Hostname: page-component-54dcc4c588-54gsr Total loading time: 0 Render date: 2025-09-15T15:50:21.486Z Has data issue: false hasContentIssue false

MODEL THEORY OF DIFFERENTIAL-HENSELIAN PRE-H-FIELDS

Published online by Cambridge University Press:  08 April 2025

NIGEL PYNN-COATES*
Affiliation:
KURT GÖDEL RESEARCH CENTER INSTITUTE OF MATHEMATICS UNIVERSITY OF VIENNAWIENAUSTRIA
Rights & Permissions [Opens in a new window]

Abstract

Pre-H-fields are ordered valued differential fields satisfying some basic axioms coming from transseries and Hardy fields. We study pre-H-fields that are differential-Hensel–Liouville closed, that is, differential-henselian, real closed, and closed under exponential integration, establishing an Ax–Kochen/Ershov theorem for such structures: the theory of a differential-Hensel–Liouville closed pre-H-field is determined by the theory of its ordered differential residue field; this result fails if the assumption of closure under exponential integration is dropped. In a two-sorted setting with one sort for a differential-Hensel–Liouville closed pre-H-field and one sort for its ordered differential residue field, we eliminate quantifiers from the pre-H-field sort, from which we deduce that the ordered differential residue field is purely stably embedded and if it has NIP, then so does the two-sorted structure. Similarly, the one-sorted theory of differential-Hensel–Liouville closed pre-H-fields with closed ordered differential residue field has quantifier elimination, is the model completion of the theory of pre-H-fields with gap $0$, and is complete, distal, and locally o-minimal.

Information

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of The Association for Symbolic Logic

1 Introduction

In [Reference Aschenbrenner, van den Dries and van der Hoeven5], Aschenbrenner, van den Dries, and van der Hoeven study the model theory of $\mathbb {T}$ as an ordered valued differential field, where $\mathbb {T}$ is the differential field of logarithmic-exponential transseries constructed in [Reference van den Dries, Macintyre and Marker23], there denoted by $\mathbb {R}(\!( t )\!)^{\operatorname {LE}}$ . As part of this project, Aschenbrenner and van den Dries introduced in [Reference Aschenbrenner and van den Dries4] the elementary class of pre-H-fields, which are ordered valued differential fields satisfying conditions capturing some of the basic interactions between the ordering, valuation, and derivation in transseries and Hardy fields. As such, the model theory of transseries leads naturally to the model theory of pre-H-fields. Among the many results of [Reference Aschenbrenner, van den Dries and van der Hoeven5], for which the three received the 2018 Karp Prize from the Association for Symbolic Logic, is an effective axiomatization of the theory of $\mathbb {T}$ as a complete theory extending the theory of pre-H-fields.

But there are pre-H-fields satisfying different elementary conditions than $\mathbb {T}$ , including $\operatorname {d}$ -henselian ones (we use “ $\operatorname {d}$ ” to abbreviate “differential” or “differentially,” as appropriate), which are the focus of this paper. This is a generalization of henselianity of a valued field to the class of valued differential fields with small derivation, a strong form of continuity of the derivation with respect to the valuation topology. To describe a family of completions of the theory of pre-H-fields, we call a pre-H-field $\operatorname {d}$ -Hensel–Liouville closed if it is $\operatorname {d}$ -henselian, real closed, and closed under exponential integration (i.e., for each f there is $y \neq 0$ such that $y'/y = f$ ). Given a complete theory T of ordered differential fields, the theory of $\operatorname {d}$ -Hensel–Liouville closed pre-H-fields whose ordered differential residue fields satisfy T is complete. Equivalently:

Theorem 1. Let $K_1$ and $K_2$ be $\operatorname {d}$ -Hensel–Liouville closed pre-H-fields with ordered differential residue fields $\boldsymbol k_1$ and $\boldsymbol k_2$ . Then $K_1 \equiv K_2$ if and only if $\boldsymbol k_1 \equiv \boldsymbol k_2$ .

More precisely, the pre-H-field language is , where, if $\mathcal O$ is the valuation ring, $f \preccurlyeq g$ means $f \in g\mathcal O$ . Each residue field is equipped with the ordering and derivation induced by those of its pre-H-field and construed as a structure in the language .

Theorem 1 is a result in the spirit of Ax–Kochen [Reference Ax and Kochen7] and Ershov [Reference Ershov11], who showed that the theory of a henselian valued field of equicharacteristic $0$ is determined by the theory of its residue field and (ordered) value group. The value groups of $K_1$ and $K_2$ do not appear in Theorem 1 because under its assumptions they are divisible ordered abelian groups, and hence elementarily equivalent. Moreover, they are even elementarily equivalent as asymptotic couples, as defined by Rosenlicht [Reference Rosenlicht19], which is to say after being expanded by the map induced by logarithmic differentiation (see Section 2). In addition to the axioms for divisible ordered abelian groups, the complete axiomatization is given by four universal axioms of asymptotic couples together with the statement that the logarithmic derivative map is surjective onto the set of negative elements of the group; this is Corollary 4.9 (where these structures are called “gap-closed H-asymptotic couples”).

Analogous AKE results for (unordered) valued differential fields have been established earlier for $\operatorname {d}$ -henselian monotone fields: first for those with many constants by Scanlon [Reference Scanlon21] and then in the general monotone case by Hakobyan [Reference Hakobyan12]. The interaction between the valuation and the derivation in a pre-H-field is opposite to the condition of monotonicity, and, indeed, $\operatorname {d}$ -henselian pre-H-fields are never monotone.

To which pre-H-fields does Theorem 1 apply? Certainly not $\mathbb {T}$ , which is not $\operatorname {d}$ -henselian.Footnote 1 To construct an example, start with an $\aleph _0$ -saturated elementary extension $\mathbb {T}^{*}$ of $\mathbb {T}$ . Although every element of $\mathbb {T}$ is exponentially bounded in the sense that it is bounded in absolute value by some finite iterate of the exponential, $\mathbb {T}^{*}$ contains transexponential (i.e., not exponentially bounded) elements. Enlarge the valuation ring $\mathcal O_{\mathbb {T}^{*}}$ of $\mathbb {T}^{*}$ to the set $\dot {\mathcal O}_{\mathbb {T}^{*}}$ of exponentially bounded elements of $\mathbb {T}^{*}$ . With this coarsened valuation, $(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ is a $\operatorname {d}$ -Hensel–Liouville closed pre-H-field whose valuation only distinguishes transexponentially different elements, and its ordered differential residue field $\operatorname {\mathrm {res}}(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ is a model of the theory of $\mathbb {T}$ . This coarsening decomposes $\mathbb {T}^{*}$ into a transexponential part, $(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ , whose model theory is the subject of this paper, and an exponentially bounded part, $\operatorname {\mathrm {res}}(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ , whose model theory we understand by [Reference Aschenbrenner, van den Dries and van der Hoeven5]. The same coarsening procedure applied to a maximal Hardy field, which contains transexponential elements [Reference Boshernitzan9], yields a similar example.

Theorem 1 follows easily from Theorem 7.2, which establishes a back-and-forth system in a two-sorted setting with a sort for a pre-H-field and a sort for its ordered differential residue field, connected by a binary residue map. This back-and-forth system allows us to eliminate quantifiers from the pre-H-field sort in Theorem 7.5, which is used to show the following two results. (In fact, we get more precise two-sorted versions.)

Theorem 2. Let K be a $\operatorname {d}$ -Hensel–Liouville closed pre-H-field with ordered differential residue field $\boldsymbol k$ .

