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

Splitting differential equations using Galois theory

Christine Eagles Christine Eagles
University of Waterloo
Department of Pure Mathematics
Mathematics & Computer
Waterloo, ON N2L 3G1
Canada
ceagles@uwaterloo.ca
 and  Léo Jimenez Léo Jimenez
The Ohio State University
Department of Mathematics
Math Tower
Columbus, OH 43210-1174
United States
jimenez.301@osu.edu
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 derivative
2020 Mathematics Subject Classification:
03C45, 03C98, 12H05, 12L12

1. 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 ω\omega-stability of the theory DCF0\mathrm{DCF}_{0} 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 pS(A)p\in S(A) is, roughly, a sequence of AA-definable maps pf1p1f2fnpnp\xrightarrow{f_{1}}p_{1}\xrightarrow{f_{2}}\cdots\xrightarrow{f_{n}}p_{n} such that pnp_{n}, along with each fiber of an fif_{i}, is semi-minimal, meaning internal to some rank one type. That this can be done in DCF0\mathrm{DCF}_{0} 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 DCF0\mathrm{DCF}_{0}: 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 pp, over some algebraically closed differential field kk, is always interdefinable with the generic type of an algebraic vector field (V,s)(V,s). Internality to the constants means that (V,s)(V,s) is, after possibly taking extensions to a differential field k<Lk<L, birationally equivalent to the trivial vector field (𝔸dim(V),0)(\mathbb{A}^{\dim(V)},0). Equivalently, after a base change, there are dim(V)\dim(V) algebraically independent rational functions on VV that are constant for the induced derivation (but are not elements of kk). 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 a1,,ana_{1},\cdots,a_{n} of (E) such that any other solution can be given as a rational function of the aia_{i} 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 f:pf(p)f:p\rightarrow f(p) with internal fibers, i.e. a definable map with all its fibers internal to a fixed definable set 𝒞\mathcal{C}. In the case of DCF0\mathrm{DCF}_{0}, we will always take 𝒞\mathcal{C} to be the field of constants. A starting question would be: suppose that f(p)f(p) is also 𝒞\mathcal{C}-internal, when is pp internal to 𝒞\mathcal{C}? 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 DCF0\mathrm{DCF}_{0}) witnessing internality of all the fibers at once.

As a guiding example, fix some differential field kk, some fk(x)f\in k(x) and consider:

δm(y)=f(δm1(y),,δ(y),y)\delta^{m}(y)=f(\delta^{m-1}(y),\cdots,\delta(y),y)

as well as the following system of differential equations:

{δm(y)=f(δm1(y),,δ(y),y)δ(x)=yx\begin{cases}\delta^{m}(y)=f(\delta^{m-1}(y),\cdots,\delta(y),y)\\ \delta(x)=yx\end{cases}

On non-zero xx, there is a logarithmic derivative map logδ:(x,y)δ(x)x=y\log_{\delta}:(x,y)\rightarrow\frac{\delta(x)}{x}=y which takes a solution of this system to a solution of (1). It is well-known (and easy to prove) that any fiber of logδ\log_{\delta} is internal to the constants, and thus this fits into our general setup. We want to understand (1) by seeing it as fibered over δm(y)=f(δm1(y),,δ(y),y)\delta^{m}(y)=f(\delta^{m-1}(y),\cdots,\delta(y),y). In particular, we want to answer:

Question 1.

If equation (1) is internal to the constants, is equation (1) also internal to the constants?

From the point of view of vector fields, we know that each fiber of logδ\log_{\delta} has a first integral, and equation (1) has mm first integrals by assumption. What the question is asking is whether from these, we can obtain m+1m+1 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 f:pf(p)f:p\rightarrow f(p), we say that ff splits if there is in AA-definable bijection ι\iota between pp and the Morley product f(p)rf(p)\otimes r, for some 𝒞\mathcal{C}-internal type rS(A)r\in S(A), and the diagram

p{p}f(p)r{f(p)\otimes r}f(p){f(p)}ι\scriptstyle{\iota}f\scriptstyle{f}

commutes. If we replace the bijection ι\iota 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 m=1m=1 in [10], but we remove that restriction here. One issue with picking m>1m>1 (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:

  1. (a)

    develop general model-theoretic tools to determine sufficient conditions for a definable fibration to split,

  2. (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 ω\omega-stable theory. Given a fixed definable set 𝒞\mathcal{C}, any 𝒞\mathcal{C}-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 FF be an algebraically closed differential subfield of a monster model 𝒰DCF0\mathcal{U}\models\mathrm{DCF}_{0} and let qS1(F)q\in S_{1}(F) be a 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-orthogonal type with binding group definably isomorphic to the 𝒞\mathcal{C}-points of a nilpotent algebraic group. Let logδ1(q)\log_{\delta}^{-1}(q) be the generic type of the pullback of qq under the logarithmic derivative. Then logδ1(q)\log_{\delta}^{-1}(q) is almost 𝒞\mathcal{C}-internal if and only if logδ:logδ1(q)q\log_{\delta}:\log_{\delta}^{-1}(q)\rightarrow q almost splits.

In [10], Jin and Moosa prove a similar theorem for qq generic type of an equation of the form δ(x)=f(x)\delta(x)=f(x) with fF(X)f\in F(X), internal to the constants and with binding group GaG_{a}, GmG_{m} or GaGmG_{a}\rtimes G_{m}. 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 GaGmG_{a}\rtimes G_{m} 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 f:pf(p)f:p\rightarrow f(p), and assume that pp is 𝒞\mathcal{C}-internal. This gives rise to a definable group morphism from the binding group of pp to the binding group of f(p)f(p). Using the type-definable groupoids of [7], we connect splitting of the fibration ff to definable splitting of the induced morphism between binding groups, thus obtaining necessary and sufficient conditions for splitting of ff.

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:

{δm(y)=f(δm1(y),,δ(y),y)δ(x)=αx\begin{cases}\delta^{m}(y)=f(\delta^{m-1}(y),\cdots,\delta(y),y)\\ \delta(x)=\alpha x\end{cases}

for some αk\alpha\in k. In [10] and [5], concrete characterizations of splitting were given if kk is a field of constants. Let qq be the generic type of the equation δm(y)=f(δm1(y),,δ(y),y)\delta^{m}(y)=f(\delta^{m-1}(y),\cdots,\delta(y),y), splitting was characterized in the following cases:

  • in [10, Theorem B] for m=1m=1 and qq internal to the constants,

  • in [5, Theorem 5.2], for arbitrary high mm, but with qq orthogonal to the constants.

It follows from [5, Corollary 5.5] that the two criteria are in fact identical when m=1m=1. This raises the possibility of a general characterization, free of assumptions on mm and qq. We achieve that goal under a mild additional assumption:

Theorem B.

Let FF be an algebraically closed differential field, a rational function fF(x0,,xm1)f\in F(x_{0},\cdots,x_{m-1}) and qq the generic type of δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)). Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q), and suppose that p(𝒰)acl(q(𝒰),𝒞,F)p(\mathcal{U})\not\subset\operatorname{acl}(q(\mathcal{U}),\mathcal{C},F). Then logδ:pq\log_{\delta}:p\rightarrow q is almost split if and only if there are a non-zero hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}), some eFe\in F and some integer k0k\neq 0 such that:

(kx0e)h=i=0m2hxixi+1+hxm1f+δF(h)(kx_{0}-e)h=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}x_{i+1}+\frac{\partial h}{\partial x_{m-1}}f+\delta^{F}(h)

where δF\delta^{F} is the unique derivation on F(x0,,xm1)F(x_{0},\cdots,x_{m-1}) such that δF|F=δ\delta^{F}|_{F}=\delta and δF(xi)=0\delta^{F}(x_{i})=0 for all ii.

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 p(𝒰)acl(logδ(p)(𝒰),𝒞,F)p(\mathcal{U})\not\subset\operatorname{acl}(\log_{\delta}(p)(\mathcal{U}),\mathcal{C},F).

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 aqa\models q generates a Picard-Vessiot extension, then logδ1(q)\log_{\delta}^{-1}(q) is never 𝒞\mathcal{C}-internal, provided that the binding group of qq is nilpotent.

Finally, our methods provide examples of internal types pp with a non-split fibration f:pf(p)f:p\rightarrow f(p). 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 𝒞\mathcal{C}-internal type pp such that the map logδ:plogδ(p)\log_{\delta}:p\rightarrow\log_{\delta}(p) does not split, with the binding group of logδ(p)\log_{\delta}(p) isomorphic to an elliptic curve:

Theorem C.

There exists an algebraically closed field F<𝒰F<\mathcal{U}, a type pS1(F)p\in S_{1}(F) such that logδ:plogδ(p)\log_{\delta}:p\rightarrow\log_{\delta}(p) does not almost-split, and the type logδ(p)\log_{\delta}(p) is strongly minimal and has a binding group isomorphic to the constant points of an elliptic curve EE.

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 𝒞\mathcal{C}-internal types p,qS(A)p,q\in S(A), it shows that their binding groups have a common definable quotient, which encodes the interactions between the two types over the definable set 𝒞\mathcal{C}. 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 f:pf(p)f:p\rightarrow f(p) with 𝒞\mathcal{C}-internal fibers, for some fixed definable set 𝒞\mathcal{C}. In Subsection 3.1, we define (almost) splitting of ff to be, roughly speaking, interdefinability (or interalgebraicity) between pp and f(p)sf(p)\otimes s, for some 𝒞\mathcal{C}-internal type ss 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 ff, in a manner compatible with composition. We recall and expand on some results of that article, in particular connecting splitting of ff and retractability, but under the very restrictive the condition that f(p)f(p) is fundamental. The main new result in that subsection is Lemma 3.16, which gives a sufficient condition to expand this when f(p)f(p) is not fundamental: no subgroup of the binding group of f(p)f(p) 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 ω\omega-stable theory. The specific example of the theory CCM\mathrm{CCM} of compact complex manifolds comes to mind as another potential area for applications. This would require a good understanding of definable Galois theory in CCM\mathrm{CCM}, 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 logδ:plogδ(p)\log_{\delta}:p\rightarrow\log_{\delta}(p), where pp is internal, and the binding group of logδ(p)\log_{\delta}(p) 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 𝒰\mathcal{U} of some totally transcendental theory TT, 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 pp is a type, we denote p(𝒰)p(\mathcal{U}) the set of realizations of pp in 𝒰\mathcal{U}, and similarly for realizations of definable sets. We will often conflate definable sets and types with their set of realizations in 𝒰\mathcal{U}. For example, we may say that that a definable group GG acts on a type pp to mean that it acts on p(𝒰)p(\mathcal{U}). For this preliminary subsection, we fix an algebraically closed set of parameters AA.

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 𝒬\mathcal{Q} be a family of partial types. A type pS(A)p\in S(A) is (almost) internal to 𝒬\mathcal{Q} if there are some BAB\supset A, some apa\models p, some c1,,cnc_{1},\cdots,c_{n} such that for all ii the type tp(ci/B)\operatorname{tp}(c_{i}/B) extends a type in 𝒬\mathcal{Q}, and:

  • adcl(c1,,cn,B)a\in\operatorname{dcl}(c_{1},\cdots,c_{n},B) (resp. acl\operatorname{acl})

  • a|ABa\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{A}B.

In particular, if XX is any partial type, the type pp is (almost) XX internal if and only if it is (almost) {X}\{X\} internal in the sense of the previous definition.

By a realization of 𝒬\mathcal{Q}, we mean any tuple realizing some partial type in 𝒬\mathcal{Q}. We denote the set of realizations of 𝒬\mathcal{Q} by 𝒬(𝒰)\mathcal{Q}(\mathcal{U}), or simply 𝒬\mathcal{Q} if the context is clear enough. Internality of pp is then equivalent to simply stating that there are parameters BAB\supset A such that p(𝒰)dcl(𝒬(𝒰),B)p(\mathcal{U})\subset\operatorname{dcl}(\mathcal{Q}(\mathcal{U}),B).

In practice, the minimal assumption for the theory to go through is that the family 𝒬\mathcal{Q} is AA-invariant: any automorphism of 𝒰\mathcal{U} fixing AA fixes the set of realizations of 𝒬\mathcal{Q} pointwise. This happens, for example, if every partial type in 𝒬\mathcal{Q} is over AA, which we assume for the rest of this subsection.

A very useful fact is that if pp is 𝒬\mathcal{Q}-internal, then the parameters BB witnessing it can be chosen equal to A{a1,,an}A\cup\{a_{1},\cdots,a_{n}\}, where a1,,anp(n)a_{1},\cdots,a_{n}\models p^{(n)}. We call the aia_{i} a fundamental system of solutions of pp.

Let AutA(𝒰/𝒬)\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{Q}) be the group of automorphisms of 𝒰\mathcal{U} fixing A𝒬A\cup\mathcal{Q} pointwise. A consequence of the previous discussion is that, if pp is 𝒬\mathcal{Q}-internal, the restriction of any σAutA(𝒰/𝒬)\sigma\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{Q}) to p(𝒰)p(\mathcal{U}) is entirely determined by σ(a1),,σ(an)\sigma(a_{1}),\cdots,\sigma(a_{n}), where the aia_{i} form a fundamental system.

In fact, we can consider the group of permutations of p(𝒰)p(\mathcal{U}) coming from the restriction of an element of AutA(𝒰/𝒬)\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{Q}), i.e. the set of maps:

AutA(p/𝒬)={σ:p(𝒰)p(𝒰)| there is σ~AutA(𝒰/𝒬) with σ~|p(𝒰)=σ} .\mathrm{Aut}_{A}(p/\mathcal{Q})=\left\{\sigma:p(\mathcal{U})\rightarrow p(\mathcal{U})|\text{ there is }\widetilde{\sigma}\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{Q})\text{ with }\widetilde{\sigma}|_{p(\mathcal{U})}=\sigma\right\}\text{ .}

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 𝒬\mathcal{Q} be a family of partial types over AA, and pS(A)p\in S(A) a complete type, internal to 𝒬\mathcal{Q}. Then AutA(p/𝒬)\mathrm{Aut}_{A}(p/\mathcal{Q}) is isomorphic to an AA-definable group, and its natural action on p(𝒰)p(\mathcal{U}) is AA-definable.

We will also denote this definable group AutA(p/𝒬)\mathrm{Aut}_{A}(p/\mathcal{Q}). Again, if the family 𝒬\mathcal{Q} 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 pS(A)p\in S(A) is

  • weakly orthogonal to 𝒬\mathcal{Q} if any realization apa\models p is independent, over AA, from any tuple of realizations of 𝒬\mathcal{Q},

  • orthogonal to 𝒬\mathcal{Q} if for any BAB\supset A, any realisation ap|Ba\models p|_{B} is independent, over BB, from any tuple of realizations of 𝒬\mathcal{Q}.

It is well-known that if pS(A)p\in S(A) is orthogonal to 𝒬\mathcal{Q}, then it is not almost 𝒬\mathcal{Q}-internal (unless it is algebraic). On the other hand, a non-algebraic type pp can be both 𝒬\mathcal{Q}-internal and weakly 𝒬\mathcal{Q}-orthogonal, and we have the well known fact:

Fact 2.4.

A 𝒬\mathcal{Q}-internal type pS(A)p\in S(A) is weakly 𝒬\mathcal{Q}-orthogonal if and only if AutA(p/𝒬)\mathrm{Aut}_{A}(p/\mathcal{Q}) acts transitively on p(𝒰)p(\mathcal{U}).

For the rest of this subsection, we fix an AA-definable set 𝒞\mathcal{C}. We will consider 𝒞\mathcal{C}-internal types.

Note that any non-forking extension of a 𝒞\mathcal{C}-internal type is 𝒞\mathcal{C}-internal, and we record another well-known fact:

Fact 2.5.

Let pS(A)p\in S(A) be 𝒞\mathcal{C}-internal. Then for any BAB\supset A, the binding group AutB(p|B/𝒞)\mathrm{Aut}_{B}(p|_{B}/\mathcal{C}) is a definable subgroup of AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}).

Proof.

Pick a Morley sequence a¯=a1,,an\bar{a}=a_{1},\cdots,a_{n} of realizations of p|Bp|_{B}. It is also a Morley sequence in pp, and we can pick nn large enough so that it is a fundamental system for both pp and p|Bp|_{B}. It is then easy to check that the map ι:AutB(p|B/𝒞)AutA(p/𝒞)\iota:\mathrm{Aut}_{B}(p|_{B}/\mathcal{C})\rightarrow\mathrm{Aut}_{A}(p/\mathcal{C}) given by sending any σAutB(p|B/𝒞)\sigma\in\mathrm{Aut}_{B}(p|_{B}/\mathcal{C}) to the (unique) ι(σ)AutA(p/𝒞)\iota(\sigma)\in\mathrm{Aut}_{A}(p/\mathcal{C}) such that ι(σ)(a¯)=σ(a¯)\iota(\sigma)(\bar{a})=\sigma(\bar{a}) is definable and injective. ∎

Recall that a relatively definable map σ\sigma from a type-definable set XX to a type-definable set YY is a relatively definable set Γ(σ)X×Y\Gamma(\sigma)\subset X\times Y that is the graph of a function from XX to YY. In the rest of this article, we forgo the adjective relative. If X,YX,Y are type-definable over AA and the map σ\sigma is defined over AA, then any τAutA(𝒰)\tau\in\mathrm{Aut}_{A}(\mathcal{U}) gives another map τ(σ):XY\tau(\sigma):X\rightarrow Y with its graph being τ(Γ(σ))\tau(\Gamma(\sigma)). The following easy and well-known observation will be useful:

Observation 2.6 (In any theory).

Let X,YX,Y be two AA-type definable sets, and let σ:XY\sigma:X\rightarrow Y be a definable map. Then for any τAutA(𝒰)\tau\in\mathrm{Aut}_{A}(\mathcal{U}), we have that τ(σ)=τ|Yστ1|X\tau(\sigma)=\tau|_{Y}\circ\sigma\circ\tau^{-1}|_{X}.

Proof.

We have:

τ(Γ(σ))\displaystyle\tau(\Gamma(\sigma)) =τ({(x,σ(x)),xX})\displaystyle=\tau\left(\left\{(x,\sigma(x)),x\in X\right\}\right)
={(τ(x),τ(σ(x))),xX}\displaystyle=\left\{(\tau(x),\tau(\sigma(x))),x\in X\right\}

and thus for any xXx\in X, we see that τ(σ)(τ(x))=τ(σ(x))\tau(\sigma)(\tau(x))=\tau(\sigma(x)). Applying this identity to τ1(x)\tau^{-1}(x), we get:

τ(σ)(x)\displaystyle\tau(\sigma)(x) =τ(σ)(τ(τ1(x)))\displaystyle=\tau(\sigma)(\tau(\tau^{-1}(x)))
=τ(σ(τ1(x)))\displaystyle=\tau(\sigma(\tau^{-1}(x)))
=τ|Yστ1|X(x)\displaystyle=\tau|_{Y}\circ\sigma\circ\tau^{-1}|_{X}(x)

In the rest of this work, we will often forget about which set one needs to restrict τ\tau to. This will allow us to write τ(σ)=τστ1=στ1\tau(\sigma)=\tau\circ\sigma\circ\tau^{-1}=\sigma^{\tau^{-1}} 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 σAutA(p/𝒞)\sigma\in\mathrm{Aut}_{A}(p/\mathcal{C}) in two ways: as a definable bijection on pp, and as an element of the AA-definable group AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}). Let us, for now, denote the first as σ|p\sigma|_{p} and the second as σ\sigma. What Fact 2.2 tells us is that there is an AA-definable group action μ:AutA(p/𝒞)×pp\mu:\mathrm{Aut}_{A}(p/\mathcal{C})\times p\rightarrow p such that μ(σ,)\mu(\sigma,\cdot) is exactly σ|p\sigma|_{p} as a bijection on pp. If τAutA(𝒰)\tau\in\mathrm{Aut}_{A}(\mathcal{U}) and a,bpa,b\models p, then:

μ(τ(σ),τ(a))=τ(b)\displaystyle\mu(\tau(\sigma),\tau(a))=\tau(b) μ(σ,a)=b as τAutA(𝒰)\displaystyle\Leftrightarrow\mu(\sigma,a)=b\text{ as }\tau\in\mathrm{Aut}_{A}(\mathcal{U})
σ|p(a)=b\displaystyle\Leftrightarrow\sigma|_{p}(a)=b
τ(σ|p)(τ(a))=τ(b) by the proof of Observation 2.6\displaystyle\Leftrightarrow\tau(\sigma|_{p})(\tau(a))=\tau(b)\text{ by the proof of Observation \ref{obs: aut applied to map}}

In other words, the actions by AutA(𝒰)\mathrm{Aut}_{A}(\mathcal{U}) on AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}), viewed as an AA-definable group or as a group of definable bijections of pp, 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 pS(A)p\in S(A) be a 𝒞\mathcal{C}-internal type. Any AA-definable subgroup H<AutA(p/𝒞)H<\mathrm{Aut}_{A}(p/\mathcal{C}) is normal.

