be almost split and thus so would . Hence in the first possibility, we deduce that is not weakly orthogonal to . But any is in the algebraic closure of some , which contradicts being weakly orthogonal to .
Therefore the kernel of the map is finite, and is definably isomorphic to an algebraic group isogenous to , which thus must be an elliptic curve. Note that by the proof of Lemma 4.8, for any , there is such that . This, and the previous discussion, forces to be strongly minimal. ∎
It is unfortunate that our method does not yield a specific differential equation with -internal set of solutions. Maybe such an equation could be found using the logarithmic derivative of a non-split semi-abelian variety.
References
- [1] James Freitag, Rémi Jaoui, and Rahim Moosa. When any three solutions are independent. Inventiones mathematicae, 230(3):1249–1265, 2022.
- [2] James Freitag and Rahim Moosa. Bounding nonminimality and a conjecture of Borovik–Cherlin. Journal of the European Mathematical Society, 2023.
- [3] John Goodrick and Alexei Kolesnikov. Groupoids, covers, and 3-uniqueness in stable theories. The Journal of Symbolic Logic, 75(3):905–929, 2010.
- [4] James E Humphreys. Linear algebraic groups, volume 21. Springer Science & Business Media, 2012.
- [5] Rémi Jaoui, Léo Jimenez, and Anand Pillay. Relative internality and definable fibrations. Advances in Mathematics, 415:108870, 2023.
- [6] Rémi Jaoui and Rahim Moosa. Abelian reduction in differential-algebraic and bimeromorphic geometry. To appear in Annales de l’Institut Fourier, 2022.
- [7] Léo Jimenez. Groupoids and relative internality. The Journal of Symbolic Logic, 84(3):987–1006, 2019.
- [8] Léo Jimenez. Internality in families. PhD thesis, University of Notre Dame, 2020.
- [9] Ruizhang Jin. The logarithmic derivative and model-theoretic analysability in differentially closed fields. PhD thesis, University of Waterloo, 2019.
- [10] Ruizhang Jin and Rahim Moosa. Internality of logarithmic-differential pullbacks. Transactions of the American Mathematical Society, 373(7):4863–4887, 2020.
- [11] SE Jones and WF Ames. Nonlinear superposition. J. Math. Anal. Appl, 17(3):484–487, 1967.
- [12] Ellis R Kolchin. Galois theory of differential fields. American Journal of Mathematics, 75(4):753–824, 1953.
- [13] Ellis R Kolchin. On the Galois theory of differential fields. American Journal of Mathematics, 77(4):868–894, 1955.
- [14] Jerald Kovacic. Geometric characterization of strongly normal extensions. Transactions of the American Mathematical Society, 358(9):4135–4157, 2006.
- [15] Serge Lang. Algebra, volume 211. Springer Science & Business Media, 2012.
- [16] David Marker, Margit Messmer, and Anand Pillay. Model theory of fields, volume 5. Cambridge University Press, 2017.
- [17] Rahim Moosa. Six lectures on model theory and differential-algebraic geometry. arXiv preprint arXiv:2210.16684, 2022.
- [18] David Mumford, Chidambaran Padmanabhan Ramanujam, and Jurij Ivanovič Manin. Abelian varieties, volume 5. Oxford university press Oxford, 1974.
- [19] Anand Pillay. Geometric stability theory. Number 32. Oxford University Press, 1996.
- [20] Anand Pillay. Remarks on Galois cohomology and definability. The Journal of Symbolic Logic, 62(2):487–492, 1997.
- [21] Bruno Poizat. Stable groups, volume 87. American Mathematical Soc., 2001.
- [22] Abraham Robinson. On the concept of a differentially closed field. Office of Scientific Research, US Air Force, 1959.
- [23] Maxwell Rosenlicht. The nonminimality of the differential closure. Pacific Journal of Mathematics, 52(2):529–537, 1974.
- [24] Omar León Sánchez and Anand Pillay. Some definable Galois theory and examples. Bulletin of Symbolic Logic, 23(2):145–159, 2017.
- [25] Jean-Pierre Serre. Algebraic groups and class fields, volume 117. Springer Science & Business Media, 2012.
- [26] Tonny Albert Springer. Linear algebraic groups. In Algebraic Geometry IV: Linear Algebraic Groups Invariant Theory, pages 1–121. Springer, 1998.