  1. (i) Every subset of $\boldsymbol k^n$ definable in K, with parameters, is definable in the ordered differential field $\boldsymbol k$ with parameters from $\boldsymbol k$ .

  2. (ii) If the ordered differential field $\boldsymbol k$ has NIP, then K has NIP.

Another of the central results of [Reference Aschenbrenner, van den Dries and van der Hoeven5] is that the class of existentially closed pre-H-fields is elementary and axiomatized by a theory $T^{\operatorname {nl}}$ ; note that $T^{\operatorname {nl}}_{\operatorname {small}} = T^{\operatorname {nl}} + \text {"small derivation"}$ axiomatizes the theory of $\mathbb {T}$ . Equivalently, $T^{\operatorname {nl}}$ is the model companion of the theory of pre-H-fields, that is, $T^{\operatorname {nl}}$ is model complete and every pre-H-field extends to a model of $T^{\operatorname {nl}}$ . Also, $T^{\operatorname {nl}}_{\operatorname {small}}$ is the model companion of the theory of H-fields with small derivation, where an H-field is a pre-H-field whose valuation ring is the convex hull of its constant field. But a pre-H-field with small derivation extends to a model of $T^{\operatorname {nl}}_{\operatorname {small}}$ if and only if the derivation induced on its residue field is trivial. This raises the question of whether the theory of pre-H-fields with nontrivial induced derivation on their residue field, which includes all $\operatorname {d}$ -henselian pre-H-fields, has a model companion.

We give a positive answer for the broader class of pre-H-fields with gap $0$ . In defining this, we use the notation $f \asymp g$ when $f \preccurlyeq g$ and $g \preccurlyeq f$ , and $f \prec g$ when $f \preccurlyeq g$ and $f \not \asymp g$ . A pre-H-field has small derivation if $f' \prec 1$ for all $f \prec 1$ and gap $0$ if it has small derivation and every $f \succ 1$ satisfies $f' \succ f$ . If a pre-H-field with gap $0$ is closed under exponential integration (or even if its ordered differential residue field is), then any $f \succ 1$ is “transexponential” in the sense that $f \succ e_n$ for each n, where $e_n$ is an nth iterated exponential integral of $1$ . To describe the axiomatization of the model completion, we recall that the theory of ordered differential fields, with no assumption on the interaction between the ordering and the derivation, has a model completion, the theory of closed ordered differential fields. This theory also has quantifier elimination and is complete [Reference Singer22]. Then:

Theorem 3. The theory $T^{\operatorname {dhl}}_{\operatorname {codf}}$ of $\operatorname {d}$ -Hensel–Liouville closed pre-H-fields with closed ordered differential residue field is the model completion of the theory of pre-H-fields with gap $0$ .

In fact, we prove a more general two-sorted version where other theories of the residue field are permitted, as in the previous two results. By Theorem 1, $T^{\operatorname {dhl}}_{\operatorname {codf}}$ is complete. This theory has quantifier elimination, so it is distal by [Reference Aschenbrenner, Chernikov, Gehret and Ziegler2] and locally o-minimal. The one-sorted results are collected in Theorem 7.9.

The assumptions of “real closed” and “exponential integration” are necessary in Theorem 3, but what of their appearance in Theorem 1? Even in the monotone setting, it is not true that the theory of a $\operatorname {d}$ -henselian field is determined by the theory of its differential residue field and its value group. Hakobyan provides an example of two $\operatorname {d}$ -henselian monotone fields that are not elementarily equivalent but have isomorphic value groups and differential residue fields [Reference Hakobyan12, Example after Corollary 4.2]. To remedy this, in his monotone AKE theorem there is an additive map from the value group to the residue field. Hence any AKE theorem for $\operatorname {d}$ -henselian fields requires either additional assumptions or extra structure on the value group or differential residue field. While it may be possible to remove the assumption that $K_1$ and $K_2$ are real closed in Theorem 1 at the expense of incorporating their asymptotic couples, the next result shows that closure under exponential integration is necessary.

Theorem 4. There exist $\operatorname {d}$ -henselian, real closed pre-H-fields $K_1$ and $K_2$ with ordered differential residue fields $\boldsymbol k_1$ and $\boldsymbol k_2$ such that $\boldsymbol k_1 \cong \boldsymbol k_2$ but $K_1 \not \equiv K_2$ .

More precisely, one of $K_1$ and $K_2$ is closed under exponential integration but the other is not. Also, their asymptotic couples are elementarily equivalent, so any three-sorted improvement of Theorem 1 would still require some assumption on the behaviour of logarithmic derivatives.

1.1 Structure of the paper

Section 2 contains basic definitions, notation, and remarks, which we keep close to [Reference Aschenbrenner, van den Dries and van der Hoeven5]. With the goal of back-and-forth arguments in mind, in subsequent sections we consider extensions controlled by the residue field, extensions controlled by the asymptotic couple, and those that involve adjoining exponential integrals.

In Section 3, we show how to extend embeddings of ordered valued differential fields by first extending the residue field.

We study extensions controlled by the asymptotic couple in Section 4, starting by studying asymptotic couples as structures in their own right in Section 4.1. We isolate the model completion of the theory of H-asymptotic couples with gap $0$ and prove that this theory has quantifier elimination in Theorem 4.8. In the rest of the section, this result, or rather its consequence Corollary 4.10, is used to study extensions of $\operatorname {d}$ -henselian pre-H-fields whose asymptotic couples are existentially closed.

The short Section 5 deals with extending the constant field and plays no role in the main results; it is used only to strengthen the statement of Theorem 6.15 in the next section.

Section 6 proves the most difficult embedding result of the paper, Theorem 6.16: the existence of $\operatorname {d}$ -Hensel–Liouville closures, which are extensions that are $\operatorname {d}$ -henselian, real closed, and closed under exponential integration, and that satisfy a semi-universal property. Uniqueness of $\operatorname {d}$ -Hensel–Liouville closures is Corollary 6.18.

Turning to the main results, in Section 7, first some embedding results from previous sections are combined in the key embedding lemma. Next we establish the two-sorted back-and-forth system, in Theorem 7.2. The AKE theorem, Theorem 1 (Corollary 7.3), follows immediately. The relative quantifier elimination is Theorem 7.5, from which the stable embeddedness of the residue field (the first part of Theorem 2) follows immediately. The two-sorted model companion result is Corollary 7.7, and the NIP transfer principle (the second part of Theorem 2) is Proposition 7.8. One-sorted results for $T^{\operatorname {dhl}}_{\operatorname {codf}}$ , including Theorem 3 but also quantifier elimination, distality, and local o-minimality, are collected in Theorem 7.9.

Section 8 documents two examples of $\operatorname {d}$ -henselian pre-H-fields. The first, in Section 8.1, provides a fuller account of the example described in the introduction of a $\operatorname {d}$ -Hensel–Liouville closed pre-H-field whose residue field is a model of $T^{\operatorname {nl}}_{\operatorname {small}}$ , which arises from a transexponential extension of $\mathbb {T}$ . This example motivates the earlier two-sorted results, in which the ordered differential residue field could have additional structure. The second, in Section 8.2, establishes Theorem 4 (Corollary 8.6), showing that the assumption of closure under exponential integration in the AKE theorem cannot be dropped. As part of the proof, we show that for every ordered differential field $\boldsymbol k$ satisfying three obviously necessary conditions, there is a $\operatorname {d}$ -Hensel–Liouville closed pre-H-field with ordered differential residue field isomorphic to $\boldsymbol k$ .