Proof.

Let HH be such a subgroup, and consider τAutA(p/𝒞)\tau\in\mathrm{Aut}_{A}(p/\mathcal{C}) and σH\sigma\in H. Then τστ1=τ~(σ)\tau\circ\sigma\circ\tau^{-1}=\widetilde{\tau}(\sigma), where τ~\widetilde{\tau} is any extension of τ\tau to 𝒰\mathcal{U}. As τ~\widetilde{\tau} fixes AA and HH is AA-definable, this implies that τ~(σ)H\widetilde{\tau}(\sigma)\in H. ∎

Given two types p,qS(A)p,q\in S(A) such that pp is 𝒞\mathcal{C}-internal, we can also form the binding group AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q). Then:

Lemma 2.8.

The group AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q) is AA-definably isomorphic to an AA-definable normal subgroup of AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}).

Proof.

The group AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q) is AA-definable, and acts definably on pp. Viewing AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) and AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q) as groups of definable maps acting on pp, we see that there is an inclusion AutA(p/𝒞,q)<AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C},q)<\mathrm{Aut}_{A}(p/\mathcal{C}).

On the definable groups side, fix some Morley sequence a1,,ana_{1},\cdots,a_{n} in pp long enough to be a fundamental system for pp over both 𝒞\mathcal{C} and {𝒞,q}\{\mathcal{C},q\}. Then we define a map ι\iota sending any σAutA(p/𝒞,q)\sigma\in\mathrm{Aut}_{A}(p/\mathcal{C},q) to the unique element of AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) having the same action on the aia_{i} (and thus on pp). This map is a1,,an,Aa_{1},\cdots,a_{n},A definable.

Consider some τAutA(𝒰)\tau\in\mathrm{Aut}_{A}(\mathcal{U}) and σAutA(p/𝒞,q)\sigma\in\mathrm{Aut}_{A}(p/\mathcal{C},q), it is easy to see, using Observation 2.6, that ι(τ(σ))=τ(ι(σ))\iota(\tau(\sigma))=\tau(\iota(\sigma)). We compute:

τ(ι)(σ)\displaystyle\tau(\iota)(\sigma) =τιτ1(σ)\displaystyle=\tau\circ\iota\circ\tau^{-1}(\sigma)
=τι(τ1(σ))\displaystyle=\tau\circ\iota(\tau^{-1}(\sigma))
=τ(τ1(ι(σ)))\displaystyle=\tau(\tau^{-1}(\iota(\sigma)))
=ι(σ)\displaystyle=\iota(\sigma)

so τ(ι)=ι\tau(\iota)=\iota. As this holds for any τAutA(𝒰)\tau\in\mathrm{Aut}_{A}(\mathcal{U}), we obtain that ι\iota is AA-definable. So AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q) is identified with an AA-definable subgroup of AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}), which is normal by Corollary 2.7. ∎

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 G1G_{1} and G2G_{2} be AA-definable groups, and HH an AA-definable subgroup of G1×G2G_{1}\times G_{2}. Consider the natural projections πi:G1×G2Gi\pi_{i}:G_{1}\times G_{2}\rightarrow G_{i} for i=1,2i=1,2, and assume that the maps πi|H\pi_{i}|_{H} are surjective. Let N1=ker(π2|H)N_{1}=\ker(\pi_{2}|_{H}) and N2=ker(π1|H)N_{2}=\ker(\pi_{1}|_{H}), which we can identify with AA-definable normal subgroups of G1G_{1} and G2G_{2}. Then G1/N1G_{1}/N_{1} and G2/N2G_{2}/N_{2} are AA-definably isomorphic.

Here is a consequence regarding the interaction between binding groups:

Lemma 2.10.

Let p,qS(A)p,q\in S(A) with pp and qq being 𝒞\mathcal{C}-internal. Then the quotients AutA(p/𝒞)/AutA(p/q,𝒞)\mathrm{Aut}_{A}(p/\mathcal{C})/\mathrm{Aut}_{A}(p/q,\mathcal{C}) and AutA(q/𝒞)/AutA(q/p,𝒞)\mathrm{Aut}_{A}(q/\mathcal{C})/\mathrm{Aut}_{A}(q/p,\mathcal{C}) are AA-definably isomorphic.

Proof.

Consider the subgroup HH of AutA(p/𝒞)×AutA(q/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C})\times\mathrm{Aut}_{A}(q/\mathcal{C}) consisting of pairs (σ1,σ2)(\sigma_{1},\sigma_{2}) having a common extension to an automorphism in AutA(𝒰)\mathrm{Aut}_{A}(\mathcal{U}). Let us first show it is AA-definable.

We start by showing it is type-definable over extra parameters. If we fix two fundamental systems of solutions a1,,ana_{1},\cdots,a_{n} and b1,,bmb_{1},\cdots,b_{m} of pp and qq, then (σ1,σ2)H(\sigma_{1},\sigma_{2})\in H if and only if a1anb1bmA𝒞σ1(a1an)σ2(b1bm)a_{1}\cdots a_{n}b_{1}\cdots b_{m}\equiv_{A\mathcal{C}}\sigma_{1}(a_{1}\cdots a_{n})\sigma_{2}(b_{1}\cdots b_{m}), by stable embeddedness of 𝒞\mathcal{C}. Still by stable embeddedness, we know that for any tuple d¯\bar{d}, we have tp(d¯/dcl(𝒞A)dcl(d¯A))tp(d¯/𝒞A)\operatorname{tp}\left(\bar{d}/\operatorname{dcl}(\mathcal{C}A)\cap\operatorname{dcl}(\bar{d}A)\right)\vdash\operatorname{tp}\left(\bar{d}/\mathcal{C}A\right). Thus, if we consider the set 𝔉\mathfrak{F} of AA-definable functions with domain containing tp(a1,,an,b1,,bm/A)\operatorname{tp}(a_{1},\cdots,a_{n},b_{1},\cdots,b_{m}/A) and with image in dcl(𝒞)\operatorname{dcl}(\mathcal{C}), we have that a1anb1bmA𝒞σ1(a1an)σ2(b1bm)a_{1}\cdots a_{n}b_{1}\cdots b_{m}\equiv_{A\mathcal{C}}\sigma_{1}(a_{1}\cdots a_{n})\sigma_{2}(b_{1}\cdots b_{m}) if and only if f(a1anb1bm)=f(σ1(a1an)σ2(b1bm))f(a_{1}\cdots a_{n}b_{1}\cdots b_{m})=f(\sigma_{1}(a_{1}\cdots a_{n})\sigma_{2}(b_{1}\cdots b_{m})) for all f𝔉f\in\mathfrak{F}, which is indeed an (A,a1,,an,b1,,bm)(A,a_{1},\cdots,a_{n},b_{1},\cdots,b_{m})-type definable condition.

By ω\omega-stability, it must then be an a1,,an,b1,,bma_{1},\cdots,a_{n},b_{1},\cdots,b_{m}-definable subgroup. We show that it is AA-definable by showing it is fixed by any τAutA(𝒰)\tau\in\mathrm{Aut}_{A}(\mathcal{U}). Let (σ1,σ2)H(\sigma_{1},\sigma_{2})\in H. By Observation 2.6, we know that τ(σi)=(σi)τ1\tau(\sigma_{i})=\left(\sigma_{i}\right)^{\tau^{-1}}. If we let σ~AutA(𝒰/𝒞)\widetilde{\sigma}\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{C}) be a common extension of σ1\sigma_{1} and σ2\sigma_{2}, then the automorphism (σ~)τ1\left(\widetilde{\sigma}\right)^{\tau^{-1}} also fixes 𝒞A\mathcal{C}\cup A pointwise, and extends (σ1)τ1(\sigma_{1})^{\tau^{-1}} and (σ2)τ1(\sigma_{2})^{\tau^{-1}}. Thus (τ(σ1),τ(σ2))H(\tau(\sigma_{1}),\tau(\sigma_{2}))\in H.

Let π1,π2\pi_{1},\pi_{2} be the projection maps from AutA(p/𝒞)×AutA(q/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C})\times\mathrm{Aut}_{A}(q/\mathcal{C}) to AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) and AutA(q/𝒞)\mathrm{Aut}_{A}(q/\mathcal{C}), respectively. By Lemma 2.9, all we need to show is that:

  1. (a)

    the restriction of the πi\pi_{i} to HH are surjective,

  2. (b)

    ker(π2|H)=AutA(p/𝒞,q)\ker(\pi_{2}|_{H})=\mathrm{Aut}_{A}(p/\mathcal{C},q)

  3. (c)

    ker(π1|H)=AutA(q/𝒞,p)\ker(\pi_{1}|_{H})=\mathrm{Aut}_{A}(q/\mathcal{C},p).

where we made the identification of ker(πi|H)\ker(\pi_{i}|_{H}) with subgroups of AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) and AutA(q/𝒞)\mathrm{Aut}_{A}(q/\mathcal{C}).

For (a), we notice that any σ1AutA(p/𝒞)\sigma_{1}\in\mathrm{Aut}_{A}(p/\mathcal{C}) extends to some σ~AutA(𝒰/𝒞)\widetilde{\sigma}\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{C}), which restricts to some σ2AutA(q/𝒞)\sigma_{2}\in\mathrm{Aut}_{A}(q/\mathcal{C}), and thus (σ1,σ2)H(\sigma_{1},\sigma_{2})\in H, which implies that π1|H\pi_{1}|_{H} is surjective. We use the same argument for π2\pi_{2}.

Statements (b) and (c) are similar, and we only prove (b). Let (σ1,id|q)ker(π2|H).(\sigma_{1},\operatorname{id}|_{q})\in\ker(\pi_{2}|_{H}). Then σ1\sigma_{1} and id|q\operatorname{id}|_{q} have a common extension σ~AutA(𝒰/𝒞)\widetilde{\sigma}\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{C}), showing σ1AutA(p/𝒞,q)\sigma_{1}\in\mathrm{Aut}_{A}(p/\mathcal{C},q). For the reverse inclusion, let σ1AutA(p/𝒞,q)\sigma_{1}\in\mathrm{Aut}_{A}(p/\mathcal{C},q), this exactly means that (σ1,id|q)H(\sigma_{1},\operatorname{id}|_{q})\in H. ∎

Consider the condition AutA(p/𝒞)=AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C})=\mathrm{Aut}_{A}(p/\mathcal{C},q), or equivalently, that any σAutA(p/𝒞)\sigma\in\mathrm{Aut}_{A}(p/\mathcal{C}) extends to σ~AutA(𝒰/𝒞)\widetilde{\sigma}\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{C}) fixing qq pointwise. An immediate consequence is that this is symmetrical: any element of AutA(q/𝒞)\mathrm{Aut}_{A}(q/\mathcal{C}) also extends to an automorphism of 𝒰\mathcal{U} fixing pp 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 AA-definable normal subgroups HAutA(p/𝒞)H\lneq\mathrm{Aut}_{A}(p/\mathcal{C}) and KAutA(q/𝒞)K\lneq\mathrm{Aut}_{A}(q/\mathcal{C}) such that AutA(p/𝒞)/H\mathrm{Aut}_{A}(p/\mathcal{C})/H and AutA(q/𝒞)/K\mathrm{Aut}_{A}(q/\mathcal{C})/K are AA-definably isomorphic, then AutA(p/𝒞)=AutA(p/q,𝒞)\mathrm{Aut}_{A}(p/\mathcal{C})=\mathrm{Aut}_{A}(p/q,\mathcal{C}). In particular, if pp is weakly orthogonal to 𝒞\mathcal{C}, then pp is weakly orthogonal to {𝒞,q}\{\mathcal{C},q\}.

Proof.

By Lemma 2.10, we obtain that the quotients AutA(p/𝒞)/AutA(p/q,𝒞)\mathrm{Aut}_{A}(p/\mathcal{C})/\mathrm{Aut}_{A}(p/q,\mathcal{C}) and AutA(q/𝒞)/AutA(q/p,𝒞)\mathrm{Aut}_{A}(q/\mathcal{C})/\mathrm{Aut}_{A}(q/p,\mathcal{C}) are AA-definably isomorphic. By assumption, this implies that AutA(p/q,𝒞)=AutA(p/𝒞)\mathrm{Aut}_{A}(p/q,\mathcal{C})=\mathrm{Aut}_{A}(p/\mathcal{C}). As pp is weakly orthogonal to 𝒞\mathcal{C}, we know that AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) acts transitively on p(𝒰)p(\mathcal{U}), and thus so does AutA(p/q,𝒞)\mathrm{Aut}_{A}(p/q,\mathcal{C}), which precisely means that pp is weakly orthogonal to {q,𝒞}\{q,\mathcal{C}\}. ∎

Opposed to this condition is isogeny: we say two groups G1G_{1} and G2G_{2} are (definably) isogenous if there are finite subgroups H1<G1H_{1}<G_{1} and H2<G2H_{2}<G_{2} such that G1/H1G_{1}/H_{1} and G2/H2G_{2}/H_{2} are (definably) isomorphic. Note that if G1G_{1} and G2G_{2} 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 p,qS(A)p,q\in S(A), we say that:

  • pp and qq are interdefinable if for any apa\models p, there is bqb\models q such that dcl(aA)=dcl(bA)\operatorname{dcl}(aA)=\operatorname{dcl}(bA). Note that by compactness, this implies that there is an AA-definable bijection between p(𝒰)p(\mathcal{U}) and q(𝒰)q(\mathcal{U}).

  • pp and qq are interalgebraic if for any apa\models p, there is bqb\models q such that acl(aA)=acl(bA)\operatorname{acl}(aA)=\operatorname{acl}(bA).

We obtain the following well-known corollary:

Corollary 2.13.

Let p,qS(A)p,q\in S(A) be two 𝒞\mathcal{C}-internal types, and suppose that they are interalgebraic over AA. Then AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) and AutA(q/𝒞)\mathrm{Aut}_{A}(q/\mathcal{C}) are AA-definably isogenous.

Proof.

By Lemma 2.10, it is enough to show that the groups AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q) and AutA(q/𝒞,p)\mathrm{Aut}_{A}(q/\mathcal{C},p) are finite. Pick a fundamental system a1,,ana_{1},\cdots,a_{n} for pp. Since pp and qq are interalgebraic over AA, there are realizations b1,,bnb_{1},\cdots,b_{n} of qq such that aiacl(bi,A)a_{i}\in\operatorname{acl}(b_{i},A) for all ii. Therefore there are only finitely many possibilities for σ(ai)\sigma(a_{i}) with σAutA(p/𝒞,q)\sigma\in\mathrm{Aut}_{A}(p/\mathcal{C},q), for all ii. Thus there are only finitely many possibilities for the action of σ\sigma on pp, implying that AutA(p/𝒞,q)\mathrm{Aut}_{A}(p/\mathcal{C},q) is finite. Similarly we show that AutA(q/𝒞,p)\mathrm{Aut}_{A}(q/\mathcal{C},p) 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 Th\mathrm{Th} eliminating imaginaries (we do not call it TT to avoid confusion with tori, which we will denote TT), some sufficiently saturated 𝒰Th\mathcal{U}\models\mathrm{Th}, and some small algebraically closed sets of parameters AB𝒰A\subset B\subset\mathcal{U}. Let 𝒞\mathcal{C} be an AA-definable purely stably embedded algebraically closed field.

We will extend the Jordan decomposition for linear solvable algebraic groups to groups definably isomorphic to the 𝒞\mathcal{C}-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 𝒞\mathcal{C}-points of a linear algebraic group.

Recall that a torus is an algebraic group that is isomorphic to GmnG^{n}_{m}, for some nn, and a d-group is a subgroup isomorphic to a definable subgroup of GmnG^{n}_{m} for some nn. If a torus is defined over some algebraically closed field FF, then this isomorphism is defined over FF (see [4, 34.3]). We call a group definably isomorphic to the 𝒞\mathcal{C}-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 𝒞\mathcal{C} 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 GG be a connected solvable linear algebraic group. There is a split short exact sequence of algebraic groups:

1GuGT11\rightarrow G_{u}\rightarrow G\rightarrow T\rightarrow 1

In particular GuGG_{u}\trianglelefteq G, and moreover, it is the unique maximal unipotent subgroup of GG and the group TT is a torus.

We prove that there is a definable Jordan decomposition:

Lemma 2.15.

Let GG be a solvable BB-definable group, and f:GG^(𝒞)f:G\rightarrow\widehat{G}(\mathcal{C}) be a definable isomorphism, where G^\widehat{G} is a connected linear algebraic group. Then there is a BB-definable normal subgroup GuGG_{u}\trianglelefteq G which is the unique maximal subgroup of GG definably isomorphic to the 𝒞\mathcal{C}-points of a unipotent group. We thus obtain an exact sequence:

1GuGG/Gu11\rightarrow G_{u}\rightarrow G\rightarrow G/G_{u}\rightarrow 1

and T:=G/GuT:=G/G_{u} is definably isomorphic to the 𝒞\mathcal{C} 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 1G^uG^T^11\rightarrow\widehat{G}_{u}\rightarrow\widehat{G}\rightarrow\widehat{T}\rightarrow 1. Let Gu=f1(G^u(𝒞)).G_{u}=f^{-1}(\widehat{G}_{u}(\mathcal{C})). This is a definable subgroup of GG which is definably isomorphic to G^u(𝒞),\widehat{G}_{u}(\mathcal{C}), the 𝒞\mathcal{C} points of the unipotent group G^u\widehat{G}_{u}.

To show that GuG_{u} is unique and maximal, let HGH\leq G be another definable subgroup, and let ll be a definable isomorphism such that l(H)=H^(𝒞)l(H)=\widehat{H}(\mathcal{C}), the 𝒞\mathcal{C}-points of an unipotent linear algebraic group H^\widehat{H}. Since 𝒞\mathcal{C} is purely stably embedded, the map fl1f\circ l^{-1} is a morphism of algebraic groups, and thus f(H)f(H) is an unipotent algebraic subgroup of G^(𝒞)\widehat{G}(\mathcal{C}). By Fact 2.14, it must be contained in G^u(𝒞)\widehat{G}_{u}(\mathcal{C}), and therefore HGuH\leq G_{u}.

In fact GuG_{u} is BB-definable since any BB-conjugate of GuG_{u} will also be a maximal definable subgroup of GG which is definably isomorphic to the 𝒞\mathcal{C} points of a unipotent algebraic group. By uniqueness of GuG_{u}, any such BB-conjugate is equal to GuG_{u}, hence GuG_{u} is BB-definable.