2 Preliminaries

We let m, n, and r range over $\mathbb {N} = \{0, 1, 2, \dots \}$ and $\rho $ , $\lambda $ , and $\mu $ be ordinals. The main objects of this paper are kinds of ordered valued differential fields; all fields in this paper are assumed to be of characteristic $0$ . A valued field is a field K equipped with a surjective map $v \colon K^{\times } \to \Gamma $ , where $\Gamma $ is a (totally) ordered abelian group, satisfying for $f, g \in K^{\times }$ :

  1. (V1) $v(fg) = v(f)+v(g)$ ;

  2. (V2) $v(f+g) \geqslant \min \{v(f), v(g)\}$ whenever $f+g \neq 0$ .

We also impose the condition that $v(\mathbb {Q}^{\times })=\{0\}$ , i.e., that K has equicharacteristic $0$ . A differential field is a field K equipped with a derivation , which satisfies for $f, g \in K$ :

  1. (D1) ;

  2. (D2) .

Let K be a valued field. We add a new symbol $\infty $ to $\Gamma $ and extend the addition and ordering to by $\infty +\gamma =\gamma +\infty =\infty $ and $\infty>\gamma $ for all $\gamma \in \Gamma $ . This allows us to extend v to K by setting . We often use the following more intuitive notation:

$$\begin{align*}\begin{array}{lc} f \preccurlyeq g\ \Leftrightarrow\ v(f)\geqslant v(g),\qquad f \prec g\ \Leftrightarrow\ v(f)>v(g),\\ f \asymp g\ \Leftrightarrow\ v(f)=v(g),\qquad f\sim g\ \Leftrightarrow\ f-g \prec g. \end{array}\end{align*}$$

The relation $\preccurlyeq $ is called a dominance relation. Both $\asymp $ and $\sim $ are equivalence relations on K and $K^{\times }$ respectively, with a consequence of (V2) being that if $f \sim g$ , then $f \asymp g$ . We set and call it the valuation ring of K. It has a (unique) maximal ideal , and we call the residue field of K, usually denoted by $\boldsymbol k$ . We also let $\overline {a}$ or $\operatorname {\mathrm {res}}(a)$ denote the image of $a \in \mathcal O$ under the natural map to $\boldsymbol k$ . For another valued field L, we denote these objects by $\mathcal O_L$ , $\Gamma _L$ , $\boldsymbol k_L$ , etc.

Let K be a differential field. For $f \in K$ , we often write $f'$ for if the derivation is clear from context and set if $f \neq 0$ , the logarithmic derivative of f. We say that K has exponential integration if $(K^{\times })^{\dagger } = K$ . The field of constants of K is . For another differential field L, we denote this object by $C_L$ . We let be the ring of differential polynomials over K and set . For $P \in K\{Y\}^{\neq }$ , the order of P is the smallest r such that $P \in K[Y, Y', \dots , Y^{(r)}]$ . We extend the derivation of K to $K\{Y\}$ in the natural way. If L is a differential field extension of K and $a \in L$ , then $K \langle a \rangle $ denotes the differential subfield of L generated by a over K. If K is additionally a valued field, we extend v to $K\{Y\}$ by setting $v(P)$ to be the minimum valuation of the coefficients of P and thus also extend the relations $\preccurlyeq $ , $\prec $ , $\asymp $ , and $\sim $ to $K\{Y\}$ .

Assumption. Suppose for the rest of the paper that K is (at least) a valued differential field, unless stated otherwise.

Relating the valuation and the derivation, we impose throughout most of this paper the condition that K has small derivation, which means that . If K has small derivation, [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 4.4.2], so induces a derivation on $\boldsymbol k$ , and we always construe $\boldsymbol k$ as a differential field with this induced derivation. In contrast to the main results of [Reference Aschenbrenner, van den Dries and van der Hoeven5], we are interested in the case that it is nontrivial. We call $\boldsymbol k$ linearly surjective if for all $a_0, \dots , a_r \in \boldsymbol k$ with $a_r \neq 0$ , the equation $1 + a_0y + a_1y' + \dots + a_r y^{(r)} = 0$ has a solution in $\boldsymbol k$ . We call K differential-henselian ( $\operatorname {d}$ -henselian for short) if K has small derivation and:

  1. (DH1) $\boldsymbol k$ is linearly surjective;

  2. (DH2) whenever $P \in \mathcal O\{Y\}$ of order r satisfies

    $$\begin{align*}P(0)\ \prec\ 1\qquad \text{and} \qquad \sum_{n=0}^r \frac{\partial P}{\partial Y^{(n)}}(0)Y^{(n)}\ \asymp\ 1,\end{align*}$$
    there is $y \prec 1$ in K with $P(y) = 0$ .

Differential-henselianity, a generalization of henselianity to valued differential fields with small derivation, was introduced in [Reference Scanlon21] and studied more systematically in [Reference Aschenbrenner, van den Dries and van der Hoeven5].

Here is another relation between the valuation and derivation fundamental to this setting. We call K asymptotic if $f \prec g \iff f' \prec g'$ for all nonzero . In the rest of this paragraph, suppose that K is asymptotic. Note that $C \subseteq \mathcal O$ . Also, if $f, g \in K^{\times }$ satisfy $f \sim g \not \asymp 1$ , then $f' \sim g'$ and $f^{\dagger } \sim g^{\dagger }$ . Similarly, for $g \in K^{\times }$ with $g \not \asymp 1$ , $v(g^{\dagger })$ and $v(g')$ depend only on $vg$ and not on g, so for $\gamma = vg$ we set and ; note that $\gamma ^{\dagger }=\gamma '-\gamma $ . For any ordered abelian group G, set , , and . Thus logarithmic differentiation induces a map

$$ \begin{align*} \psi \colon \Gamma^{\neq} &\to \Gamma\\ \gamma &\mapsto \gamma^{\dagger}. \end{align*} $$

We call $(\Gamma , \psi )$ the asymptotic couple of K; such structures were introduced by Rosenlicht [Reference Rosenlicht19], and more about them can be found in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Sections 6.5 and 9.2]. We set , and if we need to indicate the dependence on K, we denote this by $\Psi _K$ . When convenient, we extend $\psi $ to a map $\psi \colon \Gamma _{\infty } \to \Gamma _{\infty }$ by setting .

Here are two additional properties an asymptotic K can have. We say that K is H-asymptotic or of H-type if, for all $f, g \in K^{\times }$ satisfying $f \preccurlyeq g \prec 1$ , we have $f^{\dagger } \succcurlyeq g^{\dagger }$ . We say that K has gap $0$ if it has small derivation and $f^{\dagger } \succ 1$ for all $f \in K^{\times }$ with $f \prec 1$ . Properties of K are reflected in its asymptotic couple, but conversely some properties of K are determined by $(\Gamma , \psi )$ . For instance, K has small derivation if and only if $(\Gamma ^>)' \subseteq \Gamma ^>$ , and K has gap $0$ if and only if $(\Gamma ^>)' \subseteq \Gamma ^>$ and $\Psi \subseteq \Gamma ^<$ . There can be at most one $\beta \in \Gamma $ with $\Psi < \beta < (\Gamma ^>)'$ [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 9.2.4], so K having gap $0$ means that $\beta =0$ separates $\Psi $ and $(\Gamma ^>)'$ in this way. Similarly, K is of H-type just in case $\psi (\alpha )\leqslant \psi (\beta )$ whenever $\alpha \leqslant \beta <0$ in $\Gamma $ . For more on asymptotic fields, see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Chapter 9].