Finally, the algebraic group T^\widehat{T} is a torus, and TT is definably isomorphic to T^(𝒞)\widehat{T}(\mathcal{C}). ∎

We call this subgroup the definable unipotent radical of GG. We show that, just as is the case for linear algebraic groups, it is preserved by definable maps:

Lemma 2.16.

Let π:GK\pi:G\rightarrow K be a map between solvable BB-definable groups and let f:GG^(𝒞)f:G\rightarrow\widehat{G}(\mathcal{C}) and h:KK^(𝒞)h:K\rightarrow\widehat{K}(\mathcal{C}) be definable isomorphisms, where G^\widehat{G} and K^\widehat{K} are connected linear algebraic groups. Then π(Gu)<Ku\pi(G_{u})<K_{u}

Proof.

We obtain, by stable embeddedness of 𝒞\mathcal{C}, a 𝒞B\mathcal{C}\cap B-definable morphism of linear algebraic groups π^=hπf1:G^(𝒞)K^(𝒞)\widehat{\pi}=h\circ\pi\circ f^{-1}:\widehat{G}(\mathcal{C})\rightarrow\widehat{K}(\mathcal{C}), and we must have, by [4, Section 15.3] that π^(G^u)<K^u\widehat{\pi}(\widehat{G}_{u})<\widehat{K}_{u}. Applying h1h^{-1}, we obtain πf1(G^u)=h1(K^u)\pi\circ f^{-1}(\widehat{G}_{u})=h^{-1}(\widehat{K}_{u}). By construction of the definable unipotent radical, this gives us π(Gu)<Ku\pi(G_{u})<K_{u}. ∎

If the group is nilpotent, we obtain more:

Lemma 2.17.

Let GG be a nilpotent BB-definable group, and f:GG^(𝒞)f:G\rightarrow\widehat{G}(\mathcal{C}) be a definable isomorphism, where G^\widehat{G} is a connected linear algebraic group. Then there are BB-definable normal subgroups GuGG_{u}\trianglelefteq G and TGT\trianglelefteq G which are the unique maximal subgroup of GG definably isomorphic to the 𝒞\mathcal{C}-points of a unipotent group, respectively a torus. Moreover GG is BB-definably isomorphic to Gu×TG_{u}\times T.

Proof.

Here G^\widehat{G} is a nilpotent connected linear algebraic group, and thus by [4, Proposition 19.2] splits into a product Gu^×T^\widehat{G_{u}}\times\widehat{T} for some unipotent Gu^\widehat{G_{u}} and torus T^\widehat{T}. These groups are the maximal unipotent subgroup and torus of G^\widehat{G}. Using this, we can prove, as was done in Lemma 2.15, that their preimages under ff are normal BB-definable subgroups of GG, and it then follows that GG is BB-definably isomorphic to Gu×TG_{u}\times T. ∎

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 GG be a BB-definable connected group, and f:GG^(𝒞)f:G\rightarrow\widehat{G}(\mathcal{C}) a definable isomorphism, where G^\widehat{G} is an algebraic group. Suppose that there is a short exact sequence:

1KG𝜋G/K11\rightarrow K\rightarrow G\xrightarrow{\pi}G/K\rightarrow 1

where KK is a definable d-group, and G/KG/K is definably nilpotent linear. Then GG is nilpotent and G^\widehat{G} is linear. If moreover KK is a torus and the maximal tori of GG and G/KG/K are BB-definably isomorphic to the 𝒞\mathcal{C}-points of tori, then the short exact sequence is BB-definably split and G=K×G/KG=K\times G/K.

Proof.

As an extension of a solvable group by an abelian group, the group GG is solvable, and G^\widehat{G} is linear as an extension of a linear group by a linear group. By [4, Proposition 19.4], as KK is a definable d-group, this implies that CG(K)=NG(K)C_{G}(K)=N_{G}(K), and as KK is normal, is must be central. Therefore GG is nilpotent because it is a central extension of a nilpotent group.

Now suppose that KK is a torus. By Lemma 2.17, we know that GG is BB-definably isomorphic to the product of its maximal unipotent subgroup and torus, and the same is true for G/KG/K. Consider the following diagram, where all vertical and horizontal short sequences are exact and the vertical sequences are given by Lemma 2.17:

1{1}1{1}Gu{G_{u}}(G/K)u{(G/K)_{u}}1{1}K{K}G{G}G/K{G/K}1{1}T{T}S{S}1{1}1{1}ι\scriptstyle{\iota}π\scriptstyle{\pi}

The map πι\pi\circ\iota has image a subgroup of (G/K)u(G/K)_{u}. Since KK is a definable torus and GuG_{u} is definably unipotent, we can show that KGu={id}K\cap G_{u}=\{\operatorname{id}\}, thus πι\pi\circ\iota is injective. Note that its image must be contained in (G/K)u(G/K)_{u} by Lemma 2.16. Moreover, as π\pi is surjective, the Morley rank of (G/K)u(G/K)_{u} (which equals the dimension of the linear algebraic group it is isomorphic to) must be smaller or equal to the Morley rank of of GuG_{u}. Thus the map πι\pi\circ\iota must be bijective.

From the diagram, we see that there is an induced short exact sequence:

1KTS11\rightarrow K\rightarrow T\rightarrow S\rightarrow 1

By assumption, the maximal tori TT and SS are BB-definably isomorphic to the 𝒞\mathcal{C}-points of tori T^\widehat{T} and S^\widehat{S}, which are by stable embeddedness defined over 𝒞B\mathcal{C}\cap B. Using the machinery of characters of dd-groups (see [4, Section 16.2]), we can prove that the induced map T^S^\widehat{T}\rightarrow\widehat{S} is split. Moreover, because B𝒞B\cap\mathcal{C} is algebraically closed, by [26, Proposition 3.2.12], it is B𝒞B\cap\mathcal{C}-definably split. Therefore the induced short exact sequence from TT to SS is BB-definably split, implying that TT is BB-definably isomorphic to K×SK\times S.

Thus we have obtained that GG and G/KG/K are BB-definably isomorphic to Gu×S×KG_{u}\times S\times K are Gu×SG_{u}\times S, 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 G=Gu×TG=G_{u}\times T be a definable nilpotent linear algebraic group, where GuG_{u} is its unipotent radical and TT its maximal torus. Consider a definable group action of GG on some definable set XX. Then for all aXa\in X, we have GuaTa={a}G_{u}a\cap Ta=\{a\}.

Proof.

Let π1\pi_{1} and π2\pi_{2} be the projections on GuG_{u} and TT. We first prove:

Claim.

If H<Gu×TH<G_{u}\times T is a definable subgroup, then H=π1(H)×π2(H)H=\pi_{1}(H)\times\pi_{2}(H).

Proof.

By Lemma 2.9 applied to the group π1(H)×π2(H)\pi_{1}(H)\times\pi_{2}(H), there is a definable isomorphism between π1(H)/N1\pi_{1}(H)/N_{1} and π2(H)/N2\pi_{2}(H)/N_{2}, where N1=ker(π2|H)N_{1}=\ker(\pi_{2}|_{H}) and N2=ker(π1|H)N_{2}=\ker(\pi_{1}|_{H}) are normal definable subgroups of HH, which we can identify with subgroups of π1(H)\pi_{1}(H) and π2(H)\pi_{2}(H). But the group π1(H)/N1\pi_{1}(H)/N_{1} is definably unipotent and the group π2(H)/N2\pi_{2}(H)/N_{2} is a quotient of a definable torus, so there cannot be any definable isomorphism between them, unless they are both trivial. Therefore π1(H)=N1\pi_{1}(H)=N_{1} and π2(H)=N2\pi_{2}(H)=N_{2}, which is equivalent to H=π1(H)×π2(H)H=\pi_{1}(H)\times\pi_{2}(H). ∎

The action of GG on pp is definably isomorphic to its action on G/HG/H, where H<GH<G is a definable subgroup. By the claim, we know that H=π1(H)×π2(H)H=\pi_{1}(H)\times\pi_{2}(H), and thus the action is definably isomorphic to the action of G=Gu×TG=G_{u}\times T on Gu/π1(H)×T/π2(H)G_{u}/\pi_{1}(H)\times T/\pi_{2}(H). In particular for any apa\models p, we have GuaTa={a}G_{u}a\cap Ta=\{a\}. ∎

3. Splitting

3.1. Definable fibrations and uniform internality

In this section, we go back to considering a stable theory TT eliminating imaginaries, and further restrict to TT being ω\omega-stable, as to have definable binding groups. We fix an algebraically closed set of parameters AA, and some AA-definable set 𝒞\mathcal{C}.

If pS(A)p\in S(A), by an AA-definable map f:pf(p)f:p\rightarrow f(p), what we mean is an AA-definable map ff with domain containing pp, which restricts into a relatively definable map on p(𝒰)p(\mathcal{U}). The image f(p(𝒰))f(p(\mathcal{U})) is then the set of realisations of a type over AA, which we denote f(p)f(p).

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 pp, it is always interdefinable with pqp\otimes q, where qq 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 pS(A)p\in S(A) be stationary. An AA-definable map f:pf(p)f:p\rightarrow f(p) is a fibration if for any apa\models p, the type tp(a/f(a)A)\operatorname{tp}(a/f(a)A) is stationary.

Note that the set of realizations of tp(a/f(a)A)\operatorname{tp}(a/f(a)A) is exactly the fiber f1({f(a)})p(𝒰)f^{-1}(\{f(a)\})\cap p(\mathcal{U}). We introduce a notation for these types:

Notation 3.2.

If pS(A)p\in S(A) and f:pf(p)f:p\rightarrow f(p) is an AA-definable map, then for any f(a)f(p)f(a)\models f(p), we denote pf(a):=tp(a/f(a)A)p_{f(a)}:=\operatorname{tp}(a/f(a)A).

More specifically, we will mostly be interested in fibrations such that each fiber pf(a)p_{f(a)} is almost internal to some fixed definable set 𝒞\mathcal{C}. These are called relatively internal fibrations in [5].

Now suppose that the type f(p)f(p) is almost 𝒞\mathcal{C}-internal. When is pp almost 𝒞\mathcal{C}-internal as well? The answer is given by uniform internality, introduced in [5]:

Definition 3.3.

Let pS(A)p\in S(A) be stationary, and f:pf(p)f:p\rightarrow f(p) a fibration. The fibration f:pf(p)f:p\rightarrow f(p) is said to be uniformly 𝒞\mathcal{C}-internal (resp. uniformly almost 𝒞\mathcal{C}-internal) if the fibers pf(a)p_{f(a)} are 𝒞\mathcal{C}-internal (resp. almost 𝒞\mathcal{C}-internal) and there is a tuple ee such that for some (any) apa\models p, independent from ee over AA, we have that adcl(e,f(a),𝒞)a\in\operatorname{dcl}(e,f(a),\mathcal{C}) (resp. acl\operatorname{acl}).

The following is easy to prove (see [6, Proposition 3.16]):

Fact 3.4.

Let pS(A)p\in S(A) be stationary, and f:pf(p)f:p\rightarrow f(p) a fibration with almost 𝒞\mathcal{C}-internal fibers. The type pp is almost 𝒞\mathcal{C}-internal if and only if f(p)f(p) is almost 𝒞\mathcal{C}-internal and ff is uniformly almost 𝒞\mathcal{C}-internal.

If pp is interalgebraic with f(p)sf(p)\otimes s, for some almost 𝒞\mathcal{C}-internal sS(A)s\in S(A), then it is easy to prove that ff is uniformly almost 𝒞\mathcal{C}-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 f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration with almost 𝒞\mathcal{C}-internal fibers. We say that ff splits (resp. almost splits) if there is an almost 𝒞\mathcal{C}-internal type rS(A)r\in S(A) such that for any apa\models p, there is brb\models r such that (f(a),b)(f(a),b) and aa are interdefinable (resp. interalgebraic) over AA, and f(a)|Abf(a)\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{A}b.

Graphically, splitting means that there is an AA-definable bijection between pp and f(p)rf(p)\otimes r such the the diagram:

p{p}f(p)r{f(p)\otimes r}f(p){f(p)}f\scriptstyle{f}

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 f:pf(p)f:p\rightarrow f(p) be an almost split AA-definable fibration, and let rr be the type witnessing it. Then pp is interalgebraic with f(p)rf(p)\otimes r. Moreover, the fibration π:f(p)rf(p)\pi:f(p)\otimes r\rightarrow f(p) 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 p=tp(a/A)p=\operatorname{tp}(a/A) is an almost 𝒞\mathcal{C}-internal type, then there is some bdcl(aA)b\in\operatorname{dcl}(aA) such that aacl(bA)a\in\operatorname{acl}(bA) and tp(b/A)\operatorname{tp}(b/A) is 𝒞\mathcal{C}-internal. In particular, any almost 𝒞\mathcal{C}-internal type is interalgebraic with a 𝒞\mathcal{C}-internal type.

Using Remark 3.6 and applying Fact 3.7 to qq, we get:

Proposition 3.8.

Let f:pf(p)f:p\rightarrow f(p) be an almost split AA-definable fibration, and let rr be the type witnessing it. Then there is a 𝒞\mathcal{C}-internal type r~S(A)\widetilde{r}\in S(A) such that for any apa\models p, there is br~b\models\widetilde{r} such that aa and (b,f(a))(b,f(a)) are interalgebraic over AA.

As mentioned previously, it is straightforward to show that splitting implies uniform internality (see for example, [5, Proposition 3.10]):

Proposition 3.9.

If f:pf(p)f:p\rightarrow f(p) is an almost split AA-definable fibration with almost 𝒞\mathcal{C}-internal fibers, then it is uniformly almost 𝒞\mathcal{C}-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 f(p)f(p) is orthogonal to 𝒞\mathcal{C}. In that case, splitting and uniform internality coincide:

Proposition 3.10 ([5], Proposition 3.14).

Let f:pf(p)f:p\rightarrow f(p) be a fibration with almost 𝒞\mathcal{C}-internal fibers. If f(p)f(p) is orthogonal to 𝒞\mathcal{C}, then ff is almost split if and only if it is uniformly almost 𝒞\mathcal{C}-internal.

Therefore if we are interested in whether or not a uniformly almost 𝒞\mathcal{C}-internal f:pf(p)f:p\rightarrow f(p) splits, we may as well assume that f(p)⟂̸𝒞f(p)\not\perp\mathcal{C}. In this article, we will be interested in the case where f(p)f(p) is itself 𝒞\mathcal{C}-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 pS(A)p\in S(A) is 𝒞\mathcal{C}-internal and f:pf(p)f:p\rightarrow f(p) is an AA-definable map, then f(p)f(p) is also 𝒞\mathcal{C}-internal and there is a surjective AA-definable group homomorphism f~:AutA(p/𝒞)AutA(f(p)/𝒞)\widetilde{f}:\mathrm{Aut}_{A}(p/\mathcal{C})\rightarrow\mathrm{Aut}_{A}(f(p)/\mathcal{C}) such that for any σAutA(p/𝒞)\sigma\in\mathrm{Aut}_{A}(p/\mathcal{C}) and apa\models p, we have:

f~(σ)(f(a))=f(σ(a)) .\widetilde{f}(\sigma)(f(a))=f(\sigma(a))\text{ .}

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 ff, 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 f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration with 𝒞\mathcal{C}-internal fibers. For any f(a),f(b)f(p)f(a),f(b)\models f(p), we define the set of morphisms Mor(f(a),f(b))\mathrm{Mor}(f(a),f(b)) to be the set of bijections from pf(a)p_{f(a)} to pf(b)p_{f(b)} which extend to an automorphism of 𝒰\mathcal{U} fixing A𝒞A\cup\mathcal{C} pointwise. We define the groupoid associated to ff, denoted GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}), to be the groupoid with set of objects f(p)f(p) and morphisms the sets Mor(f(a),f(b))\mathrm{Mor}(f(a),f(b)). This groupoid acts naturally on pp in the following way: by construction, any element of Mor(f(a),f(b))\mathrm{Mor}(f(a),f(b)) is a bijection from pf(a)p_{f(a)} to pf(b)p_{f(b)}, and we define its (partial) action on pp to be that bijection. The following is [7, Theorem 1.3]:

Theorem 3.12.

The groupoid GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is isomorphic to an AA-type-definable groupoid, and its natural action on pp is relatively AA-definable.

We recall the definition of retractability from [3]:

Definition 3.13.

The groupoid GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable if there is an AA-definable map r:f(p)2GdA(p,f/𝒞)r:f(p)^{2}\rightarrow\mathrm{Gd}_{A}(p,f/\mathcal{C}) such that for all a,b,cf(p)a,b,c\models f(p):

  • r(f(a),f(b))Mor(f(a),f(b))r(f(a),f(b))\in\mathrm{Mor}(f(a),f(b)),

  • r(f(b),f(c))r(f(a),f(b))=r(f(a),f(c))r(f(b),f(c))\circ r(f(a),f(b))=r(f(a),f(c)),

Roughly speaking, retractability means that we can pick, uniformly definably, morphisms between the fibers of ff, in a way that is compatible with composition. Note that it implies, in particular, that GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is connected: any two of its objects have a morphism between them. By Fact 2.4 this is equivalent to f(p)f(p) being weakly 𝒞\mathcal{C}-orthogonal.

We will use the following:

Lemma 3.14.

Let f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration with 𝒞\mathcal{C}-internal fibers. The groupoid GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable if and only if there exists an AA-definable map π:pπ(p)\pi:p\rightarrow\pi(p) such that:

  • f(p)f(p) is weakly orthogonal to {𝒞,π(p)}\{\mathcal{C},\pi(p)\},

  • π×f:pπ(p)×f(p)\pi\times f:p\rightarrow\pi(p)\times f(p) is a definable bijection.

In particular retractability of GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) implies splitting of ff.

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 GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable, we have a family of map r(f(a),f(b))r(f(a),f(b)). We define an equivalence relation EE on pp as E(x,y)E(x,y) if and only if r(f(x),f(y))(x)=yr(f(x),f(y))(x)=y. The compatibility conditions of retractability exactly mean that this is an equivalence relation. The map π\pi is the quotient map of EE. The other direction is similar to the right to left direction of the next lemma.

When f(p)f(p) 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 pS(A)p\in S(A) be a 𝒞\mathcal{C}-internal type and f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration. Suppose f(p)f(p) is weakly 𝒞\mathcal{C}-orthogonal and fundamental. Then the short exact sequence

1KAutA(p/𝒞)f~AutA(f(p)/𝒞)11\rightarrow K\rightarrow\mathrm{Aut}_{A}(p/\mathcal{C})\xrightarrow{\widetilde{f}}\mathrm{Aut}_{A}(f(p)/\mathcal{C})\rightarrow 1

is AA-definably split if and only if the groupoid GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable (which implies ff splits).

Moreover in that case, the binding group AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) is AA-definably isomorphic to AutA(f(p)/𝒞)×AutA(π(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C})\times\mathrm{Aut}_{A}(\pi(p)/\mathcal{C}), where π\pi is the map of Lemma 3.14.

Proof.

Suppose first that the short exact sequence is AA-definably split. Then there exists an AA-definable map s:AutA(f(p)/𝒞)AutA(p/𝒞)s:\mathrm{Aut}_{A}(f(p)/\mathcal{C})\rightarrow\mathrm{Aut}_{A}(p/\mathcal{C}) such that f~s=id\widetilde{f}\circ s=\operatorname{id}. For any f(a),f(b)f(p)f(a),f(b)\models f(p), because f(p)f(p) is fundamental, there exists a unique σf(a),f(b)AutA(f(p)/𝒞)\sigma_{f(a),f(b)}\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) such that σf(a),f(b)(f(a))=f(b)\sigma_{f(a),f(b)}(f(a))=f(b). It is easy to see that the family of maps {s(σf(a),f(b))|pf(a):f(a),f(b)f(p)}\{s(\sigma_{f(a),f(b)})|_{p_{f(a)}}:f(a),f(b)\models f(p)\} witnesses retractability of GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}).