By the above, all asymptotic fields with gap $0$ are clearly pre- $\operatorname {d}$ -valued, where we say that K is pre-differential-valued (pre- $\operatorname {d}$ -valued for short) if:

  1. (PDV) for all $f, g \in K^{\times }$ with $f \preccurlyeq 1$ and $g \prec 1$ , we have $f' \prec g^{\dagger }$ .

More on pre- $\operatorname {d}$ -valued fields can be found in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 10.1], including a characterization of them as those K that satisfy $C \subseteq \mathcal O$ and a valuation-theoretic analog of l’Hôpital’s Rule [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.1.4]. In the rest of this paragraph, let K be pre- $\operatorname {d}$ -valued. If K has small derivation and the derivation induced on $\boldsymbol k$ is nontrivial (for instance, if K is $\operatorname {d}$ -henselian), then K has gap $0$ . Having gap $0$ is inherited by valued differential subfields, and thus if K has a $\operatorname {d}$ -henselian valued differential field extension, then K has gap $0$ . These remarks apply in particular to pre-H-fields with gap $0$ , which we now define.

This paper is primarily concerned with certain ordered pre- $\operatorname {d}$ -valued fields called pre-H-fields. Here, K is an ordered valued differential field if, in addition to its valuation and derivation, K is equipped with a (total) ordering $\leqslant $ making it an ordered field. Relating the ordering, valuation, and derivation, we call K a pre- H-field if:

  1. (PH1) K is pre- $\operatorname {d}$ -valued;

  2. (PH2) $\mathcal O$ is convex (with respect to $\leqslant $ );

  3. (PH3) for all $f \in K$ , if $f> \mathcal O$ , then $f'>0$ .

Condition (PH2) holds if and only if is convex, which holds if and only if . Hence if (PH2) holds, then $\leqslant $ induces an ordering on $\boldsymbol k$ making it an ordered field. We thus construe the residue fields of pre-H-fields with small derivation as ordered differential fields. By [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.5.2(ii)], pre-H-fields are of H-type.

We now discuss extensions of valued differential fields and ordered valued differential fields. Given an extension L of K, we identify $\Gamma $ with an ordered subgroup of $\Gamma _L$ and $\boldsymbol k$ with a subfield of $\boldsymbol k_L$ in the obvious way. Here and in general we use the word extension as follows: if F is a valued differential field, “extension of F” means “valued differential field extension of F”; if F is an ordered valued differential field, “extension of F” means “ordered valued differential field extension of F”; etc. “Embedding,” “isomorphic,” and “isomorphism” are used similarly. Where there is particular danger of confusion, we are explicit.

An important class of extensions are the immediate extensions: We say that an extension L of K is immediate if $\Gamma _L = \Gamma $ and $\boldsymbol k_L = \boldsymbol k$ ; equivalently, for every $b \in L^{\times }$ there is $a \in K^{\times }$ such that $b \sim a$ . If K is a pre-H-field and L is an immediate valued differential field extension of K that is asymptotic, then L can be given an ordering making it a pre-H-field extension of K; in fact, this is the unique ordering with respect to which $\mathcal O_L$ is convex, and hence any valued differential field embedding of L into a pre-H-field extension M of K is automatically an ordered valued differential field embedding [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.5.8].

3 Extensions controlled by the residue field

In this section we prove Lemma 3.1, which yields ordered variants of results from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Sections 6.3 and 7.1] for use in Section 7. In the first lemma, K need not be equipped with a derivation.

Lemma 3.1. Suppose that K is an ordered valued field with convex valuation ring. Let L be a valued field extension of K with $\Gamma _L=\Gamma $ and suppose that $\boldsymbol k_L$ is an ordered field extension of $\boldsymbol k$ . Then there exists a unique ordering on L making it an ordered field extension of K with convex valuation ring such that the induced ordering on $\boldsymbol k_L$ agrees with the given one.

If K is a pre-H-field with gap $0$ and L is moreover a differential field extension of K with small derivation, then L is also a pre-H-field with gap $0$ .

Proof. Suppose that L is equipped with an ordering making it an ordered field extension of K with convex valuation ring such that the induced ordering on $\boldsymbol k_L$ agrees with the given one. Let $a \in L^{\times }$ . Then $a=su$ with $s \in K^>$ and $u \asymp 1$ in L. Since $u>0 \iff \overline {u}>0$ , there is at most one such ordering on L, and this also shows how to define the ordering on L: $a>0 \iff \overline {u}>0$ . This is independent of the choice of s and u: If $a=tu_1$ with $t \in K^>$ and $u_1 \asymp 1$ in L, then $u_1 = st^{-1}u$ and $\overline {u_1} = \overline {st^{-1}}\cdot \overline {u}$ , so $\overline {u_1}>0 \iff \overline {u}>0$ . Obviously, $a>0$ or $-a>0$ .

Next, assume that $a, b \in L^{>}$ ; we will show that $a+b>0$ and $ab>0$ . Then $a=su_1$ and $b=tu_2$ with $s, t \in K^>$ and $u_1, u_2 \asymp 1$ in $L^>$ . Without loss of generality, $s \preccurlyeq t$ , so $a+b=t(st^{-1}u_1+u_2)$ and

$$\begin{align*}\overline{st^{-1}u_1+u_2}\ =\ \overline{st^{-1}}\cdot\overline{u_1}+\overline{u_2} > 0. \end{align*}$$

Thus $st^{-1}u_1+u_2 \asymp 1$ and $a+b>0$ . Also, $ab = stu_1u_2$ with $\overline {u_1u_2}=\overline {u_1}\cdot \overline {u_2}>0$ , so $ab>0$ . Similarly, $a^2>0$ . Thus we have defined an ordering on L making it an ordered field extension of K. Obviously, if $a \prec 1$ , then $-1<a<1$ , so the valuation ring of L is convex with respect to this ordering, and by construction it induces the given ordering on $\boldsymbol k_L$ .

Finally, suppose that K is a pre-H-field with gap $0$ . Let $a,b \in L^{\times }$ with $a \preccurlyeq 1$ and $b \not \asymp 1$ . Since L has small derivation, we have $a' \preccurlyeq 1$ . Write $b=su$ with $s \in K^>$ and $u \asymp 1$ in L, so then $b^{\dagger } = s^{\dagger }+u^{\dagger }\sim s^{\dagger } \succ 1$ , since K has gap $0$ . Thus $a' \prec b^{\dagger }$ , showing that L is pre- $\operatorname {d}$ -valued. It has the same asymptotic couple as K, so still has gap $0$ . To see that L is a pre-H-field, it remains to check (PH3): If $b>\mathcal O_L$ , then likewise $b'\sim s'u> 0$ .

Lemma 3.2. Suppose that K is an ordered valued differential field with small derivation and convex valuation ring. Let $\boldsymbol k\langle y \rangle $ be an ordered differential field extension of $\boldsymbol k$ with y $\operatorname {d}$ -algebraic over $\boldsymbol k$ . Then there exists an ordered valued differential field extension $K\langle a \rangle $ of K such that:

  1. (i) $\Gamma _{K \langle a \rangle } = \Gamma $ ;

  2. (ii) $K \langle a \rangle $ has small derivation and convex valuation ring;

  3. (iii) $a\asymp 1$ and $\operatorname {\mathrm {res}}(K\langle a\rangle )=\boldsymbol k\langle \overline {a} \rangle \cong \boldsymbol k\langle y \rangle $ over $\boldsymbol k$ (as ordered differential fields);

  4. (iv) for any ordered valued differential field extension M of K with convex valuation ring that is $\operatorname {d}$ -henselian, every embedding $\boldsymbol k\langle \overline {a}\rangle \to \boldsymbol k_M$ over $\boldsymbol k$ is induced by an embedding $K\langle a\rangle \to M$ over K.

Moreover, if K is a pre-H-field with gap $0$ , then so is $K\langle a \rangle $ .

Proof. The existence of the valued differential field $K\langle a\rangle $ and its embedding property as a valued differential field are provided by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 6.3.2 and Lemma 7.1.4]. Equipping $K\langle a\rangle $ with the ordering from Lemma 3.1 gives the rest, with the embedding property as an ordered valued differential field following from the uniqueness of the ordering on $K\langle a\rangle $ .

For the $\operatorname {d}$ -transcendental analog, which likewise follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.3.1] and Lemma 3.1, we construe the fraction field $K \langle Y \rangle $ of $K\{Y\}$ as a valued differential field extension of K by extending and the map $P \mapsto v(P)$ to $K \langle Y \rangle $ in the obvious way. Then $\Gamma _{K \langle Y \rangle }=\Gamma $ , and $K\langle Y\rangle $ has small derivation if K does, in which case $\overline {Y}$ is $\operatorname {d}$ -transcendental over $\boldsymbol k$ and $\operatorname {\mathrm {res}}(K\langle Y\rangle ) = \boldsymbol k\langle \overline {Y}\rangle $ (see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 6.3]).

Lemma 3.3. Suppose that K is an ordered valued differential field with small derivation and convex valuation ring. Suppose that $\boldsymbol k\langle \overline {Y} \rangle $ is an ordered differential field extension of $\boldsymbol k$ and equip $K \langle Y \rangle $ with the ordering from Lemma 3.1. Let M be an ordered valued differential field extension of K with small derivation, convex valuation ring, and $a \in M$ with $a \asymp 1$ and $\overline {a}$ $\operatorname {d}$ -transcendental over $\boldsymbol k$ . Then there exists a unique embedding $K \langle Y \rangle \to M$ over K with $Y \mapsto a$ . If K is a pre-H-field with gap $0$ , then so is $K\langle Y \rangle $ .

Similarly, we have an ordered variant of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 7.1.5].

Corollary 3.4. Suppose that K is an ordered valued differential field with small derivation and convex valuation ring. Let $\boldsymbol k_L$ be an ordered differential field extension of $\boldsymbol k$ . Then K has an ordered valued differential field extension L with the following properties:

  1. (i) $\Gamma _L = \Gamma $ ;

  2. (ii) L has small derivation and convex valuation ring;

  3. (iii) $\operatorname {\mathrm {res}}(L) \cong \boldsymbol k_L$ over $\boldsymbol k$ (as ordered differential fields);

  4. (iv) for any ordered valued differential field extension M of K with convex valuation ring that is $\operatorname {d}$ -henselian, every embedding $\operatorname {\mathrm {res}}(L) \to \boldsymbol k_M$ over $\boldsymbol k$ is induced by an embedding $L \to M$ over K.

Moreover, if K is a pre-H-field with gap $0$ , then so is L.

4 Extensions controlled by the asymptotic couple

Towards our quantifier elimination and model completion results for pre-H-fields with gap $0$ , in this section we study extensions controlled by the asymptotic couple. We begin by studying H-asymptotic couples with gap $0$ in their own right and finding the model completion of this theory, which has quantifier elimination.

4.1 Asymptotic couples with small derivation

The material in this subsection is based on [Reference Aschenbrenner, van den Dries and van der Hoeven6], which, apart from its new results, revisits quantifier elimination for the theory of closed H-asymptotic couples from [Reference Aschenbrenner and van den Dries3], introducing several new lemmas that simplify the arguments. For convenience, we use [Reference Aschenbrenner, van den Dries and van der Hoeven5, Sections 6.5 and 9.2] as a reference instead of the original sources [Reference Aschenbrenner and van den Dries3, Reference Aschenbrenner and van den Dries4, Reference Rosenlicht18, Reference Rosenlicht19, Reference Rosenlicht20] for asymptotic couples.

In contrast with the results of [Reference Aschenbrenner and van den Dries3, Reference Aschenbrenner, van den Dries and van der Hoeven6], here we do not need to expand the language by a predicate for the $\Psi $ -set or by functions for divisibility by nonzero natural numbers. Additionally, those authors work over an arbitrary ordered scalar field $\boldsymbol k$ , but here we work over $\mathbb Q$ for concreteness (the results of this section hold in that setting in the language $\mathcal L_{\operatorname {ac}}$ , described later, expanded by functions for scalar multiplication).

4.1.1 Preliminaries

We suspend in this subsection the convention that $\Gamma $ is the value group of K. Instead, $(\Gamma , \psi )$ is an H-asymptotic couple, which means that $\Gamma $ is an ordered abelian group and $\psi \colon \Gamma ^{\neq } \to \Gamma $ is a map satisfying, for all $\gamma , \delta \in \Gamma ^{\neq }$ :

  1. (AC1) if $\gamma +\delta \neq 0$ , then $\psi (\gamma +\delta ) \geqslant \min \{\psi (\gamma ),\psi (\delta )\}$ ;

  2. (AC2) $\psi (k\gamma ) = \psi (\gamma )$ for all $k \in \mathbb {Z}^{\neq }$ ;

  3. (AC3) if $\gamma>0$ , then $\gamma +\psi (\gamma )>\psi (\delta )$ ;

  4. (HC) if $0<\gamma \leqslant \delta $ , then $\psi (\gamma ) \geqslant \psi (\delta )$ .

Keeping in mind that later $(\Gamma , \psi )$ will be the asymptotic couple of an H-asymptotic field (such as a pre-H-field), we let and for $\gamma \in \Gamma ^{\neq }$ . The map $\gamma \mapsto \gamma '$ for $\gamma \in \Gamma ^{\neq }$ is strictly increasing [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.5.4(iii)]. We let and for any ordered abelian group G we set , having earlier defined $G^>$ and $G^<$ . Thus (AC3) says that $\Psi <(\Gamma ^>)'$ .

A fundamental result is that $\Gamma \setminus (\Gamma ^{\neq })'$ has at most one element [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 9.2.1], and moreover exactly one of the following holds [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 9.2.16]:

  1. (1) there exists $\beta \in \Gamma $ such that $\Psi <\beta <(\Gamma ^>)'$ , in which case $(\Gamma , \psi )$ has gap $\beta $ ;

  2. (2) there exists $\beta \in \Gamma $ such that $\max \Psi =\beta $ , in which case $(\Gamma , \psi )$ has max $\beta $ ;

  3. (3) $(\Gamma ^{\neq })'=\Gamma $ , in which case $(\Gamma , \psi )$ has asymptotic integration.

We are primarily concerned with H-asymptotic couples having gap $0$ , although we also consider the case of max $0$ in this subsection. The material on H-asymptotic couples with max $0$ is only used in one later theorem that itself is not used in the main results, but fits naturally with that of the gap $0$ case. In contrast, the main results of [Reference Aschenbrenner and van den Dries3, Reference Aschenbrenner, van den Dries and van der Hoeven6] concern asymptotic couples with asymptotic integration, such as the asymptotic couple of $\mathbb {T}$ . In adapting their arguments, we try to highlight how the substitution of “gap $0$ ” or “max $0$ ” for “asymptotic integration” (or similar changes) alters the proofs.