Conversely, suppose that GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable, and consider the map π\pi given by Lemma 3.14. The maps ff and π\pi induce an AA-definable group morphism π~×f~:AutA(p/𝒞)AutA(π(p)/𝒞)×AutA(f(p)/𝒞)\widetilde{\pi}\times\widetilde{f}:\mathrm{Aut}_{A}(p/\mathcal{C})\rightarrow\mathrm{Aut}_{A}(\pi(p)/\mathcal{C})\times\mathrm{Aut}_{A}(f(p)/\mathcal{C}). As pp is interdefinable with π(p)×f(p)\pi(p)\times f(p), this map is injective.

As f(p)f(p) is fundamental, the action of AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) is free and transitive. As f(p)f(p) is weakly orthogonal to {π(p),𝒞}\left\{\pi(p),\mathcal{C}\right\}, so is the action of AutA(f(p)/𝒞,π(p))\mathrm{Aut}_{A}(f(p)/\mathcal{C},\pi(p)), and therefore we must have AutA(f(p)/𝒞,π(p))=AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C},\pi(p))=\mathrm{Aut}_{A}(f(p)/\mathcal{C}), and by Lemma 2.10 we also obtain AutA(π(p)/𝒞,f(p))=AutA(π(p)/𝒞)\mathrm{Aut}_{A}(\pi(p)/\mathcal{C},f(p))=\mathrm{Aut}_{A}(\pi(p)/\mathcal{C}).

It is easy to see that AutA(π(p)/𝒞,f(p))×AutA(f(p)/𝒞,π(p))<Im(π~×f~)\mathrm{Aut}_{A}(\pi(p)/\mathcal{C},f(p))\times\mathrm{Aut}_{A}(f(p)/\mathcal{C},\pi(p))<\operatorname{Im}(\widetilde{\pi}\times\widetilde{f}): given (σ,τ)(\sigma,\tau) in the former group, we can extend each to σ^,τ^AutA(p/𝒞)\widehat{\sigma},\widehat{\tau}\in\mathrm{Aut}_{A}(p/\mathcal{C}), and then (π~×f~)(σ^τ^)=(σ,τ)(\widetilde{\pi}\times\widetilde{f})(\widehat{\sigma}\widehat{\tau})=(\sigma,\tau). Therefore the map π~×f~\widetilde{\pi}\times\widetilde{f} is an AA-definable isomorphism between AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}) and AutA(π(p)/𝒞)×AutA(f(p)/𝒞)\mathrm{Aut}_{A}(\pi(p)/\mathcal{C})\times\mathrm{Aut}_{A}(f(p)/\mathcal{C}). ∎

Note that even when f(p)f(p) is not fundamental, if there is an AA-definable section ss, by Corollary 2.7, the group Im(s)\operatorname{Im}(s) is an AA-definable normal subgroup of AutA(p/𝒞)\mathrm{Aut}_{A}(p/\mathcal{C}), and this group is AA-definably isomorphic to AutA(f(p)/𝒞)×Im(s)\mathrm{Aut}_{A}(f(p)/\mathcal{C})\times\operatorname{Im}(s).

There is one important case where the type f(p)f(p) is fundamental: if its binding group is abelian. Indeed, the action of AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) is faithful, and all stabilisers are conjugate, hence equal. So if some σAutA(f(p)/𝒞)\sigma\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) fixes some f(a)f(p)f(a)\models f(p), then its fixes every realisation of f(p)f(p), and thus must be the identity by faithfullness. We will use this to deal with binding groups isomorphic to tori.

The assumption that f(p)f(p) is fundamental can always be obtained, without changing the short exact sequence, by taking a high enough Morley power of pp (see [6, Fact 2.4]). However, applying Lemma 3.15 would yield a splitting of a Morley power of pp, which is not what we want. Note that the right to left direction of the proof only uses that f(p)f(p) is fundamental to obtain the equality AutA(f(p)/𝒞,π(p))=AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C},\pi(p))=\mathrm{Aut}_{A}(f(p)/\mathcal{C}). We could therefore replace the assumption that f(p)f(p) is fundamental by this and obtain the same implication.

Outside of the abelian case, we cannot expect the type f(p)f(p) 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 pS(A)p\in S(A) be a 𝒞\mathcal{C}-internal type and f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration. Suppose f(p)f(p) is weakly 𝒞\mathcal{C}-orthogonal and that the short exact sequence 1KAutA(p/𝒞)f~AutA(f(p)/𝒞)11\rightarrow K\rightarrow\mathrm{Aut}_{A}(p/\mathcal{C})\xrightarrow{\widetilde{f}}\mathrm{Aut}_{A}(f(p)/\mathcal{C})\rightarrow 1 is AA-definably split via some AA-definable section ss. If the following holds:

(\bigstar): for any f(a)f(p)f(a)\models f(p) and σAutA(f(p)/𝒞)\sigma\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}), if σ(f(a))=f(a)\sigma(f(a))=f(a), then s(σ)|pf(a)=ids(\sigma)|_{p_{f(a)}}=\operatorname{id}.

then GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable (and in particular ff splits).

Before starting with the proof, it is worth pointing out what assumption (\bigstar) means. If σ\sigma fixes f(a)f(a), then s(σ)s(\sigma) fixes pf(a)p_{f(a)}, the fiber above f(a)f(a), as a set. This assumption says that it in fact fixes pf(a)p_{f(a)} pointwise.

Proof.

Let f(a),f(b)f(p)f(a),f(b)\models f(p) and σ,τAutA(f(p)/𝒞)\sigma,\tau\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) such that σ(f(a))=f(b)=τ(f(a))\sigma(f(a))=f(b)=\tau(f(a)). Then s(σ)|pf(a)s(\sigma)|_{p_{f(a)}} and s(τ)|pf(a)s(\tau)|_{p_{f(a)}} are in Mor(f(a),f(b))\mathrm{Mor}(f(a),f(b)). By assumption (s(σ)|pf(a))1s(τ)|pf(a)=s(στ1)|pf(a)=id|pf(a)(s(\sigma)|_{p_{f(a)}})^{-1}\circ s(\tau)|_{p_{f(a)}}=s(\sigma\circ\tau^{-1})|_{p_{f(a)}}=\operatorname{id}|_{p_{f(a)}}, so s(σ)|pf(a)=s(τ)|pf(a)s(\sigma)|_{p_{f(a)}}=s(\tau)|_{p_{f(a)}}. Therefore, we can define, for any f(a),f(b)f(p)f(a),f(b)\models f(p), the map r(f(a),f(b))=s(σ)pf(a)r(f(a),f(b))=s(\sigma)_{p_{f(a)}}, for any σAutA(f(p)/𝒞)\sigma\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) with σ(f(a))=f(b)\sigma(f(a))=f(b). This is well-defined by the previous discussion, and it is straightforward to show that it witnesses retractability of GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}). ∎

Note that if f(p)f(p) is fundamental, then any σAutA(f(p)/𝒞)\sigma\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) fixing some f(a)f(p)f(a)\models f(p) must be the identity. Therefore in that case, we always have (\bigstar). If f(p)f(p) is not fundamental, we find the following sufficient condition for (\bigstar):

Corollary 3.17.

Under the same assumptions as Lemma 3.16, if for some (any) f(a)f(p)f(a)\models f(p), there is no non-trivial (i.e. with image not equal to the trivial subgroup) definable morphism from any definable subgroup of AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) to AutAf(a)(pf(a)/𝒞)\mathrm{Aut}_{Af(a)}(p_{f(a)}/\mathcal{C}) then assumption (\bigstar) holds. In particular ff splits.

Proof.

Fix f(a)f(p)f(a)\models f(p), let HH be the definable subgroup of elements AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) fixing f(a)f(a). Then ss gives rise to a definable map s:HAutf(a)A(pf(a)/𝒞)s^{\prime}:H\rightarrow\mathrm{Aut}_{f(a)A}(p_{f(a)}/\mathcal{C}), which must have trivial image. This is exactly saying that if σ(f(a))=f(a)\sigma(f(a))=f(a), then s(σ)s(\sigma) fixes pf(a)p_{f(a)} pointwise. ∎

To go from unipotent and tori to nilpotent, we need to remark that (\bigstar) is preserved under direct product. More precisely:

Lemma 3.18.

Let pS(A)p\in S(A) be a 𝒞\mathcal{C}-internal type and f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration. Suppose f(p)f(p) is weakly 𝒞\mathcal{C}-orthogonal and that the short exact sequence 1KAutA(p/𝒞)f~AutA(f(p)/𝒞)11\rightarrow K\rightarrow\mathrm{Aut}_{A}(p/\mathcal{C})\xrightarrow{\widetilde{f}}\mathrm{Aut}_{A}(f(p)/\mathcal{C})\rightarrow 1 is AA-definably split via some AA-definable section ss. Suppose that AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) is AA-definably isomorphic to a direct product H1×H2H_{1}\times H_{2}, that for all f(a)f(p)f(a)\models f(p) we have H1f(a)H2f(a)={f(a)}H_{1}f(a)\cap H_{2}f(a)=\{f(a)\}, and that the induced actions of H1H_{1} and H2H_{2} on f(p)f(p) both satisfy (\bigstar). Then the action of AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) on f(p)f(p) satisfies (\bigstar).

Proof.

We can write any σAutA(f(p)/𝒞)\sigma\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) as σ1σ2\sigma_{1}\sigma_{2} with σiHi\sigma_{i}\in H_{i}. Suppose that σ(f(a))=f(a)\sigma(f(a))=f(a) for some f(a)f(p)f(a)\models f(p). Then σ1(f(a))=σ21(f(a))\sigma_{1}(f(a))=\sigma_{2}^{-1}(f(a)), therefore σ1(f(a))=σ2(f(a))=f(a)\sigma_{1}(f(a))=\sigma_{2}(f(a))=f(a). By our (\bigstar) assumptions, we obtain that s(σi)|pf(a)=ids(\sigma_{i})|_{p_{f(a)}}=\operatorname{id}, and therefore s(σ)|pf(a)=ids(\sigma)|_{p_{f(a)}}=\operatorname{id}. ∎

Let us say a word about the assumptions of these lemmas before continuing. First, the assumption that the short exact sequence is AA-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 𝒰DCF0\mathcal{U}\models\mathrm{DCF}_{0} a sufficiently saturated model, and 𝒞\mathcal{C} its field of constants. If ss is a differential transcendental, Jin considers the 𝒞\mathcal{C}-internal types pp generic of δ1(s)\delta^{-1}(s) and q=δ1(p)q=\delta^{-1}(p). He shows that the short exact sequence obtained is isomorphic to an extension of Ga(𝒞)G_{a}(\mathcal{C}) by some unipotent group, and that the map δ:qp\delta:q\rightarrow p 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 qq.

As for the necessity of (\bigstar), 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 pS(A)p\in S(A) be a 𝒞\mathcal{C}-internal type and f:pf(p)f:p\rightarrow f(p) be an AA-definable fibration. Suppose that f(p)f(p) is weakly 𝒞\mathcal{C}-orthogonal and:

  • the short exact sequence 1KAutA(p/𝒞)f~AutA(f(p)/𝒞)11\rightarrow K\rightarrow\mathrm{Aut}_{A}(p/\mathcal{C})\xrightarrow{\widetilde{f}}\mathrm{Aut}_{A}(f(p)/\mathcal{C})\rightarrow 1 is AA-definably split via some AA-definable section ss and KK is finite,

  • the action of AutA(f(p)/𝒞)\mathrm{Aut}_{A}(f(p)/\mathcal{C}) satisfies (\bigstar).

Then the groupoid GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable, thus by Lemma 3.14 we get a map π:pπ(p)\pi:p\rightarrow\pi(p) such that pp is interdefinable with π(p)×f(p)\pi(p)\times f(p). Moreover π(p)(𝒰)acl(𝒞,A)\pi(p)(\mathcal{U})\subset\operatorname{acl}(\mathcal{C},A).

Proof.

By Lemma 3.16, we know that the groupoid GdA(p,f/𝒞)\mathrm{Gd}_{A}(p,f/\mathcal{C}) is retractable by the map rr defined by r(f(a),f(b))=s(σ)|pf(a)r(f(a),f(b))=s(\sigma)|_{p_{f(a)}}, for some (any) σAutA(f(p)/𝒞)\sigma\in\mathrm{Aut}_{A}(f(p)/\mathcal{C}) with σ(f(a))=f(b)\sigma(f(a))=f(b).

The map π\pi is defined to be the quotient map of the equivalence relation EE given by E(x,y)E(x,y) if and only if r(f(a),f(b))(a)=br(f(a),f(b))(a)=b, for any a,bpa,b\models p. By the previous paragraph, this is equivalent to s(σ)|pf(a)(a)=bs(\sigma)|_{p_{f(a)}}(a)=b.

Let σAutA(𝒰/𝒞)\sigma\in\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{C}), and apa\models p. Then there is τK\tau\in K such that σ|p=s(σ|f(p))τ\sigma|_{p}=s(\sigma|_{f(p)})\circ\tau. We then have:

σ(a)\displaystyle\sigma(a) =s(σ|f(p))(τ(a))\displaystyle=s(\sigma|_{f(p)})(\tau(a))
=s(σ|f(p))|pf(τ(a))(τ(a))\displaystyle=s(\sigma|_{f(p)})|_{p_{f(\tau(a))}}(\tau(a))
=r(f(τ(a)),f(σ(τ(a))))(τ(a))\displaystyle=r\Bigl{(}f\bigl{(}\tau(a)\bigr{)},f\bigl{(}\sigma(\tau(a))\bigr{)}\Bigr{)}(\tau(a))
=r(f(τ(a)),f(σ(a)))(τ(a)) as τK\displaystyle=r\Bigl{(}f\bigl{(}\tau(a)\bigr{)},f\bigl{(}\sigma(a)\bigr{)}\Bigr{)}(\tau(a))\text{ as }\tau\in K

which implies that E(τ(a),σ(a))E(\tau(a),\sigma(a)). As KK is finite, this means that the orbit of π(a)\pi(a) under AutA(𝒰/𝒞)\mathrm{Aut}_{A}(\mathcal{U}/\mathcal{C}) is finite. Hence π(p)(𝒰)acl(𝒞,A)\pi(p)(\mathcal{U})\subset\operatorname{acl}(\mathcal{C},A). ∎

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 DCF0\mathrm{DCF}_{0}, for which a reference is [16]. Fix some sufficiently saturated 𝒰DCF0\mathcal{U}\models\mathrm{DCF}_{0} with its derivation δ\delta. Also fix some algebraically closed differential subfield F<𝒰F<\mathcal{U}.

We let 𝒞\mathcal{C} be its field of constants. It is a pure algebraically closed field, and is stably embedded. In particular, this implies that definable groups in 𝒞\mathcal{C} 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:

logδ:𝒰{0}\displaystyle\log_{\delta}:\mathcal{U}\setminus\{0\} 𝒰\displaystyle\rightarrow\mathcal{U}
x\displaystyle x δ(x)x\displaystyle\rightarrow\frac{\delta(x)}{x}

which is a surjective morphism from (𝒰,)(\mathcal{U},\cdot) to (𝒰,+)(\mathcal{U},+) with kernel (𝒞,)(\mathcal{C},\cdot).

Definition 4.1.

Let pS1(F)p\in S_{1}(F). We define its pullback under the logarithmic derivative logδ\log_{\delta} to be q:=logδ1(p)q:=\log_{\delta}^{-1}(p), the type of any uu such that logδ(u)p\log_{\delta}(u)\models p and uacl(F,logδ(u))u\not\in\operatorname{acl}(F,\log_{\delta}(u)).

It is a known fact that this defines a unique complete type, and that each fiber qlogδ(u)q_{\log_{\delta}(u)} is a strongly minimal 𝒞\mathcal{C}-internal type (see [9, Proposition 5.3]).

In his thesis [9], Jin conjectures (Conjecture 5.4) the following, for such types pp and qq:

Conjecture 4.2.

If pp is almost 𝒞\mathcal{C}-internal, the following are equivalent:

  1. (1)

    qq is almost 𝒞\mathcal{C}-internal

  2. (2)

    logδ:qp\log_{\delta}:q\rightarrow p is almost split

  3. (3)

    there is an integer k0k\neq 0 such that for some uqu\models q, there are w1,w2w_{1},w_{2} such that:

    • uk=w1w2u^{k}=w_{1}w_{2}

    • w1dcl(F,logδu)w_{1}\in\operatorname{dcl}(F,\log_{\delta}u)

    • logδ(w2)dcl(F)\log_{\delta}(w_{2})\in\operatorname{dcl}(F)

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 (3)(2)(1)(3)\Rightarrow(2)\Rightarrow(1) (Remark 5.5). In their article [10], Jin and Moosa completely answer this question in the case where pp is the generic type of an equation δ(x)=f(x)\delta(x)=f(x), where fF(x)f\in F(x) and is 𝒞\mathcal{C}-internal. Their answer depends on the following fact, which itself is a consequence, in DCF0\mathrm{DCF}_{0}, of [16, Chapter 3, Corollary 1.6]:

Fact 4.3.

Let (G,S)(G,S) be a definable (in DCF0\mathrm{DCF}_{0}), faithful, transitive group action on a strongly minimal set SS. Then (G,S)(G,S) is definably isomorphic to either:

  1. (1)

    Ga(𝒞)G_{a}(\mathcal{C}), Gm(𝒞)G_{m}(\mathcal{C}) or E(𝒞)E(\mathcal{C}) for some elliptic curve EE over the constants, acting regularly on itself,

  2. (2)

    Ga(𝒞)Gm(𝒞)G_{a}(\mathcal{C})\rtimes G_{m}(\mathcal{C}), acting on 𝒞\mathcal{C} by affine transformations,

  3. (3)

    PSL2(𝒞)\mathrm{PSL}_{2}(\mathcal{C}) acting on (𝒞)\mathbb{P}(\mathcal{C}) by projective transformations.

Suppose that pS1(F)p\in S_{1}(F) is the generic type of δ(x)=f(x)\delta(x)=f(x), for some fF(X)f\in F(X), and is 𝒞\mathcal{C}-internal. They prove:

Theorem A of [10].

If the binding group of pp is not of dimension 33, then logδ1(p)\log_{\delta}^{-1}(p) is almost 𝒞\mathcal{C}-internal if and only if it product-splits (which implies that it is almost-split).

When the dimension is 33, 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 FF and a minimal type pS(F)p\in S(F) internal to the constants, with binding group PSL2(𝒞)\mathrm{PSL}_{2}(\mathcal{C}), such that logδ1(p)\log_{\delta}^{-1}(p) is 𝒞\mathcal{C}-internal but logδ:logδ1(p)p\log_{\delta}:\log_{\delta}^{-1}(p)\rightarrow p 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 𝒞\mathcal{C}-internality of logδ1(p)\log_{\delta}^{-1}(p), we need not assume pp to be almost 𝒞\mathcal{C}-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 logδ\log_{\delta} are not weakly 𝒞\mathcal{C}-orthogonal (and will be useful later):

Lemma 4.4.

Let qS1(F)q\in S_{1}(F) be a type over any algebraically closed differential subfield of 𝒰\mathcal{U}. Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q). If for some apa\models p, the fiber plogδ(a)p_{\log_{\delta}(a)} is not weakly 𝒞\mathcal{C}-orthogonal, then there is an integer k0k\neq 0, some constant μ\mu and some w1w_{1} such that:

  • ak=μw1a^{k}=\mu w_{1}

  • w1dcl(F,logδa)w_{1}\in\operatorname{dcl}(F,\log_{\delta}a).

In particular logδ:pq\log_{\delta}:p\rightarrow q product-splits.