Note that if $(\Gamma , \psi )$ is the asymptotic couple of an asymptotic field K, then K has gap $0$ (in the sense of Section 2) if and only if $(\Gamma , \psi )$ has gap $0$ . It follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 9.2.1 and Corollary 9.2.4] that $\sup \Psi =0 \iff (\Gamma ^>)' = \Gamma ^>$ and that $\sup \Psi =0 \notin \Psi $ if and only if $(\Gamma , \psi )$ has gap $0$ . Thus $\sup \Psi =0$ if and only if $(\Gamma , \psi )$ has gap $0$ or max $0$ .

It follows from (AC2) and (HC) that $\psi $ is constant on archimedean classes of $\Gamma $ . For $\gamma \in \Gamma $ , we let denote its archimedean class, and set , ordering it by $[\delta ]<[\gamma ]$ if $n|\delta |<|\gamma |$ for all n, where $\gamma , \delta \in \Gamma $ . The map $\psi $ extends uniquely to the divisible hull $\mathbb {Q}\Gamma $ of $\Gamma $ , defined by $\psi (q\gamma )=\psi (\gamma )$ for $\gamma \in \Gamma ^{\neq }$ and $q \in \mathbb {Q}^{\times }$ (use [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.5.3] to get (AC3)), and in this way we always construe $\mathbb {Q}\Gamma $ as an H-asymptotic couple $(\mathbb {Q}\Gamma , \psi )$ extending $(\Gamma , \psi )$ . It satisfies $\psi ((\mathbb {Q}\Gamma )^{\neq }) = \psi (\Gamma ^{\neq })$ , so if $(\Gamma , \psi )$ has gap $0$ (respectively, max $0$ ), then so does $(\mathbb {Q}\Gamma , \psi )$ . We also use that if $(\Gamma , \psi )$ has gap $0$ , then for every $\gamma \in \Gamma ^{\neq }$ , we have $[\gamma ]>[\gamma ^{\dagger }]$ by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.2.10(iv)] and so $[\gamma ]=[\gamma ']$ ; in particular, if $\gamma <0$ , then $\gamma <\gamma ^{\dagger }<0$ .

We call $(\Gamma , \psi )$ gap-closed if $\Gamma $ is nontrivial and divisible, and $\Psi =\Gamma ^<$ . The goal of this section is Theorem 4.8, which states that the theory of gap-closed H-asymptotic couples has quantifier elimination and is the model completion of the theory of H-asymptotic couples with gap $0$ . Likewise, the theory of max-closed H-asymptotic couples, where $(\Gamma , \psi )$ is max-closed if $\Gamma $ is divisible and $\Psi =\Gamma ^{\leqslant }$ , has quantifier elimination and is the model completion of the theory of H-asymptotic couples $(\Gamma , \psi )$ with $\sup \Psi =0$ . The language for these results is the language $\mathcal L_{\operatorname {ac}}=\{+, -, \leqslant , 0, \infty , \psi \}$ of asymptotic couples. The underlying set of $(\Gamma , \psi )$ in this language is , and we interpret $\infty $ in the following way: for all $\gamma \in \Gamma $ , and $\gamma <\infty $ ; ; ; . The other symbols have the expected interpretations.

One of the new extension lemmas of [Reference Aschenbrenner, van den Dries and van der Hoeven6] is [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.7]. Here is a variant that avoids assuming asymptotic integration by making a stronger cofinality assumption.

Lemma 4.1. Suppose that $\Psi $ is downward closed in $\Gamma $ . Let $(\Gamma _1, \psi _1)$ and $(\Gamma _*, \psi _*)$ be H-asymptotic couples extending $(\Gamma , \psi )$ such that $\Gamma ^<$ is cofinal in $\Gamma _1^<$ . Suppose that $\gamma _1 \in \Gamma _1 \setminus \Gamma $ and $\gamma _* \in \Gamma _* \setminus \Gamma $ realize the same cut in $\Gamma $ and $\gamma _1^{\dagger } \notin \Gamma $ . Then $\gamma _*^{\dagger } \notin \Gamma $ and $\gamma _*^{\dagger }$ realizes the same cut in $\Gamma $ as $\gamma _1^{\dagger }$ .

Proof. Let $\alpha \in \Gamma ^{\neq }$ . We first show that:

$$\begin{align*}\gamma_1^{\dagger} < \alpha^{\dagger} \implies \gamma_*^{\dagger} < \alpha^{\dagger} \qquad \text{and} \qquad \gamma_1^{\dagger}> \alpha^{\dagger} \implies \gamma_*^{\dagger} > \alpha^{\dagger}. \end{align*}$$

The first implication follows exactly as in [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.7], which uses that $\Psi $ is downward closed but not that $(\Gamma , \psi )$ has asymptotic integration. The second implication has a hidden use of asymptotic integration, so here are the modifications.

Suppose that $\gamma _1^{\dagger }>\alpha ^{\dagger }$ , so by the cofinality assumption, there is $\delta \in \Gamma $ with $\gamma _1^{\dagger }> \delta > \alpha ^{\dagger }$ . By the reasoning in the first implication (see [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.7]), it suffices to show that $\delta \in \Psi $ ; asymptotic integration is used in [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.7] to get this. By the stronger cofinality assumption here, take $\beta \in \Gamma ^{\neq }$ with $|\gamma _1|>|\beta |$ , so $\delta <\gamma _1^{\dagger }< \beta ^{\dagger } \in \Psi $ . Thus $\delta \in \Psi $ , since $\Psi $ is downward closed.

Finally, the last paragraph also shows that $\gamma _*^{\dagger } \notin \Gamma $ , since $\gamma _{*}^{\dagger }\leqslant \beta ^{\dagger } \in \Psi $ and $\Psi $ is downward closed, and therefore it realizes the same cut as $\gamma _1^{\dagger }$ by the displayed implications.

Alternatively, we can weaken the cofinality assumption from $\Gamma ^<$ being cofinal in $\Gamma _1^<$ to $\Gamma ^<$ being cofinal in $(\Gamma + \mathbb {Q}\gamma _1^{\dagger })^<$ (as in [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.7]) in the case that $\sup \Psi _1=0$ ; the only change in the proof is that now $\delta \in \Psi $ follows from $\delta <\gamma _1^{\dagger } \leqslant 0$ .

Another of the new lemmas is the following, [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.8], which we also use in Section 4.2.

Lemma 4.2. Suppose that $(\Gamma _1, \psi _1)$ is an H-asymptotic couple extending $(\Gamma , \psi )$ , and let $\gamma _1 \in \Gamma _1 \setminus \Gamma $ and $\alpha \in \Gamma $ . Suppose that $\gamma _1$ and satisfy $\gamma _1^{\dagger } \notin \Gamma $ and $\beta ^{\dagger } \notin \Psi $ , and that $|\gamma _1|>|\gamma |$ for some $\gamma \in \Gamma ^{\neq }$ . Then $\gamma _1^{\dagger } < \beta ^{\dagger }$ .

4.1.2 Embedding lemmas

We now turn to the proof of quantifier elimination for gap-closed and max-closed H-asymptotic couples, which we handle uniformly as much as possible. To that end, suppose that $(\Gamma , \psi )$ is a divisible H-asymptotic couple with $\sup \Psi =0$ , and let $(\Gamma _1, \psi _1)$ and $(\Gamma _*, \psi _*)$ be divisible H-asymptotic couples extending $(\Gamma , \psi )$ such that $(\Gamma _*, \psi _*)$ is $|\Gamma |^+$ -saturated. Let $\gamma _1 \in \Gamma _1 \setminus \Gamma $ and $(\Gamma \langle \gamma _1 \rangle , \psi _1)$ be the divisible H-asymptotic couple generated by $\Gamma \cup \{\gamma _1\}$ in $(\Gamma _1, \psi _1)$ .

The first two lemmas are proved similarly to [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemmas 3.4 and 3.5], so we only sketch the proofs, omitting some details that are the same. The third lemma is particular to the case of gap $0$ . For convenience, we set , so $\Gamma ^{\dagger } = \Psi \cup \{\infty \}$ .

In the next lemma, the $\sup \Psi =0$ assumption simplifies the case distinctions needed in [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.4], and the assumption of asymptotic integration there is used only in Cases 1–2 of that lemma, which do not occur here. We also do not need the “H-cuts” of that paper for quantifier elimination since the $\Psi $ -set in a gap-closed or a max-closed H-asymptotic couple is already quantifier-free definable without parameters.

Lemma 4.3. Suppose that $(\Gamma + \mathbb Q\gamma _1)^{\dagger } = \Gamma ^{\dagger }$ . Then $(\Gamma \langle \gamma _1 \rangle , \psi _1)$ can be embedded into $(\Gamma _*, \psi _*)$ over $\Gamma $ .

Proof. From $(\Gamma +\mathbb Q\gamma _1)^{\dagger } = \Gamma ^{\dagger }$ , we get $\Gamma \langle \gamma _1 \rangle = \Gamma +\mathbb Q\gamma _1$ .

In case $[\Gamma +\mathbb Q\gamma _1] = [\Gamma ]$ , realizing in $\Gamma _*$ the cut in $\Gamma $ realized by $\gamma _1$ yields an embedding $i \colon \Gamma +\mathbb Q\gamma _1 \to \Gamma _*$ of ordered $\mathbb {Q}$ -vector spaces fixing $\Gamma $ , and the assumption $[\Gamma +\mathbb Q\gamma _1] = [\Gamma ]$ ensures that it is an embedding of H-asymptotic couples (see Case 3 of [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.4]).

In case $[\Gamma +\mathbb Q\gamma _1] \neq [\Gamma ]$ but $\Gamma ^<$ is cofinal in $(\Gamma +\mathbb Q\gamma _1)^<$ , take $\beta \in \Gamma _1 \setminus \Gamma $ with $\beta>0$ and $[\beta ] \notin [\Gamma ]$ . Letting D 1 be the cut in $\Gamma $ realized by $\beta $ and , it follows from the cofinality assumption that D 1 has no greatest element and D 2 has no least element. Then saturation gives $\beta _* \in \Gamma _*$ realizing the same cut as $\beta $ in $\Gamma $ and with $\beta ^{\dagger } = \beta _*^{\dagger }$ , yielding an embedding $i \colon \Gamma +\mathbb Q\gamma _1 \to \Gamma _*$ of H-asymptotic couples fixing $\Gamma $ (see Cases 4–6 of [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.4]).

In case $\Gamma ^<$ is not cofinal in $(\Gamma +\mathbb Q\gamma _1)^<$ , take $\beta \in \Gamma +\mathbb Q\gamma _1$ satisfying $0<\beta <\Gamma ^>$ . In this case, $(\Gamma , \psi )$ must have max $0$ , since if $(\Gamma , \psi )$ has gap $0$ , then $\Psi <\psi _1(\beta ) \in \Psi $ , a contradiction. By saturation, take $\beta _* \in \Gamma _*$ with $0<\beta _*<\Gamma ^>$ and $\beta _*^{\dagger }=\beta ^{\dagger }=0$ , so we get an embedding $i \colon \Gamma +\mathbb Q\gamma _1 \to \Gamma _*$ of H-asymptotic couples as before.

In [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.5], the cofinality of $\Gamma ^<$ in $\Gamma _1^<$ follows from asymptotic integration; In the max $0$ case it is also automatic, but needs to be added as an assumption in the gap $0$ case, (i) in the next lemma.

Lemma 4.4. Suppose that $(\Gamma +\mathbb Q\gamma )^{\dagger } \neq \Gamma ^{\dagger }$ for all $\gamma \in \Gamma _1 \setminus \Gamma $ . Also, suppose that either:

  1. (i) $(\Gamma , \psi )$ is gap-closed, $(\Gamma _1, \psi _1)$ has gap $0$ , and $\Gamma ^<$ is cofinal in $\Gamma _1^<$ ; or

  2. (ii) $(\Gamma , \psi )$ is max-closed and $(\Gamma _1,\psi _1)$ has max $0$ .

Then $(\Gamma \langle \gamma _1\rangle , \psi _1)$ can be embedded into $(\Gamma _*, \psi _*)$ over $\Gamma $ .

Proof. Note that in case (ii), the cofinality assumption comes for free, for if $\gamma \in \Gamma _1 \setminus \Gamma $ satisfies $0<\gamma <\Gamma ^>$ , then $\gamma ^{\dagger } = 0$ and so $(\Gamma +\mathbb Q\gamma )^{\dagger } = \Gamma ^{\dagger }$ , a contradiction.

Take $\alpha _1 \in \Gamma $ such that $(\gamma _1-\alpha _1)^{\dagger } \notin \Gamma ^{\dagger }$ . Then $(\gamma _1-\alpha _1)^{\dagger } \notin \Gamma $ , since $(\gamma _1-\alpha _1)^{\dagger }<0$ and in case (i) and case (ii), $\Psi =\Gamma ^{<}$ and $\Psi =\Gamma ^{\leqslant }$ , respectively. (In [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.5], this step uses asymptotic integration.)

The proof now proceeds exactly as in [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.5], substituting Lemma 4.1 for [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 2.7] (and $\mathbb {Q}$ for $\boldsymbol k$ ); here is a sketch. Let $n \geqslant 1$ . Continue this procedure to construct sequences $\alpha _1, \alpha _2, \dots $ in $\Gamma $ and $\beta _1, \beta _2, \dots $ in $\Gamma \langle \gamma _1 \rangle \setminus \Gamma $ with $\beta _1 = \gamma _1-\alpha _1$ and $\beta _{n+1} = \beta _n^{\dagger }-\alpha _{n+1}$ . Then $\beta _n^{\dagger }<\beta _{n+1}^{\dagger }$ by Lemma 4.2 (with $\beta _n$ in place of $\gamma _1$ and $\beta _{n+1}$ in place of $\beta $ ), so $[\beta _n]>[\beta _{n+1}]$ , so

$$\begin{align*}\Gamma\langle \gamma_1 \rangle\ =\ \Gamma \oplus \mathbb Q\beta_1 \oplus \mathbb Q\beta_2 \oplus \cdots. \end{align*}$$

Saturation gives $\gamma _* \in \Gamma _* \setminus \Gamma $ realizing the same cut in $\Gamma $ as $\gamma _1$ , so use the same $\alpha _n$ sequence to likewise define $\beta _{*n} \in (\Gamma _*)_{\infty }$ for $n \geqslant 1$ . Inductively construct an increasing sequence of $\mathbb {Q}$ -vector space embeddings

$$\begin{align*}i_n \colon \Gamma + \mathbb Q\beta_1 + \dots + \mathbb Q\beta_n \to \Gamma_* \end{align*}$$

fixing $\Gamma $ and sending $\beta _n$ to $\beta _{*n}$ in two steps: Given that $\beta _n$ and $\beta _{*n}$ realize the same cut in $\Gamma $ , use Lemma 4.1 to get that $\beta _n^{\dagger }$ and $\beta _{*n}^{\dagger }$ realize the same cut in $\Gamma $ . Then argue with archimedean classes, using Lemma 4.2 in $(\Gamma _*,\psi _*)$ now, to extend the embedding to $i_{n+1}$ . The union of these maps is the desired embedding of H-asymptotic couples.