Proof.

Let L=Flogδ(a)=dcl(F,logδ(a))L=F\langle\log_{\delta}(a)\rangle=\operatorname{dcl}(F,\log_{\delta}(a)). Non-weak 𝒞\mathcal{C}-orthogonality implies that dcl(aL)𝒞L𝒞\operatorname{dcl}(aL)\cap\mathcal{C}\neq L\cap\mathcal{C}. As logδ(a)L\log_{\delta}(a)\in L, we see that dcl(aL)=L(a)\operatorname{dcl}(aL)=L(a), the field generated by aa over LL, so there are f(X),g(X)L[X]f(X),g(X)\in L[X] and e𝒞Le\in\mathcal{C}\setminus L such that f(a)g(a)=e\frac{f(a)}{g(a)}=e. Write f(x)=i=0nbixif(x)=\sum\limits_{i=0}^{n}b_{i}x^{i} and g(x)=j=0mdjxjg(x)=\sum\limits_{j=0}^{m}d_{j}x^{j} for some bi,djLb_{i},d_{j}\in L (with bn0b_{n}\neq 0, dm0d_{m}\neq 0 and nmn\geq m, which we may assume by replacing e0e\neq 0 by 1e\frac{1}{e}). Taking a derivative, we obtain:

δ(f(a))g(a)δ(g(a))f(a)=0\delta(f(a))g(a)-\delta(g(a))f(a)=0

which is a polynomial identity in aa, with coefficients in LL. Recall that logδ1(q)\log_{\delta}^{-1}(q) is chosen so that aacl(F,logδ(a))=acl(L)a\not\in\operatorname{acl}(F,\log_{\delta}(a))=\operatorname{acl}(L). Therefore all coefficients of this polynomial are zero. Computing the dominant coefficient gives:

dm(δ(bn)+logδ(a)nbn)=bn(δ(dm)+logδ(a)mdm)d_{m}\left(\delta(b_{n})+\log_{\delta}(a)nb_{n}\right)=b_{n}\left(\delta(d_{m})+\log_{\delta}(a)md_{m}\right)

and as bn0b_{n}\neq 0 and dm0d_{m}\neq 0 we get:

logδ(anm)=logδ(dmbn)\log_{\delta}\left(a^{n-m}\right)=\log_{\delta}\left(\frac{d_{m}}{b_{n}}\right)

and thus there is μ𝒞\mu\in\mathcal{C} such that:

anm=μdmbna^{n-m}=\mu\frac{d_{m}}{b_{n}}

If nmn\neq m, this gives the result by picking w1=dmbnw_{1}=\frac{d_{m}}{b_{n}} and w2=μw_{2}=\mu. Otherwise, we see that dmbm=μ\frac{d_{m}}{b_{m}}=\mu, for some non-zero μ𝒞\mu\in\mathcal{C}. In this case, using euclidian division, we find hL[X]h\in L[X] with deg(h)<deg(f)\deg(h)<\deg(f) and such that h(a)g(a)=eμ\frac{h(a)}{g(a)}=e-\mu. As eLe\not\in L and μL\mu\in L, we get eμLe-\mu\not\in L. Since μ𝒞\mu\in\mathcal{C} and e𝒞e\in\mathcal{C}, we also get eμ𝒞e-\mu\in\mathcal{C}. 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 qS1(F)q\in S_{1}(F) be a type over any algebraically closed differential subfield of 𝒰\mathcal{U}. Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q). Suppose that p(𝒰)acl(logδ(p)(𝒰),𝒞,F)p(\mathcal{U})\not\subset\operatorname{acl}(\log_{\delta}(p)(\mathcal{U}),\mathcal{C},F). Then the fibration logδ:pq\log_{\delta}:p\to q is almost split if and only if it is product-split, i.e. there is an integer k0k\neq 0 such that for some apa\models p, there are w1,w2w_{1},w_{2} such that:

  • ak=w1w2a^{k}=w_{1}w_{2}

  • w1dcl(F,logδa)w_{1}\in\operatorname{dcl}(F,\log_{\delta}a)

  • logδ(w2)F\log_{\delta}(w_{2})\in F.

Proof.

As stated previously, that product-splitting implies almost splitting is essentially contained in [9, Remark 5.5].

Suppose that the map logδ:plogδ(p)\log_{\delta}:p\rightarrow\log_{\delta}(p) is almost split, so there is sS(F)s\in S(F) that is 𝒞\mathcal{C}-internal and such that any apa\models p is interalgebraic with logδ(a),b\log_{\delta}(a),b, for some bsb\models s with logδ(a)|Fb\log_{\delta}(a)\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}b. We fix such aa and let L=Flogδ(a)=dcl(F,logδ(a))L=F\langle\log_{\delta}(a)\rangle=\operatorname{dcl}(F,\log_{\delta}(a)), so that b|FLb\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}L.

As the types s|Ls|_{L} and plogδ(a)p_{\log_{\delta}(a)} are interalgebraic, their binding groups must be isogenous by Corollary 2.13. By Lemma 4.4, we may assume that plogδ(a)p_{\log_{\delta}(a)} is weakly orthogonal to 𝒞\mathcal{C} and do so for the rest of the proof. Hence the binding group of plogδ(a)p_{\log_{\delta}(a)} is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}), and the binding group of s|Ls|_{L} must be as well. We want to show that AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) is also definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}).

First, remark that AutF(s/𝒞,logδ(p))\mathrm{Aut}_{F}(s/\mathcal{C},\log_{\delta}(p)) is infinite. Suppose, on the contrary, that it is finite. Then s(𝒰)acl(𝒞,logδ(p)(𝒰),F)s(\mathcal{U})\subset\operatorname{acl}(\mathcal{C},\log_{\delta}(p)(\mathcal{U}),F), and as pp is interalgebraic with slogδ(p)s\otimes\log_{\delta}(p) over FF, this would imply that p(𝒰)acl(𝒞,logδ(p)(𝒰),F)p(\mathcal{U})\subset\operatorname{acl}(\mathcal{C},\log_{\delta}(p)(\mathcal{U}),F), contradicting our assumption.

By Fact 2.5, the binding group AutL(s|L/𝒞)\mathrm{Aut}_{L}(s|_{L}/\mathcal{C}) is a definable subgroup of AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}). Moreover by Lemma 2.8 the binding group AutF(s/𝒞,logδ(p))\mathrm{Aut}_{F}(s/\mathcal{C},\log_{\delta}(p)) is a normal FF-definable subgroup of AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}), and, by a proof similar to that of Fact 2.5, is also seen to be a definable subgroup of AutL(s|L/𝒞)\mathrm{Aut}_{L}(s|_{L}/\mathcal{C}). Since AutF(s/𝒞,logδ(p))\mathrm{Aut}_{F}(s/\mathcal{C},\log_{\delta}(p)) is infinite, as Gm(𝒞)G_{m}(\mathcal{C}) is connected and strongly minimal, it must be equal to AutL(s|L/𝒞)\mathrm{Aut}_{L}(s|_{L}/\mathcal{C}). In particular AutL(s|L/𝒞)\mathrm{Aut}_{L}(s|_{L}/\mathcal{C}) is normal in AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}).

Note that RM(s)=RM(s|L)=RM(plogδ(a))=1\operatorname{RM}(s)=\operatorname{RM}(s|_{L})=\operatorname{RM}(p_{\log_{\delta}(a)})=1, and as FF is algebraically closed, the type ss is strongly minimal. Moreover, as s|Ls|_{L} is weakly 𝒞\mathcal{C}-orthogonal, forking calculus along with an automorphism argument yields that ss is weakly 𝒞\mathcal{C}-orthogonal, and in particular AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) acts transitively and faithfully on ss. As it has a normal definable subgroup isomorphic to Gm(𝒞)G_{m}(\mathcal{C}), Fact 4.3 implies that AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}).

By [10, Lemma 4.1 (b)], there is edcl(bF)Fe\in\operatorname{dcl}(bF)\setminus F such that logδ(e)F\log_{\delta}(e)\in F (and in particular tp(e/F)\operatorname{tp}(e/F) is rr-internal). Note that e|Fbe\not\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}b, so in particular we also have bacl(eF)b\in\operatorname{acl}(eF), as RM(b/F)=1\operatorname{RM}(b/F)=1. So ee and bb are interalgebraic over FF. For the rest of the proof we can therefore replace bb by ee, and assume that logδ(b)F\log_{\delta}(b)\in F.

Let L(b)=LbL(b)=L\langle b\rangle be the (differential) field generated by bb over LL (the two are equal since logδ(b)F\log_{\delta}(b)\in F). We have that logδ(a)L(b)\log_{\delta}(a)\in L(b), but also aacl(L(b))a\in\operatorname{acl}(L(b)), which implies that aa is field-algebraic over L(b)L(b). Thus there is PL(b)[X]P\in L(b)[X] such that P(a)=0P(a)=0, we take PP monic and of minimal degree, so P=Xn+i=0n1ciXiP=X^{n}+\sum\limits_{i=0}^{n-1}c_{i}X^{i}. Then we compute:

0=δ(P(a))=nlogδ(a)an+i=0n1(δ(ci)+icilogδ(a))ai0=\delta(P(a))=n\log_{\delta}(a)a^{n}+\sum\limits_{i=0}^{n-1}(\delta(c_{i})+ic_{i}\log_{\delta}(a))a^{i}

and thus:

0=nlogδ(a)P(a)δ(P(a))=i=0n1(nlogδ(a)ciδ(ci)icilogδ(a))ai0=n\log_{\delta}(a)P(a)-\delta(P(a))=\sum\limits_{i=0}^{n-1}(n\log_{\delta}(a)c_{i}-\delta(c_{i})-ic_{i}\log_{\delta}(a))a^{i}

As we picked nn minimal, we obtain, for all in1i\leq n-1:

nlogδ(a)ciδ(ci)icilogδ(a)=0n\log_{\delta}(a)c_{i}-\delta(c_{i})-ic_{i}\log_{\delta}(a)=0

and thus, for any ci0c_{i}\neq 0:

logδ(ani)=logδ(ci) .\log_{\delta}\left(a^{n-i}\right)=\log_{\delta}(c_{i})\text{ .}

If there is some ci0c_{i}\neq 0, we obtain that anici𝒞L(b)(a)\frac{a^{n-i}}{c_{i}}\in\mathcal{C}\cap L(b)(a). Recall that plogδ(a)=tp(a/L)p_{\log_{\delta}(a)}=\operatorname{tp}(a/L) is, by assumption, weakly orthogonal to 𝒞\mathcal{C}. As bacl(aL)b\in\operatorname{acl}(aL), so is tp(a,b/L)\operatorname{tp}(a,b/L), which implies L(b)(a)𝒞=L𝒞L(b)(a)\cap\mathcal{C}=L\cap\mathcal{C}, and thus aniciL\frac{a^{n-i}}{c_{i}}\in L, so aniL(b)a^{n-i}\in L(b). If all cic_{i} are zero, then an=0L(b)a^{n}=0\in L(b). In any case, there is k>0k>0 such that akL(b)a^{k}\in L(b).

Thus there are f,gL[x]f,g\in L[x] such that ak=f(b)g(b)a^{k}=\frac{f(b)}{g(b)}. Taking derivatives, we obtain

ak=δ(f(b))g(b)f(b)δ(g(g))klogδ(a)g(b)2a^{k}=\frac{\delta(f(b))g(b)-f(b)\delta(g(g))}{k\log_{\delta}(a)g(b)^{2}}

and thus, by equating our two expressions for aka^{k}:

δ(f(b))g(b)(klogδ(a)g(b)+δ(g(b)))f(b)=0 .\delta(f(b))g(b)-(k\log_{\delta}(a)g(b)+\delta(g(b)))f(b)=0\text{ .}

Write f(x)=i=0neixif(x)=\sum\limits_{i=0}^{n}e_{i}x^{i} and g(x)=j=0mdjxjg(x)=\sum\limits_{j=0}^{m}d_{j}x^{j} for some ei,djLe_{i},d_{j}\in L (with en0e_{n}\neq 0 and dm0d_{m}\neq 0), we compute:

δ(f(b))\displaystyle\delta(f(b)) =i=0n(δ(ei)+ieilogδ(b))bi\displaystyle=\sum\limits_{i=0}^{n}\left(\delta(e_{i})+ie_{i}\log_{\delta}(b)\right)b^{i}
δ(g(b))\displaystyle\delta(g(b)) =j=0m(δ(dj)+jdjlogδ(b))bj\displaystyle=\sum\limits_{j=0}^{m}\left(\delta(d_{j})+jd_{j}\log_{\delta}(b)\right)b^{j}

In particular, since logδ(b)L\log_{\delta}(b)\in L, these expressions are given by polynomials in L(x)L(x) applied to bb. Therefore the equation (4.1):

δ(f(b))g(b)(klogδ(a)g(b)+δ(g(b)))f(b)=0\delta(f(b))g(b)-(k\log_{\delta}(a)g(b)+\delta(g(b)))f(b)=0

is a polynomial in bb over LL. Since RM(b/L)=1\operatorname{RM}(b/L)=1, all coefficients must be zero. In particular, examining the dominant coefficient, and using that cn0c_{n}\neq 0 and dm0d_{m}\neq 0, we get:

logδ(en)+nlogδ(b)klogδ(a)logδ(dm)mlogδ(b)=0\log_{\delta}(e_{n})+n\log_{\delta}(b)-k\log_{\delta}(a)-\log_{\delta}(d_{m})-m\log_{\delta}(b)=0

which simplifies into:

logδ(ak)=logδ(endmbnm)\log_{\delta}\left(a^{k}\right)=\log_{\delta}\left(\frac{e_{n}}{d_{m}}b^{n-m}\right)

and implies that there is μ𝒞\mu\in\mathcal{C} such that:

ak\displaystyle a^{k} =μendmbnm\displaystyle=\mu\frac{e_{n}}{d_{m}}b^{n-m}
=endm(μbnm)\displaystyle=\frac{e_{n}}{d_{m}}\left(\mu b^{n-m}\right)

Note that logδ(μbnm)=(nm)logδ(b)F\log_{\delta}\left(\mu b^{n-m}\right)=(n-m)\log_{\delta}(b)\in F. So setting w1=endmw_{1}=\frac{e_{n}}{d_{m}} and w2=μbnmw_{2}=\mu b^{n-m} gives us what we want. ∎

Over a field of constants, we obtain the equivalence in general :

Corollary 4.6.

Let qS1(F)q\in S_{1}(F) be a type over any algebraically closed differential field of constants. Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q). Then logδ:pq\log_{\delta}:p\to q is almost split if and only if there is an integer k0k\neq 0 such that for some apa\models p, there are w1,w2w_{1},w_{2} such that:

  • ak=w1w2a^{k}=w_{1}w_{2}

  • w1dcl(F,logδa)w_{1}\in\operatorname{dcl}(F,\log_{\delta}a)

  • logδ(w2)F\log_{\delta}(w_{2})\in F.

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 ss was definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}). Keeping the notation of that proof, we now show how to obtain this if FF is a field of constants.

By Fact 2.5, the binding group AutL(s|L/𝒞)\mathrm{Aut}_{L}(s|_{L}/\mathcal{C}) is a definable subgroup of AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}). Moreover, we still know that AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) acts transitively and faitfhfully on the strongly minimal type ss. Over the constants, the possible only binding groups in Fact 4.3 are Ga,GmG_{a},G_{m} and an elliptic curve (see the proof of [1, Theorem 3.9]), and therefore the only possible case is AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) being definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}). ∎

We also always obtain the equivalence if the base type is orthogonal to the constants:

Corollary 4.7.

Let qS1(F)q\in S_{1}(F) be a type over any algebraically closed differential field, and assume that qq is orthogonal to the constants. Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q). Then logδ:pq\log_{\delta}:p\to q is almost split if and only if there is an integer k0k\neq 0 such that for some apa\models p, there are w1,w2w_{1},w_{2} such that:

  • ak=w1w2a^{k}=w_{1}w_{2}

  • w1dcl(F,logδa)w_{1}\in\operatorname{dcl}(F,\log_{\delta}a)

  • logδ(w2)F\log_{\delta}(w_{2})\in F.

Proof.

Again, we keep the notation of the proof of Theorem 4.5. By that theorem, we may assume that p(𝒰)acl(logδ(p)(𝒰),𝒞,F)p(\mathcal{U})\subset\operatorname{acl}(\log_{\delta}(p)(\mathcal{U}),\mathcal{C},F), which implies that s(𝒰)acl(logδ(p)(𝒰),𝒞,F)s(\mathcal{U})\subset\operatorname{acl}(\log_{\delta}(p)(\mathcal{U}),\mathcal{C},F).

As q=logδ(p)q=\log_{\delta}(p) is orthogonal to the constants, we have that AutF(s/𝒞,logδ(p))=AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C},\log_{\delta}(p))=\mathrm{Aut}_{F}(s/\mathcal{C}). The former group is, by assumption, finite. Thus ss is 𝒞\mathcal{C}-algebraic. This implies, in particular, that for some (any) apa\models p, the fiber plogδ(a)p_{\log_{\delta}(a)} is not weakly 𝒞\mathcal{C}-orthogonal. By Lemma 4.4, we get product-splitting. ∎

The following lemma will be key in applications:

Lemma 4.8.

Let FF be an algebraically closed field, let qS(F)q\in S(F) be an almost 𝒞\mathcal{C}-internal type and f:qf(q)=pf:q\rightarrow f(q)=p be a fibration, with pp being 𝒞\mathcal{C}-internal and weakly 𝒞\mathcal{C}-orthogonal. Suppose that for some (any) aqa\models q:

  • qf(a)q_{f(a)} is 𝒞\mathcal{C}-internal and weakly 𝒞\mathcal{C}-orthogonal,

  • Autf(a)F(qf(a)/𝒞)\mathrm{Aut}_{f(a)F}(q_{f(a)}/\mathcal{C}) is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}).

Then there exists a type qq^{\prime}, internal to 𝒞\mathcal{C}, and a fibration logδ:qlogδ(q)\log_{\delta}:q^{\prime}\rightarrow\log_{\delta}(q^{\prime}) with logδ(q)dcl(p(𝒰)F)\log_{\delta}(q^{\prime})\subset\operatorname{dcl}(p(\mathcal{U})F). Moreover if logδ:qlogδ(q)\log_{\delta}:q^{\prime}\rightarrow\log_{\delta}(q^{\prime}) almost splits then f:qf(q)f:q\rightarrow f(q) almost splits.

Proof.

By Fact 3.7, there is a finite-to-one definable map π:qπ(q)\pi:q\rightarrow\pi(q) such that π(q)\pi(q) is 𝒞\mathcal{C}-internal. In particular, for any aqa\models q, it restricts to a finite-to-one map from qf(a)q_{f(a)} to tp(π(a)/f(a)F)\operatorname{tp}(\pi(a)/f(a)F). By Lemma 2.10, there is a finite-to-one definable morphism Autf(a)F(qf(a)/𝒞)Autf(a)F(tp(π(a)/f(a)F)/𝒞)\mathrm{Aut}_{f(a)F}(q_{f(a)}/\mathcal{C})\rightarrow\mathrm{Aut}_{f(a)F}(\operatorname{tp}(\pi(a)/f(a)F)/\mathcal{C}). As the former group is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}), so is the latter.

Fix some aqa\models q. Since qf(a)q_{f(a)} is weakly 𝒞\mathcal{C}-orthogonal, so is tp(π(a)/f(a)F)\operatorname{tp}(\pi(a)/f(a)F). Moreover, as p=f(q)p=f(q) is weakly orthogonal to 𝒞\mathcal{C}, we have that 𝒞Ff(a)=𝒞F\mathcal{C}\cap F\langle f(a)\rangle=\mathcal{C}\cap F, which is algebraically closed. By [10, Lemma 4.1], there is bFf(a),π(a)Ff(a)algb\in F\langle f(a),\pi(a)\rangle\setminus F\langle f(a)\rangle^{\mathrm{alg}} such that logδ(b)Ff(a)\log_{\delta}(b)\in F\langle f(a)\rangle. We let q=tp(b/F)q^{\prime}=\operatorname{tp}(b/F) and obtain a fibration logδ:qlogδ(q)\log_{\delta}:q^{\prime}\rightarrow\log_{\delta}(q^{\prime}) with 𝒞\mathcal{C}-internal strongly minimal fibers. Finally as bdcl(f(a),π(a),F)b\in\operatorname{dcl}(f(a),\pi(a),F), we see that qq^{\prime} is 𝒞\mathcal{C}-internal.

For the moreover part, we assume that logδ:qlogδ(q)\log_{\delta}:q^{\prime}\rightarrow\log_{\delta}(q^{\prime}) almost splits. Then there exists crS(F)c\models r\in S(F), with rr a 𝒞\mathcal{C}-internal type, such that bb is interalgebraic, over FF, with (logδ(b),c)(\log_{\delta}(b),c) and c|Flogδ(b)c\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}\log_{\delta}(b). We are going to show that (c,f(a))(c,f(a)) witnesses splitting of f:qf(q)f:q\rightarrow f(q).

Claim.

We have U(c/F)=1U(c/F)=1.

Proof of claim.

We know that bacl(f(a),F)b\not\in\operatorname{acl}(f(a),F), and as logδ(b)acl(f(a),F)\log_{\delta}(b)\in\operatorname{acl}(f(a),F), we also have bacl(logδ(b),F)b\not\in\operatorname{acl}(\log_{\delta}(b),F), thus U(logδ(b)/F)=U(b/F)1U(\log_{\delta}(b)/F)=U(b/F)-1. We obtain:

U(c/F)\displaystyle U(c/F) =U(c/logδ(b)F) as c|Flogδ(b)\displaystyle=U(c/\log_{\delta}(b)F)\text{ as }c\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}\log_{\delta}(b)
=U(clogδ(b)/F)U(logδ(b)/F)\displaystyle=U(c\log_{\delta}(b)/F)-U(\log_{\delta}(b)/F)
=U(b/F)U(b/F)+1\displaystyle=U(b/F)-U(b/F)+1
=1\displaystyle=1

On the other hand, note that logδ(b)Ff(a)\log_{\delta}(b)\in F\langle f(a)\rangle, therefore U(b/f(a)F)=1U(b/f(a)F)=1, thus U(c/f(a)F)=1U(c/f(a)F)=1, which implies that c|Ff(a)c\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}f(a).

By construction bdcl(aF)b\in\operatorname{dcl}(aF), and cacl(bF)c\in\operatorname{acl}(bF), so (c,f(a))acl(aF)(c,f(a))\in\operatorname{acl}(aF). Note that U(c,f(a)/F)=U(c/F)+U(f(a)/F)=1+U(f(a)/F)U(c,f(a)/F)=U(c/F)+U(f(a)/F)=1+U(f(a)/F). Since Autf(a)F(qf(a)/𝒞)\mathrm{Aut}_{f(a)F}(q_{f(a)}/\mathcal{C}) is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}) and qf(a)q_{f(a)} is weakly 𝒞\mathcal{C}-orthogonal, we see that U(qf(a))=1U(q_{f(a)})=1, which implies that U(f(a)/F)=U(a/F)1U(f(a)/F)=U(a/F)-1, and combining this with the previous equality we get U(c,f(a)/F)=U(a/F)U(c,f(a)/F)=U(a/F). As (c,f(a))acl(aF)(c,f(a))\in\operatorname{acl}(aF), this implies that aa and (c,f(a))(c,f(a)) are interalgebraic over FF, and f:qf(q)f:q\rightarrow f(q) splits. ∎

Finally, we prove a criteria for product-splitting of generic types of differential equations of arbitrary high order. More specifically, if FF is some algebraically closed differential subfield of 𝒰\mathcal{U}, we will be interested definable sets given by solution sets of equations of the form δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)), with fF(x0,,xm1)f\in F(x_{0},\cdots,x_{m-1}). Note that we can write f=PQf=\frac{P}{Q} with P,QF[x0,,xm1]P,Q\in F[x_{0},\cdots,x_{m-1}] without common factors, and that our definable set is then given by solutions of the equation δm(y)Q(y,δ(y),,δm1(y))P(y,δ(y),,δm1(y))=0\delta^{m}(y)Q(y,\delta(y),\cdots,\delta^{m-1}(y))-P(y,\delta(y),\cdots,\delta^{m-1}(y))=0. As the polynomial xmP(x0,,xm1)Q(x0,,xm1)F[x0,,xm]x_{m}P(x_{0},\cdots,x_{m-1})-Q(x_{0},\cdots,x_{m-1})\in F[x_{0},\cdots,x_{m}] is irreducible, this definable set has a unique generic type.

Recall that if (F,δ)(F,\delta) is any differential field, then the field F(x0,,xm)F(x_{0},\cdots,x_{m}) can be equipped with the derivations xi\frac{\partial}{\partial x_{i}} with respect to xix_{i}, as well as the derivation δF\delta^{F}, which treats all xix_{i} as constants and equals δ\delta on FF. If PF[x0,,xm]P\in F[x_{0},\cdots,x_{m}], then δF(P)=Pδ\delta^{F}(P)=P^{\delta}, i.e. the polynomial obtained by differentiating the coefficients of PP. Using this notation, we can compute, for any tuple (a0,,am1)𝒰m(a_{0},\cdots,a_{m-1})\in\mathcal{U}^{m} and hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}), that δ(h(a¯))=i=0m1hxiδ(ai)+δF(h)(a¯)\delta(h(\overline{a}))=\sum\limits_{i=0}^{m-1}\frac{\partial h}{\partial x_{i}}\delta(a_{i})+\delta^{F}(h)(\overline{a}). We obtain the following:

Lemma 4.9.

Let FF be an algebraically closed differential field, a rational function fF(x0,,xm1)f\in F(x_{0},\cdots,x_{m-1}) and qq the generic type of δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)). Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q), then logδ:pq\log_{\delta}:p\rightarrow q is product-split if and only if there are a non-zero hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}), some eFe\in F and some integer k0k\neq 0 such that

(kx0e)h=i=0m2hxixi+1+hxm1f+δF(h) .(kx_{0}-e)h=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}x_{i+1}+\frac{\partial h}{\partial x_{m-1}}f+\delta^{F}(h)\text{ .}
Proof.

Suppose that there are such h,eh,e and kk, consider some upu\models p and a=logδ(u)qa=\log_{\delta}(u)\models q. Also denote (a,δ(a),,δm1(a))=a¯(a,\delta(a),\cdots,\delta^{m-1}(a))=\overline{a}. As aa realizes the generic type of δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)), which is of order mm, we have that h(a¯)h(\overline{a}) is well-defined and non-zero, as otherwise aa would satisfy a differential equation of order m1m-1.

Let w1=h(a¯)w_{1}=h(\overline{a}) and w2=ukh(a¯)w_{2}=\frac{u^{k}}{h(\overline{a})}, so that uk=w1w2u^{k}=w_{1}w_{2}, we need to show that w1dcl(F,logδ(u))w_{1}\in\operatorname{dcl}(F,\log_{\delta}(u)) and logδ(w2)F\log_{\delta}(w_{2})\in F. The first part is immediate, as a=logδ(u)a=\log_{\delta}(u) and hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}). We also compute:

logδ(w2)\displaystyle\log_{\delta}(w_{2}) =kalogδ(h(a¯))\displaystyle=ka-\log_{\delta}(h(\overline{a}))
=kaδ(h(a¯))h(a¯)\displaystyle=ka-\frac{\delta(h(\overline{a}))}{h(\overline{a})}
=kai=0m1hxi(a¯)δi+1(a)+δF(h)(a¯)h(a¯)\displaystyle=ka-\frac{\sum\limits_{i=0}^{m-1}\frac{\partial h}{\partial x_{i}}(\overline{a})\delta^{i+1}(a)+\delta^{F}(h)(\overline{a})}{h(\overline{a})}
=kai=0m2hxi(a¯)δi+1(a)+hxm1(a¯)f(a¯)+δF(h)(a¯)h(a¯)\displaystyle=ka-\frac{\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}(\overline{a})\delta^{i+1}(a)+\frac{\partial h}{\partial x_{m-1}}(\overline{a})f(\overline{a})+\delta^{F}(h)(\overline{a})}{h(\overline{a})}
=ka(kae)\displaystyle=ka-(ka-e)
=eF .\displaystyle=e\in F\text{ .}

For the converse, suppose that logδ:pq\log_{\delta}:p\rightarrow q is product-split. There are upu\models p, some integer k0k\neq 0 and w1,w2w_{1},w_{2} such that:

  • uk=w1w2u^{k}=w_{1}w_{2}

  • w1dcl(F,logδu)w_{1}\in\operatorname{dcl}(F,\log_{\delta}u)

  • logδ(w2)F\log_{\delta}(w_{2})\in F.

Note that neither w1w_{1} or w2w_{2} is zero as u0u\neq 0. We let a=logδ(u)a=\log_{\delta}(u) and e=logδ(w2)e=\log_{\delta}(w_{2}). Again denote a¯=(a,δ(a),,δm1(a))\overline{a}=(a,\delta(a),\cdots,\delta^{m-1}(a)). As δm(a)=f(a¯)\delta^{m}(a)=f(\overline{a}) and w1dcl(F,a)w_{1}\in\operatorname{dcl}(F,a), there is a non-zero hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}) such that w1=h(a¯)w_{1}=h(\overline{a}). Applying the logarithmic derivative to the equality uk=w1w2u^{k}=w_{1}w_{2}, we compute:

(kae)h(a¯)\displaystyle(ka-e)h(\overline{a}) =δ(h(a¯))\displaystyle=\delta(h(\overline{a}))
=i=0m1hxi(a¯)δi+1(a)+δF(h)(a¯)\displaystyle=\sum\limits_{i=0}^{m-1}\frac{\partial h}{\partial x_{i}}(\overline{a})\delta^{i+1}(a)+\delta^{F}(h)(\overline{a})
=i=0m2hxi(a¯)δi+1(a)+hxm1(a¯)f(a¯)+δF(h)(a¯) .\displaystyle=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}(\overline{a})\delta^{i+1}(a)+\frac{\partial h}{\partial x_{m-1}}(\overline{a})f(\overline{a})+\delta^{F}(h)(\overline{a})\text{ .}

Because aa realizes the generic type of δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)) and this equation is of order m1m-1, this implies:

(kx0e)h=i=0m2hxixi+1+hxm1f+δF(h) .(kx_{0}-e)h=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}x_{i+1}+\frac{\partial h}{\partial x_{m-1}}f+\delta^{F}(h)\text{ .}

Under the additional assumption of Theorem 4.5, we get a criteria for almost splitting:

Corollary 4.10.

Let FF be an algebraically closed differential field, some fF(x0,,xm1)f\in F(x_{0},\cdots,x_{m-1}) and qq the generic type of δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)). Let p:=logδ1(q)p:=\log_{\delta}^{-1}(q) and assume that p(𝒰)acl(logδ(p)(𝒰),𝒞,F)p(\mathcal{U})\not\subset\operatorname{acl}(\log_{\delta}(p)(\mathcal{U}),\mathcal{C},F). Then logδ:pq\log_{\delta}:p\rightarrow q is almost split if and only if there are a non-zero hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}), some eFe\in F and some integer k0k\neq 0 such that

(kx0e)h=i=0m2hxixi+1+hxm1f+δF(h) .(kx_{0}-e)h=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}x_{i+1}+\frac{\partial h}{\partial x_{m-1}}f+\delta^{F}(h)\text{ .}
Proof.

Immediate consequence of Theorem 4.5 and Lemma 4.12. ∎

4.2. Splitting when the binding group is nilpotent

In this subsection, we prove that for any 𝒞\mathcal{C}-internal and weakly 𝒞\mathcal{C}-orthogonal type pp over any algebraically closed field FF, if pp has nilpotent binding group, then logδ1(p)\log_{\delta}^{-1}(p) is almost 𝒞\mathcal{C}-internal if and only if logδ:logδ1(p)p\log_{\delta}:\log_{\delta}^{-1}(p)\rightarrow p 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 𝒞\mathcal{C}-points of a nilpotent linear algebraic group.

Theorem 4.11.

Let FF be an algebraically closed differential field and let pS1(F)p\in S_{1}(F) be a 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-orthogonal type with binding group definably isomorphic to the 𝒞\mathcal{C}-points of a nilpotent linear algebraic group. Then the following are equivalent:

  1. (1)

    logδ1(p)\log_{\delta}^{-1}(p) is almost 𝒞\mathcal{C}-internal,

  2. (2)

    logδ:logδ1(p)p\log_{\delta}:\log_{\delta}^{-1}(p)\rightarrow p product-splits,

  3. (3)

    logδ:logδ1(p)p\log_{\delta}:\log_{\delta}^{-1}(p)\rightarrow p almost splits.

Proof.

The implication (2)(3)(2)\Rightarrow(3), as previously discussed, is essentially given by [9, Remark 5.5], and the implication (3)(1)(3)\Rightarrow(1) is immediate. Therefore we only prove (1)(2)(1)\Rightarrow(2). In fact, we will prove (1)(3)(1)\Rightarrow(3) and ((1) and (3))(2)\bigl{(}(1)\text{ and }(3)\bigr{)}\Rightarrow(2).

Assume that q=logδ1(p)q=\log_{\delta}^{-1}(p) is almost 𝒞\mathcal{C}-internal. We will start by showing that this implies almost splitting of logδ\log_{\delta}. To use binding groups, we need to reduce to the case where qq is 𝒞\mathcal{C}-internal.

First note that if qlogδ(a)q_{\log_{\delta}(a)} is not weakly 𝒞\mathcal{C}-orthogonal, by Lemma 4.4 logδ:logδ1(p)p\log_{\delta}:\log_{\delta}^{-1}(p)\rightarrow p product-splits. So we now assume that qlogδ(a)q_{\log_{\delta}(a)} is weakly 𝒞\mathcal{C}-orthogonal.

By Lemma 4.8, there is another type qq^{\prime}, internal to 𝒞\mathcal{C}, and a fibration logδ:qlogδ(q)\log_{\delta}:q^{\prime}\rightarrow\log_{\delta}(q^{\prime}) with logδ(q)dcl(p(𝒰)F)\log_{\delta}(q^{\prime})\subset\operatorname{dcl}(p(\mathcal{U})F). This last part implies that the binding group of logδ(q)\log_{\delta}(q^{\prime}) is also definably isomorphic to the 𝒞\mathcal{C}-points of a nilpotent linear algebraic group. Moreover if logδ:qlogδ(q)\log_{\delta}:q^{\prime}\rightarrow\log_{\delta}(q^{\prime}) almost splits, so does logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q). Therefore, replacing qq by qq^{\prime}, we may assume that qq is 𝒞\mathcal{C}-internal. Because p=logδ(q)p=\log_{\delta}(q) is weakly 𝒞\mathcal{C}-orthogonal and each fiber qlogδ(a)q_{\log_{\delta}(a)} is weakly 𝒞\mathcal{C}-orthogonal, the type qq is also weakly 𝒞\mathcal{C}-orthogonal. By [6, Corollary 2.5], the binding group AutF(q/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C}) is connected. We get a short exact sequence:

1KAutF(q/𝒞)logδ~AutF(p/𝒞)11\rightarrow K\rightarrow\mathrm{Aut}_{F}(q/\mathcal{C})\xrightarrow{\widetilde{\log_{\delta}}}\mathrm{Aut}_{F}(p/\mathcal{C})\rightarrow 1

By [6, Lemma 2.7] KK is definably isomorphic to AutFlogδ(a¯)(tp(a¯/Flogδ(a¯))/𝒞)\mathrm{Aut}_{F\log_{\delta}(\bar{a})}(\operatorname{tp}(\bar{a}/F\log_{\delta}(\bar{a}))/\mathcal{C}), where a¯=a1,,an\bar{a}=a_{1},\cdots,a_{n} is Morley sequence in qq such that a¯\bar{a} is a fundamental system of realizations of qq, and logδ(a¯)\log_{\delta}(\bar{a}) is a fundamental system of realizations for pp. As AutFlogδ(a)(tp(a/Flogδ(a))/𝒞)\mathrm{Aut}_{F\log_{\delta}(a)}(\operatorname{tp}(a/F\log_{\delta}(a))/\mathcal{C}) is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}) for any aqa\models q, this implies that KK is definably isomorphic to a subgroup of Gm(𝒞)nG_{m}(\mathcal{C})^{n}. We would like to apply Lemma 2.18 to get a definable splitting of this short exact sequence. However KK need not be connected in general, and thus we cannot get splitting. We will now show how to replace KK by its connected component by replacing qq with an interalgebraic type.

The kernel KK is a definable d-group, and by Lemma 2.18, the binding group AutF(q/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C}) is a definable nilpotent linear algebraic group. Moreover, by [4, Theorem 16.2], it is equal to K×RK^{\circ}\times R, where RR is a finite FF-definable subgroup and KK^{\circ} is a torus.

The subgroup RR induces an equivalence relation EE on qq by E(x,y)E(x,y) if and only if there is σR\sigma\in R such that σ(x)=y\sigma(x)=y. Let π\pi be the quotient map. We can consider the quotient type π(q)\pi(q), we have an induced group morphism π~:AutF(q/𝒞)AutF(π(q)/𝒞)\widetilde{\pi}:\mathrm{Aut}_{F}(q/\mathcal{C})\rightarrow\mathrm{Aut}_{F}(\pi(q)/\mathcal{C}).

Claim.

ker(π~)=R\ker(\widetilde{\pi})=R.

Proof.

The kernel of π~\widetilde{\pi} is the set of τAutF(q/𝒞)\tau\in\mathrm{Aut}_{F}(q/\mathcal{C}) such that for all xqx\models q, there is σR\sigma\in R such that τ(x)=σ(x)\tau(x)=\sigma(x). Note that as qq is 𝒞\mathcal{C}-internal and weakly 𝒞\mathcal{C}-orthogonal, it is isolated, and therefore the action of AutF(q/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C}) is an action of a definable nilpotent linear algebraic group on a definable set. In particular, for any xqx\models q, by Lemma 2.19, if τ\tau is in its unipotent radical and σ\sigma is its maximal torus, then τ(x)=σ(x)\tau(x)=\sigma(x) implies σ(x)=x=τ(x)\sigma(x)=x=\tau(x). As RR is a subgroup of the maximal torus of AutF(q/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C}) and the action on qq is faithful, this implies that ker(π~)\ker(\widetilde{\pi}) intersects the unipotent radical of AutF(q/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C}) trivially. Again by Lemma 2.19 (or rather, its proof), we see that ker(π~)\ker(\widetilde{\pi}) is a subgroup of the maximal torus of AutF(q/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C}). Now, if τker(π~)\tau\in\ker(\widetilde{\pi}), there must be some xqx\models q and σR\sigma\in R such that σ(x)=τ(x)\sigma(x)=\tau(x). But the maximal torus is abelian, thus its action on qq is regular. This implies τ=σ\tau=\sigma. ∎

Since R<ker(logδ~)R<\ker(\widetilde{\log_{\delta}}), we obtain a definable map f:π(q)logδ(q)f:\pi(q)\rightarrow\log_{\delta}(q) by defining f(π(x))=logδ(x)f(\pi(x))=\log_{\delta}(x). This map induces the short exact sequence:

1KAutF(q/𝒞)/Rf~AutF(p/𝒞)11\rightarrow K^{\circ}\rightarrow\mathrm{Aut}_{F}(q/\mathcal{C})/R\xrightarrow{\widetilde{f}}\mathrm{Aut}_{F}(p/\mathcal{C})\rightarrow 1

The group AutF(q/𝒞)/R=AutF(π(q)/𝒞)\mathrm{Aut}_{F}(q/\mathcal{C})/R=\mathrm{Aut}_{F}(\pi(q)/\mathcal{C}) 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 𝒞\mathcal{C}-points of a nilpotent linear algebraic group. Since KK^{\circ} is connected, we now can apply the second part of Lemma 2.18.

Note that if we can show that f:π(q)logδ(q)f:\pi(q)\rightarrow\log_{\delta}(q) splits, then we automatically obtain almost splitting of logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q), because π\pi has finite fibers. This is what we will now prove.

The binding group AutF(π(q)/𝒞)\mathrm{Aut}_{F}(\pi(q)/\mathcal{C}) 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 𝒞\mathcal{C}-point of a nilpotent linear algebraic group.

Moreover, remark that the maximal tori of AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) and AutF(π(q)/𝒞)\mathrm{Aut}_{F}(\pi(q)/\mathcal{C}) are FF-definably isomorphic to the 𝒞\mathcal{C}-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 FF-definably isomorphic to the 𝒞\mathcal{C}-points of tori.

By Lemma 2.18, the short exact sequence of binding groups splits FF-definably. We know that G=AutF(p/𝒞)=Gu×TG=\mathrm{Aut}_{F}(p/\mathcal{C})=G_{u}\times T, where TT is its maximal torus. By Lemma 2.19, for any apa\models p, we have GuaTa={a}G_{u}a\cap Ta=\{a\}. Because GuG_{u} is unipotent, by Lemma 3.17, its action on pp satisfies (\bigstar). Because TT is abelian, its action also satisfies (\bigstar). Therefore by Lemma 3.18 the action of GG satisfies (\bigstar). By Lemma 3.16, the map f:π(q)pf:\pi(q)\rightarrow p splits, thus logδ:qp\log_{\delta}:q\rightarrow p almost splits.

We still need to obtain product-splitting. By Theorem 4.5, we may assume that q(𝒰)acl(logδ(q)(𝒰),𝒞,F)q(\mathcal{U})\subset\operatorname{acl}(\log_{\delta}(q)(\mathcal{U}),\mathcal{C},F). This also implies that π(q)(𝒰)acl(logδ(q)(𝒰),𝒞,F)\pi(q)(\mathcal{U})\subset\operatorname{acl}(\log_{\delta}(q)(\mathcal{U}),\mathcal{C},F). Using similar ideas as for the proof of Corollary 2.13, we see that this implies that AutF(π(q)/logδ(q),𝒞)\mathrm{Aut}_{F}(\pi(q)/\log_{\delta}(q),\mathcal{C}) is finite. This group is the kernel of the map f~:AutF(π(q)/𝒞)AutF(logδ(q)/𝒞)\widetilde{f}:\mathrm{Aut}_{F}(\pi(q)/\mathcal{C})\rightarrow\mathrm{Aut}_{F}(\log_{\delta}(q)/\mathcal{C}). By Lemma 3.19, we obtain that the fibers of ff are not weakly 𝒞\mathcal{C}-orthogonal, and thus neither are the fibers of logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q). But we assumed that they were, a contradiction. ∎

We deduce some internality criteria:

Corollary 4.12.

Let FF be an algebraically closed differential field, some rational function fF(x0,,xm1)f\in F(x_{0},\cdots,x_{m-1}), and consider qq, the generic type of δm(y)=f(y,δ(y),,δm1(y))\delta^{m}(y)=f(y,\delta(y),\cdots,\delta^{m-1}(y)). If qq is 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-orthogonal and has a nilpotent binding group, then p:=logδ1(q)p:=\log_{\delta}^{-1}(q) is almost 𝒞\mathcal{C}-internal if and only if there are a non-zero hF(x0,,xm1)h\in F(x_{0},\cdots,x_{m-1}), some eFe\in F and some integer k0k\neq 0 such that

(kx0e)h=i=0m2hxixi+1+hxm1f+δF(h) .(kx_{0}-e)h=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}x_{i+1}+\frac{\partial h}{\partial x_{m-1}}f+\delta^{F}(h)\text{ .}
Proof.

This is an immediate consequence of Lemma 4.9 and Theorem 4.11. ∎

In practice, we can now sometimes reduce the question of internality of a logarithmic differential pullback to some valuation computation. Consider some differential field FF and the rational function field F(x0,,xm1)F(x_{0},\cdots,x_{m-1}), recall that it is equipped with the derivations xi\frac{\partial}{\partial x_{i}}, as well as the derivation δF\delta^{F} that is equal to δ\delta on FF, and δF(xi)=0\delta^{F}(x_{i})=0 for all ii.

We can also equip it with valuations: see F(x0,,xm)F(x_{0},\cdots,x_{m}) as F(x1,,xm)(x0)F(x_{1},\cdots,x_{m})(x_{0}), and equip it with the valuation given, for any P,QF(x1,,xm)[x0]P,Q\in F(x_{1},\cdots,x_{m})[x_{0}], by v(P)=deg(P)v(P)=-\mathrm{deg}(P), and v(PQ)=v(P)v(Q)v(\frac{P}{Q})=v(P)-v(Q). We have, for all P,QF(x1,,xm)[x0]P,Q\in F(x_{1},\cdots,x_{m})[x_{0}] and ii, that v(Pxi)v(P)v\left(\frac{\partial P}{\partial x_{i}}\right)\geq v(P) and v(δF(P))v(P)v(\delta^{F}(P))\geq v(P), from which we deduce that v(PQxi)v(PQ)v\left(\frac{\partial\frac{P}{Q}}{\partial x_{i}}\right)\geq v\left(\frac{P}{Q}\right) and v(δF(PQ))v(PQ)v(\delta^{F}(\frac{P}{Q}))\geq v(\frac{P}{Q}). We could do this for any 0im0\leq i\leq m, and we denote viv_{i} the valuation obtained.

As an example of application, we look at linear differential equations:

Corollary 4.13.

Let FF be an algebraically closed differential field, and consider L(x0,,xm1)=i=0m1cixi+cF[x0,,xm1]L(x_{0},\cdots,x_{m-1})=\sum\limits_{i=0}^{m-1}c_{i}x_{i}+c\in F[x_{0},\cdots,x_{m-1}]. Let qq be the generic type of δm(x)=L(δm1(x),,δ(x),x)\delta^{m}(x)=L(\delta^{m-1}(x),\cdots,\delta(x),x), which is always 𝒞\mathcal{C}-internal. If qq is weakly 𝒞\mathcal{C}-orthogonal and has nilpotent binding group, then logδ1(q)\log_{\delta}^{-1}(q) is not almost 𝒞\mathcal{C}-internal.

Proof.

Suppose, for a contradiction, that logδ1(q)\log_{\delta}^{-1}(q) is almost 𝒞\mathcal{C}-internal. By Corollary 4.12, there are a non-zero hF(x0,,xm)h\in F(x_{0},\cdots,x_{m}), some eFe\in F and some integer k0k\neq 0 such that we have the following equality, in F(x0,,xm1)F(x_{0},\cdots,x_{m-1}):

(kx0e)h=i=0m2hxixi+1+hxm1(j=0m1cjxj+c)+δF(h) .(kx_{0}-e)h=\sum\limits_{i=0}^{m-2}\frac{\partial h}{\partial x_{i}}x_{i+1}+\frac{\partial h}{\partial x_{m-1}}\left(\sum\limits_{j=0}^{m-1}c_{j}x_{j}+c\right)+\delta^{F}(h)\text{ .}

Let v0v_{0} be the valuation in F(x0,,xm1)F(x_{0},\cdots,x_{m-1}) with respect to x0x_{0}. The coefficients of x0v0(h)+1x_{0}^{-v_{0}(h)+1} must be equal on both sides of the equality. We can write h=h~x0v0(h)+g(x0,,xm1)h=\widetilde{h}x_{0}^{-v_{0}(h)}+g(x_{0},\cdots,x_{m-1}) with h~F(x1,,xm1)\widetilde{h}\in F(x_{1},\cdots,x_{m-1}) and v0(g)>v0(h)v_{0}(g)>v_{0}(h). We have:

kh~=c0h~xm1 .k\widetilde{h}=c_{0}\frac{\partial\widetilde{h}}{\partial x_{m-1}}\text{ .}

Now consider the valuation vm1v_{m-1} with respect to xm1x_{m-1}, the left-hand side has valuation vm1(h~)v_{m-1}(\widetilde{h}), but the right-hand side has valuation strictly greater than vm1(h~)v_{m-1}(\widetilde{h}), unless h~=0\widetilde{h}=0, which is a contradiction, as then g=hg=h, but v0(g)>v0(h)v_{0}(g)>v_{0}(h) (note that this also works in the case where m=1m=1 and h~F\widetilde{h}\in F). ∎

If FF is a field of constants, then this would apply as long as the generic type of δm(x)=L(δm1(x),,δ(x),x)\delta^{m}(x)=L(\delta^{m-1}(x),\cdots,\delta(x),x) is weakly 𝒞\mathcal{C}-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 δm(x)=0\delta^{m}(x)=0 is never weakly 𝒞\mathcal{C}-orthogonal as long as m>1m>1.

Over non-constant parameters, we could pick some differential transcendental cc, and let F=calgF=\mathbb{Q}\langle c\rangle^{\mathrm{alg}}. Then this would apply to the generic type of δm(x)=c\delta^{m}(x)=c, which one can easily show is 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-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 pS1(F)p\in S_{1}(F), with FF algebraically closed, such that pp is 𝒞\mathcal{C}-internal with binding group an elliptic curve, with the type logδ1(p)\log_{\delta}^{-1}(p) almost 𝒞\mathcal{C}-internal, but the map logδ:logδ1(p)p\log_{\delta}:\log_{\delta}^{-1}(p)\rightarrow p 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 Ext(,)\mathrm{Ext}(\cdot,\cdot) 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 kk, over which everything will be defined, and a |k||k|-saturated algebraically closed field 𝒞k\mathcal{C}\supset k (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:

1BCA11\rightarrow B\rightarrow C\rightarrow A\rightarrow 1

of algebraic groups, where the morphisms are maps of algebraic groups. Two such extensions CC and CC^{\prime} are isomorphic if there exists a map f:CCf:C\rightarrow C^{\prime} making the diagram:

1{1}B{B}C{C}A{A}1{1}1{1}B{B}C{C^{\prime}}A{A}1{1}id\scriptstyle{\operatorname{id}}f\scriptstyle{f}id\scriptstyle{\operatorname{id}}

commute. In that case, ff must be an isomorphism. For any two commutative algebraic groups AA and BB, denote Ext(A,B)\mathrm{Ext}(A,B) 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 Ext(,)\mathrm{Ext}(\cdot,\cdot) into a functor.

Given a morphism f:BBf:B\rightarrow B^{\prime}, there exists a unique extension CExt(A,B)C^{\prime}\in\mathrm{Ext}(A,B^{\prime}) and a map F:CCF:C\rightarrow C^{\prime} making the following diagram:

0{0}B{B}C{C}A{A}0{0}0{0}B{B^{\prime}}C{C^{\prime}}A{A}0{0}f\scriptstyle{f}F\scriptstyle{F}id\scriptstyle{\operatorname{id}}

commute. We denote this extension f(C)f_{*}(C). If g:AAg:A^{\prime}\rightarrow A is a morphism, there is a similar construction of g(C)g^{*}(C). We give details on the construction, as we will need them later. Let:

0BC𝑓A00\rightarrow B\rightarrow C\xrightarrow{f}A\rightarrow 0

be an element of Ext(A,B)\mathrm{Ext}(A,B). There is a unique CExt(A,B)C^{\prime}\in\mathrm{Ext}(A^{\prime},B) and map G:CCG:C^{\prime}\rightarrow C making the diagram:

0{0}B{B}C{C^{\prime}}A{A^{\prime}}0{0}0{0}B{B}C{C}A{A}0{0}id\scriptstyle{\operatorname{id}}G\scriptstyle{G}g\scriptstyle{g}f\scriptstyle{f}

commute. Moreover, the group CC^{\prime} is the subgroup of A×CA^{\prime}\times C consisting of pairs (a,c)(a,c) such that g(a)=f(c)g(a)=f(c), and the maps CAC^{\prime}\rightarrow A^{\prime} and G:CCG:C^{\prime}\rightarrow C are the natural projections.

The following summarizes the properties we will need (see [25, Chapter 7, 1.1]):

Fact 4.14.

Let 𝒜\mathcal{A} be the category of commutative algebraic groups. The previous construction makes Ext(,)\mathrm{Ext}(\cdot,\cdot) into an additive bifunctor on 𝒜×𝒜\mathcal{A}\times\mathcal{A}, contravariant in the first coordinate and covariant in the second. In particular Ext(A,B)\mathrm{Ext}(A,B) is always an algebraic group, and the maps ff_{*} and gg^{*} are morphisms of algebraic groups.

We will also need to identify the specific algebraic group Ext(E,Gm)\mathrm{Ext}(E,G_{m}), where EE is an elliptic curve. Recall that for any abelian variety AA, we can form its dual abelian variety A~\widetilde{A}, which parametrizes the topologically trivial line bundles on AA (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 Ext(E,Gm)\mathrm{Ext}(E,G_{m}) is isomorphic to the dual elliptic curve E^\widehat{E}.

Finally, if f:GmGmf:G_{m}\rightarrow G_{m} is a non-trivial morphism of algebraic groups, we will need information on the kernel of the map f:Ext(E,Gm)Ext(E,Gm)f_{*}:\mathrm{Ext}(E,G_{m})\rightarrow\mathrm{Ext}(E,G_{m}). 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 EE be an elliptic curve, and g:GmGmg:G_{m}\rightarrow G_{m} be a non-trivial morphism of algebraic groups. Then g:Ext(E,Gm)Ext(E,Gm)g_{*}:\mathrm{Ext}(E,G_{m})\rightarrow\mathrm{Ext}(E,G_{m}) has finite kernel.

Proof.

The map g:GmGmg:G_{m}\rightarrow G_{m} has finite kernel, denote it HH. By [25, Chapter 7, 4.23, Theorem 12], the functor Ext(E,)\mathrm{Ext}(E,\cdot) is exact on the category of linear algebraic groups, and we thus obtain a short exact sequence:

0Ext(E,H)Ext(E,Gm)gExt(E,Gm)00\rightarrow\mathrm{Ext}(E,H)\rightarrow\mathrm{Ext}(E,G_{m})\xrightarrow{g_{*}}\mathrm{Ext}(E,G_{m})\rightarrow 0

It is therefore enough to show that Ext(E,H)\mathrm{Ext}(E,H) is finite, whenever HH is a finite algebraic group.

To do so, we show that there is a surjective morphism from Hom(En,H)\operatorname{Hom}(\prescript{}{n}{E},H) to Ext(E,H)\mathrm{Ext}(E,H), where En\prescript{}{n}{E} is the nn-torsion of EE. This torsion group is well-known to be finite (see [18, II.4]), so Hom(En,H)\operatorname{Hom}(\prescript{}{n}{E},H) is finite as well.

Let nn\in\mathbb{N}, we have maps νn:HH\nu_{n}:H\rightarrow H and μn:EE\mu_{n}:E\rightarrow E given by sending xx to nxnx. It is well known (see [18, II.4]) that μn\mu_{n} is surjective and has finite kernel. Pick nn to be a multiple of the order of HH, so that νn\nu_{n} is the zero map, or in other words, so that nh=0nh=0 for all hHh\in H. We have a short exact sequence:

0En𝜄EμnE00\rightarrow\prescript{}{n}{E}\xrightarrow{\iota}E\xrightarrow{\mu_{n}}E\rightarrow 0

which is an element ee of Ext(E,En)\mathrm{Ext}(E,\prescript{}{n}{E}). Any ϕHom(En,G)\phi\in\operatorname{Hom}(\prescript{}{n}{E},G) gives rise to a map ϕ:Ext(E,En)Ext(E,G)\phi_{*}:\mathrm{Ext}(E,\prescript{}{n}{E})\rightarrow\mathrm{Ext}(E,G), and in particular we obtain ϕ(e)Ext(E,G)\phi_{*}(e)\in\mathrm{Ext}(E,G). To summarize, we have constructed a map:

d:Hom(En,G)\displaystyle d:\operatorname{Hom}(\prescript{}{n}{E},G) Ext(E,G)\displaystyle\rightarrow\mathrm{Ext}(E,G)
ϕ\displaystyle\phi ϕ(e)\displaystyle\rightarrow\phi_{*}(e)

By [25, Chapter 7, 1.2, Proposition 2], we obtain an exact sequence:

0{0}Hom(E,H){\operatorname{Hom}(E,H)}Hom(E,H){\operatorname{Hom}(E,H)}Hom(En,H){\operatorname{Hom}(\prescript{}{n}{E},H)}Ext(E,H){\mathrm{Ext}(E,H)}Ext(E,H){\mathrm{Ext}(E,H)}Ext(En,H){\mathrm{Ext}(\prescript{}{n}{E},H)}μn\scriptstyle{\cdot\circ\mu_{n}}ι\scriptstyle{\cdot\circ\iota}d\scriptstyle{d}μn\scriptstyle{\mu_{n}^{*}}ι\scriptstyle{\iota^{*}}

We show that dd is surjective by proving that the map μn\mu_{n}^{*} is the zero map. Let 0HC𝑓E00\rightarrow H\rightarrow C\xrightarrow{f}E\rightarrow 0 be an element of Ext(E,H)\mathrm{Ext}(E,H). We consider the group C<E×CC^{\prime}<E\times C consisting of pairs (e,c)(e,c) such that ne=f(c)ne=f(c) and obtain the extension μn(C)\mu_{n}^{*}(C) as the top row of the commutative diagram:

0{0}H{H}C{C^{\prime}}E{E}0{0}0{0}H{H}C{C}E{E}0{0}id\scriptstyle{\operatorname{id}}f\scriptstyle{f^{\prime}}G\scriptstyle{G}μn\scriptstyle{\mu_{n}}f\scriptstyle{f}

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 eEe\in E, there is a unique xnCx\in nC^{\prime} such that f(x)=ef^{\prime}(x)=e, this will define a section. We know that μn\mu_{n} and ff are surjective, so there are e~E\widetilde{e}\in E and cCc\in C such that ne~=e=f(c)n\widetilde{e}=e=f(c). Then (e~,c)C(\widetilde{e},c)\in C^{\prime} and f(n(e~,c))=ef^{\prime}(n(\widetilde{e},c))=e, giving existence. For uniqueness, suppose that f(n(e1,c1))=f(n(e2,c2))f^{\prime}(n(e_{1},c_{1}))=f^{\prime}(n(e_{2},c_{2})). This implies that ne1=ne2ne_{1}=ne_{2}, and as (e1,c1),(e2,c2)C(e_{1},c_{1}),(e_{2},c_{2})\in C^{\prime}, that f(c1)=f(c2)f(c_{1})=f(c_{2}). Hence c1c2Hc_{1}-c_{2}\in H, and by choice of nn, we obtain that nc1=nc2nc_{1}=nc_{2}. Therefore n(e1,c1)=n(e2,c2)n(e_{1},c_{1})=n(e_{2},c_{2}).

Thus we have obtained a section to ff^{\prime}, 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 𝒞\mathcal{C}-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 K,LK,L be differential subfields of 𝒰\mathcal{U}, and denote 𝒞K=𝒞K\mathcal{C}_{K}=\mathcal{C}\cap K and 𝒞L=𝒞L\mathcal{C}_{L}=\mathcal{C}\cap L. We say that K<LK<L is a strongly normal extension if:

  1. (1)

    𝒞K=𝒞L\mathcal{C}_{K}=\mathcal{C}_{L} is algebraically closed,

  2. (2)

    L/KL/K is finitely generated,

  3. (3)

    for any σAutK(𝒰)\sigma\in\mathrm{Aut}_{K}(\mathcal{U}), we have that L,𝒞=σ(L),𝒞\langle L,\mathcal{C}\rangle=\langle\sigma(L),\mathcal{C}\rangle.

The differential Galois group G(L/K)G(L/K) is the group of differential automorphisms of LL fixing KK. The full differential Galois group Gal(L/K)\mathrm{Gal}(L/K) is the group of differential automorphisms of L,𝒞\langle L,\mathcal{C}\rangle fixing K,𝒞\langle K,\mathcal{C}\rangle.

Since L/KL/K is strongly normal, there is some tuple a¯\overline{a} such that L=Ka¯L=K\langle\overline{a}\rangle. Let p=tp(a¯/K)p=\operatorname{tp}(\overline{a}/K). Marker’s proof of [16, Theorem 9.5] (and the discussion following it) shows that tp(a¯/K)\operatorname{tp}(\overline{a}/K) is 𝒞\mathcal{C}-internal, and that there is an algebraic group GG such that G(L/K)G(L/K) (resp. Gal(L/K)\mathrm{Gal}(L/K)) is isomorphic to G(𝒞K)G(\mathcal{C}_{K}) (resp. G(𝒞)G(\mathcal{C})).

It is easy to see that Gal(L/K)\mathrm{Gal}(L/K) is definably isomorphic to the binding group AutK(p/𝒞)\mathrm{Aut}_{K}(p/\mathcal{C}). In other words, the binding group of pp is isomorphic to the 𝒞\mathcal{C}-points of the Galois group of the strongly normal extension L/KL/K. 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 FF, which we can assume to be algebraically closed. We can then take the type pp obtained previously, which is weakly 𝒞\mathcal{C}-orthogonal as 𝒞K=𝒞L\mathcal{C}_{K}=\mathcal{C}_{L}. We have obtained:

Fact 4.17.

Fix a connected algebraic group GG defined over 𝒞\mathcal{C}. There exists an algebraically closed differential field FF and some pS(F)p\in S(F) such that pp is 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-orthogonal, and AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) is definably isomorphic to G(𝒞)G(\mathcal{C}).

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 FF be an algebraically closed field, and pS(F)p\in S(F) a 𝒞\mathcal{C}-internal, fundamental and weakly 𝒞\mathcal{C}-orthogonal type. Fix some apa\models p. If HH is an FF-definable normal subgroup of AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}), then there is bdcl(aF)b\in\operatorname{dcl}(aF) such that:

  • H={σAutF(p/𝒞), some (any) lift of σ to 𝒰 fixes b}H=\{\sigma\in\mathrm{Aut}_{F}(p/\mathcal{C}),\text{ some (any) lift of }\sigma\text{ to }\mathcal{U}\text{ fixes }b\},

  • tp(b/F)\operatorname{tp}(b/F) is 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-orthogonal and fundamental, and the binding group AutF(tp(b/F)/𝒞)\mathrm{Aut}_{F}(\operatorname{tp}(b/F)/\mathcal{C}) is definably isomorphic to AutF(p/𝒞)/H\mathrm{Aut}_{F}(p/\mathcal{C})/H,

  • tp(a/bF)\operatorname{tp}(a/bF) is 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C}-orthogonal and fundamental, and the binding group AutF(tp(a/bF)/𝒞)\mathrm{Aut}_{F}(\operatorname{tp}(a/bF)/\mathcal{C}) is definably isomorphic to HH.

If HH is connected, we also obtain that tp(a/bF)\operatorname{tp}(a/bF) is stationary (see [19, Chapter 1, Lemma 6.16]).

From this we obtain:

Fact 4.19.

Let FF be an algebraically closed field, let pS(F)p\in S(F) be a 𝒞\mathcal{C}-internal, fundamental and weakly 𝒞\mathcal{C}-orthogonal type. If HH is a connected FF-definable normal subgroup of AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}), then there is an FF-definable fibration f:pf(p)f:p\rightarrow f(p) with 𝒞\mathcal{C}-internal fibers, such that f(p)f(p) and pf(a)p_{f(a)} (for any f(a)f(p)f(a)\models f(p)) are 𝒞\mathcal{C}-internal and fundamental, with AutF(pf(a)/𝒞)\mathrm{Aut}_{F}(p_{f(a)}/\mathcal{C}) definably isomorphic to HH, and giving rise to the following short exact sequence:

1HAutF(p/𝒞)f~AutF(f(p)/𝒞)1 .1\rightarrow H\rightarrow\mathrm{Aut}_{F}(p/\mathcal{C})\xrightarrow{\widetilde{f}}\mathrm{Aut}_{F}(f(p)/\mathcal{C})\rightarrow 1\text{ .}

We are now equipped to prove the following:

Theorem 4.20.

There exist an algebraically closed field FF, some type pS(F)p\in S(F) that is 𝒞\mathcal{C}-internal, fundamental and weakly 𝒞\mathcal{C}-orthogonal, and a definable fibration f:pf(p)f:p\rightarrow f(p), such that:

  • AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) is definably isomorphic to D(𝒞)D(\mathcal{C}), where DD is a non-definably split extension of an elliptic curve by GmG_{m},

  • ff does not almost split.

  • for any apa\models p, the type pf(a)p_{f(a)} is 𝒞\mathcal{C}-internal, weakly 𝒞\mathcal{C} orthogonal, fundamental, with binding group definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}).

Proof.

By Fact 4.15, there is an algebraic group DD that is a non-definably split extension of an elliptic curve EE by GmG_{m}, i.e. we have a map g:DEg:D\rightarrow E with kernel GmG_{m}. Moreover, we can assume that DD does not belong to any finite subgroup of the elliptic curve Ext(E,Gm)\mathrm{Ext}(E,G_{m}) (equivalently DD is not a torsion point).

By Fact 4.17, there is an algebraically closed field FF and pS(F)p\in S(F) such that pp is 𝒞\mathcal{C}-internal and weakly 𝒞\mathcal{C}-orthogonal, with AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) definably isomorphic to D(𝒞)D(\mathcal{C}), where GG is a non-definably split extension of an elliptic curve by GmG_{m}. As was observed in [6, Fact 2.4], we may assume that pp is fundamental by taking a Morley power.

In particular the group AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) has a definable Chevalley decomposition in the sense of [6, Fact 2.8], meaning a subgroup maximal among definable subgroups FF-definably isomorphic to the 𝒞\mathcal{C}-points of a linear algebraic group. This subgroup LL is called the linear part of AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}). It is easy to show, in this case, that the linear part of AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) is the kernel of the map induced by gg on AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) (in particular it is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C})). Moreover, the linear part is FF-definable and normal by [6, Fact 2.8].

By Fact 4.19, there is an FF-definable fibration f:pf(p)f:p\rightarrow f(p) giving rise to the short exact sequence:

1LAutF(p/𝒞)f~AutF(f(p)/𝒞)11\rightarrow L\rightarrow\mathrm{Aut}_{F}(p/\mathcal{C})\xrightarrow{\widetilde{f}}\mathrm{Aut}_{F}(f(p)/\mathcal{C})\rightarrow 1

and AutF(f(p)/𝒞)\mathrm{Aut}_{F}(f(p)/\mathcal{C}) is definably isomorphic to E(𝒞)E(\mathcal{C}).

Suppose, for a contradiction, that ff does almost split. Then there exist, by Proposition 3.8, some 𝒞\mathcal{C} internal type sS(F)s\in S(F), some apa\models p and bsb\models s such that aa is interalgebraic with (f(a),b)(f(a),b) over FF, with b|Ff(a)b\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{F}f(a).

Since both pp and f(p)f(p) are fundamental, by [6, Lemma 2.7] Autf(a)F(pf(a)/𝒞)\mathrm{Aut}_{f(a)F}(p_{f(a)}/\mathcal{C}) is definably isomorphic to LL, and hence to Gm(𝒞)G_{m}(\mathcal{C}). By Lemma 2.10 the binding groups Autf(a)F(pf(a)/𝒞)\mathrm{Aut}_{f(a)F}(p_{f(a)}/\mathcal{C}) and Autf(a)F(tp(b/f(a)F)/𝒞)\mathrm{Aut}_{f(a)F}(\operatorname{tp}(b/f(a)F)/\mathcal{C}) are isogenous, thus Autf(a)𝒞(tp(b/f(a)F)/𝒞)\mathrm{Aut}_{f(a)\mathcal{C}}(\operatorname{tp}(b/f(a)F)/\mathcal{C}) is isogenous to Gm(𝒞)G_{m}(\mathcal{C}). Also Autf(a)F(tp(b/f(a)F)/𝒞)\mathrm{Aut}_{f(a)F}(\operatorname{tp}(b/f(a)F)/\mathcal{C}) is a definable subgroup of AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) by Fact 2.5.

As LL acts regularly on pf(a)p_{f(a)}, we see that U(a/f(a)F)=1U(a/f(a)F)=1. We deduce from this that U(s)=1U(s)=1. As ss is 𝒞\mathcal{C}-internal, this implies that RM(s)=1\mathrm{RM}(s)=1, and as FF is algebraically closed, the type ss is strongly minimal.

Fact 4.3 rules out the cases of the additive group and an elliptic curve for AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}). By Lemma 2.10, the groups AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) and AutF(sf(p)/𝒞)\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C}) are isogenous. The projection map sf(p)ss\otimes f(p)\rightarrow s yields a definable surjective morphism AutF(sf(p)/𝒞)AutF(s/𝒞)\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C})\rightarrow\mathrm{Aut}_{F}(s/\mathcal{C}). As the former is isogenous to a semi-abelian variety, it cannot have any definable morphism to GaGmG_{a}\ltimes G_{m} nor PSL2\mathrm{PSL_{2}}, ruling out the other two cases. Thus AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}), and by connectedness, we get Autf(a)F(tp(b/f(a)F)/𝒞)=AutF(s/𝒞)\mathrm{Aut}_{f(a)F}(\operatorname{tp}(b/f(a)F)/\mathcal{C})=\mathrm{Aut}_{F}(s/\mathcal{C}).

So AutF(f(p)/𝒞)\mathrm{Aut}_{F}(f(p)/\mathcal{C}) is definably isomorphic to E(𝒞)E(\mathcal{C}) and AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}) is definably isomorphic to Gm(𝒞)G_{m}(\mathcal{C}). Corollary 2.11 implies that AutF(s/𝒞,f(p))=AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C},f(p))=\mathrm{Aut}_{F}(s/\mathcal{C}), from which we deduce:

AutF(f(p)s/𝒞,f(p))\displaystyle\mathrm{Aut}_{F}(f(p)\otimes s/\mathcal{C},f(p)) =AutF(s/𝒞,f(p))\displaystyle=\mathrm{Aut}_{F}(s/\mathcal{C},f(p))
=AutF(s/𝒞)\displaystyle=\mathrm{Aut}_{F}(s/\mathcal{C})

As f(p)f(p) is weakly orthogonal to 𝒞\mathcal{C}, Corollary 2.11 also implies that f(p)f(p) is weakly orthogonal to {s,𝒞}\{s,\mathcal{C}\}. Therefore the short exact sequence:

1AutF(s/𝒞)AutF(f(p)s/𝒞)AutF(f(p)/𝒞)11\rightarrow\mathrm{Aut}_{F}(s/\mathcal{C})\rightarrow\mathrm{Aut}_{F}(f(p)\otimes s/\mathcal{C})\rightarrow\mathrm{Aut}_{F}(f(p)/\mathcal{C})\rightarrow 1

is FF-definably split by Lemmas 3.14 and 3.15.

By Lemma 2.10, the groups AutF(sf(p)/𝒞)\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C}) and AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) are isogenous, let GG be the FF-definable group witnessing it. The kernels of the maps to GG are given by elements fixing pp (resp. f(p)sf(p)\otimes s), and therefore must in particular be contained in the kernels of the maps to AutF(f(p)/𝒞)\mathrm{Aut}_{F}(f(p)/\mathcal{C}), i.e. LL and AutF(s/𝒞)\mathrm{Aut}_{F}(s/\mathcal{C}). Therefore the two groups are also isogenous, and we have an FF-definable group HH and a commutative diagram of FF-definable maps:

1{1}L{L}AutF(p/𝒞){\mathrm{Aut}_{F}(p/\mathcal{C})}AutF(f(p)/𝒞){\mathrm{Aut}_{F}(f(p)/\mathcal{C})}1{1}1{1}H{H}G{G}AutF(f(p)/𝒞){\mathrm{Aut}_{F}(f(p)/\mathcal{C})}1{1}1{1}AutF(s/𝒞){\mathrm{Aut}_{F}(s/\mathcal{C})}AutF(sf(p)/𝒞){\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C})}AutF(f(p)/𝒞){\mathrm{Aut}_{F}(f(p)/\mathcal{C})}1{1}f~\scriptstyle{\widetilde{f}}

Both AutF(p/𝒞)\mathrm{Aut}_{F}(p/\mathcal{C}) and AutF(f(p)s/𝒞)\mathrm{Aut}_{F}(f(p)\otimes s/\mathcal{C}) are abelian, and therefore FF-definably isomorphic to the 𝒞\mathcal{C}-points of algebraic groups (see [6, Lemma 2.2] for a proof of that well-known fact). Denote the 𝒞\mathcal{C}-points of these algebraic groups by AutF(p/𝒞)¯\overline{\mathrm{Aut}_{F}(p/\mathcal{C})} and AutF(f(p)s/𝒞)¯\overline{\mathrm{Aut}_{F}(f(p)\otimes s/\mathcal{C})}. They are defined over F𝒞F\cap\mathcal{C}, are isogenous, and by the same reasoning as in the previous paragraph, we obtain an F𝒞F\cap\mathcal{C}-definable algebraic group G¯\overline{G} and a commutative diagram:

1{1}Gm(𝒞){G_{m}(\mathcal{C})}AutF(p/𝒞)¯{\overline{\mathrm{Aut}_{F}(p/\mathcal{C})}}E(𝒞){E(\mathcal{C})}1{1}1{1}Gm(𝒞){G_{m}(\mathcal{C})}G¯{\overline{G}}E(𝒞){E(\mathcal{C})}1{1}1{1}Gm(𝒞){G_{m}(\mathcal{C})}AutF(sf(p)/𝒞)¯{\overline{\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C})}}E(𝒞){E(\mathcal{C})}1{1}ι\scriptstyle{\iota}ζ\scriptstyle{\zeta}

where everything is defined over 𝒞F\mathcal{C}\cap F.

By Fact 4.14, the maps ι\iota and ζ\zeta give rise to two automorphisms ι\iota_{*} and ζ\zeta_{*} of Ext(E,Gm)\mathrm{Ext}(E,G_{m}). We obtain that ι(AutF(p/𝒞)¯)=G¯=ζ(AutF(sf(p)/𝒞)¯)\iota_{*}(\overline{\mathrm{Aut}_{F}(p/\mathcal{C})})=\overline{G}=\zeta_{*}(\overline{\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C})}).

We know that AutF(sf(p)/𝒞)¯\overline{\mathrm{Aut}_{F}(s\otimes f(p)/\mathcal{C})} is a 𝒞F\mathcal{C}\cap F-definably split extension of E(𝒞)E(\mathcal{C}), therefore it is trivial as an element of Ext(E,Gm)\mathrm{Ext}(E,G_{m}), and so is its image G¯\overline{G}, hence G¯\overline{G} is the identity of Ext(E,Gm)\mathrm{Ext}(E,G_{m}), and thus the extension AutF(p/𝒞)¯\overline{\mathrm{Aut}_{F}(p/\mathcal{C})} belongs to the kernel of ι\iota_{*}. By Proposition 4.16, the map ι\iota_{*} has finite kernel. Therefore the extension AutF(p/𝒞)¯\overline{\mathrm{Aut}_{F}(p/\mathcal{C})}, which is isomorphic to the extension DD, belongs to a finite subgroup of Ext(E,Gm)\mathrm{Ext}(E,G_{m}). This contradicts our choice of DD, as it was assumed that it did not belong to any finite subgroup of Ext(Gm,E)=E^\mathrm{Ext}(G_{m},E)=\widehat{E}. ∎

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 F<𝒰F<\mathcal{U}, a type qS1(F)q\in S_{1}(F) such that logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q) does not almost-split, the type logδ(q)\log_{\delta}(q) is strongly minimal and has a binding group isomorphic to the constant points of an elliptic curve EE.

Proof.

Applying Lemma 4.8 to the algebraically closed field FF and pS(F)p\in S(F) obtained in Theorem 4.20, we obtain a type qS1(F)q\in S_{1}(F), internal to 𝒞\mathcal{C}, a fibration logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q) with logδ(q)dcl(f(p)(𝒰)F)\log_{\delta}(q)\subset\operatorname{dcl}(f(p)(\mathcal{U})F). Because ff does not almost split, the map logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q) does not almost split either. To conclude, we just need to show that logδ(q)\log_{\delta}(q) is strongly minimal and AutF(logδ(q)/𝒞)\mathrm{Aut}_{F}(\log_{\delta}(q)/\mathcal{C}) definably isomorphic to the constant points of an elliptic curve.

Because logδ(q)dcl(f(p)(𝒰)F)\log_{\delta}(q)\subset\operatorname{dcl}(f(p)(\mathcal{U})F), we have that AutF(logδ(q)/f(p),𝒞)={id}\mathrm{Aut}_{F}(\log_{\delta}(q)/f(p),\mathcal{C})=\{\operatorname{id}\}, therefore by Lemma 2.10 we obtain an FF-definable surjective map AutF(f(p)/𝒞)AutF(logδ(q)/𝒞)\mathrm{Aut}_{F}(f(p)/\mathcal{C})\rightarrow\mathrm{Aut}_{F}(\log_{\delta}(q)/\mathcal{C}). As AutF(f(p)/𝒞)\mathrm{Aut}_{F}(f(p)/\mathcal{C}) is definably isomorphic to E(𝒞)E(\mathcal{C}), and thus strongly minimal, its kernel is either the whole group, which would imply that logδ(q)\log_{\delta}(q) is 𝒞\mathcal{C}-definable (i.e. logδ(q)dcl(𝒞,F)\log_{\delta}(q)\subset\operatorname{dcl}(\mathcal{C},F)), or finite.

Note that the type logδ(q)\log_{\delta}(q) cannot be algebraic, as if it was, then logδ:qlogδ(q)\log_{\delta}:q\rightarrow\log_{\delta}(q) would be almost split and thus so would ff. Hence in the first possibility, we deduce that logδ(q)\log_{\delta}(q) is not weakly orthogonal to 𝒞\mathcal{C}. But any logδ(b)logδ(q)\log_{\delta}(b)\models\log_{\delta}(q) is in the algebraic closure of some apa\models p, which contradicts pp being weakly orthogonal to 𝒞\mathcal{C}.

Therefore the kernel of the map is finite, and AutF(logδ(q)/𝒞)\mathrm{Aut}_{F}(\log_{\delta}(q)/\mathcal{C}) is definably isomorphic to an algebraic group isogenous to E(𝒞)E(\mathcal{C}), which thus must be an elliptic curve. Note that by the proof of Lemma 4.8, for any logδ(b)q\log_{\delta}(b)\models q, there is f(a)f(p)f(a)\models f(p) such that logδ(b)dcl(f(a),F)\log_{\delta}(b)\in\operatorname{dcl}(f(a),F). This, and the previous discussion, forces logδ(q)\log_{\delta}(q) to be strongly minimal. ∎

It is unfortunate that our method does not yield a specific differential equation with 𝒞\mathcal{C}-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.