In case (i) of the previous lemma, we added a cofinality assumption that was not present in [Reference Aschenbrenner, van den Dries and van der Hoeven6, Lemma 3.5]. The next lemma handles the non-cofinal case, which is particular to gap $0$ .

Lemma 4.5. Suppose that $\Gamma ^< < \gamma _1 <0$ , and $(\Gamma _1,\psi _1)$ and $(\Gamma _*,\psi _*)$ have gap $0$ . Then $(\Gamma \langle \gamma _1\rangle , \psi _1)$ can be embedded into $(\Gamma _*, \psi _*)$ over $\Gamma $ .

Proof. Set and for all n. We have

$$\begin{align*}[\gamma_1]>[\gamma_1^{\langle 1 \rangle}]>[\gamma_1^{\langle 2 \rangle}]>\cdots, \end{align*}$$

and so

$$\begin{align*}\Gamma^< < \gamma_1 < \gamma_1^{\langle 1 \rangle} < \gamma_1^{\langle 2 \rangle}<\dots<0 \qquad \text{and} \qquad [\gamma_1^{\langle n \rangle}] \notin [\Gamma]\ \text{for all}\ n. \end{align*}$$

Hence the family $(\gamma _1^{\langle n \rangle })_{n \in \mathbb N}$ is $\mathbb Q$ -linearly independent over $\Gamma $ and

$$\begin{align*}\Gamma\langle \gamma_1 \rangle\ =\ \Gamma \oplus \mathbb Q\gamma_1 \oplus \mathbb Q\gamma_1^{\langle 1 \rangle} \oplus \mathbb Q\gamma_1^{\langle 2 \rangle} \oplus \cdots. \end{align*}$$

By saturation, we may take $\gamma _* \in \Gamma _* \setminus \Gamma $ with $\Gamma ^< < \gamma _* < 0$ . The above holds in $\Gamma _*$ with $\gamma _*$ replacing $\gamma _1$ (and $\gamma _*^{\langle n \rangle }$ defined analogously), so we obtain an embedding of $(\Gamma \langle \gamma _1 \rangle , \psi _1)$ into $(\Gamma _*, \psi _*)$ over $\Gamma $ that sends $\gamma _1$ to $\gamma _*$ .

4.1.3 Quantifier elimination

We call an H-asymptotic couple $(\Gamma _1, \psi _1)$ extending $(\Gamma , \psi )$ a gap-closure of $(\Gamma , \psi )$ if it is gap-closed and it embeds over $(\Gamma , \psi )$ into every gap-closed H-asymptotic couple extending $(\Gamma , \psi )$ . Similarly, we call an H-asymptotic couple $(\Gamma _1, \psi _1)$ extending $(\Gamma , \psi )$ a max-closure of $(\Gamma , \psi )$ if it is max-closed and it embeds over $(\Gamma , \psi )$ into every max-closed H-asymptotic couple extending $(\Gamma , \psi )$ . By the embedding lemmas of the previous subsection and a standard quantifier elimination test (see for example [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary B.11.11]), quantifier elimination for gap-closed and max-closed H-asymptotic couples reduces to the existence of gap-closures and max-closures, respectively.

For that, we need one more embedding lemma, Let $\Psi ^{\downarrow }$ be the downward closure of $\Psi $ in $\Gamma $ .

Lemma 4.6. Let $\beta \in \Psi ^{\downarrow } \setminus \Psi $ or $\beta $ be a gap in $(\Gamma , \psi )$ . Then there is an H-asymptotic couple $(\Gamma \oplus \mathbb Z\alpha , \psi ^\alpha )$ extending $(\Gamma , \psi )$ such that:

  1. (i) $\alpha>0$ and $\psi ^\alpha (\alpha )=\beta $ ;

  2. (ii) given an embedding i of $(\Gamma , \psi )$ into an H-asymptotic couple $(\Gamma ^{*}, \psi ^{*})$ and $\alpha ^{*} \in \Gamma ^{*}$ with $\alpha ^{*}>0$ and $\psi ^{*}(\alpha ^{*})=i(\beta )$ , there is a unique extension of i to an embedding $j \colon (\Gamma \oplus \mathbb Z\alpha , \psi ^\alpha ) \to (\Gamma ^{*}, \psi ^{*})$ with $j(\alpha )=\alpha ^{*}$ .

Proof. This follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.8.7] with $C=\{[\gamma ] : \gamma \in \Gamma ^{\neq },\ \psi (\gamma )>\beta \}$ , but here is an outline. Define the ordered abelian group $\Gamma \oplus \mathbb Z\alpha $ so that

$$\begin{align*}\{0\} \cup \{ \gamma \in \Gamma^> : \psi(\gamma)>\beta \}\ <\ \alpha\ <\ \{ \gamma \in \Gamma^> : \psi(\gamma)<\beta \}, \end{align*}$$

and thus $[\alpha ] \notin [\Gamma ^{\neq }]$ . Extend $\psi $ to $\psi ^{\alpha } \colon (\Gamma \oplus \mathbb Z\alpha )^{\neq } \to \Gamma $ by , where $\gamma \in \Gamma $ and $k \in \mathbb Z^{\neq }$ . To verify that this makes $(\Gamma \oplus \mathbb Z\alpha , \psi ^\alpha )$ an H-asymptotic couple involves tedious case distinctions. The axioms (AC1), (AC2), and (HC) are straightforward, while (AC3) is more subtle and uses [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.5.4]. The universal property of $(\Gamma \oplus \mathbb Z\alpha , \psi ^\alpha )$ is easy.

Corollary 4.7. Every H-asymptotic couple $(\Gamma , \psi )$ with gap $0$ has a gap-closure. Every H-asymptotic couple $(\Gamma , \psi )$ with $\sup \Psi =0$ has a max-closure.

Proof. If $(\Gamma ,\psi )$ has max $0$ , then by applying Lemma 4.6 for $\beta \in \Psi ^{\downarrow } \setminus \Psi $ , taking the divisible hull, and repeating these procedures, we get a max-closure of $(\Gamma , \psi )$ . If $(\Gamma ,\psi )$ has gap $0$ and $\Gamma \neq \{0\}$ , the same procedure yields a gap-closure of $(\Gamma , \psi )$ . To obtain a max-closure of $(\Gamma ,\psi )$ when it has gap $0$ , first apply Lemma 4.6 with $\beta =0$ and then repeat the previous argument.

It remains to show that the trivial H-asymptotic couple $(\{0\},\psi )$ , where $\psi \colon \emptyset \to \{0\}$ is the empty function, has a gap-closure. Let be the ordered $\mathbb {Q}$ -vector space satisfying $\gamma _n<0$ and $[\gamma _n]>[\gamma _{n+1}]$ for all n, and equip it with the function $\psi _0$ defined by $\psi _0(q_1\gamma _{n_1}+\dots +q_m\gamma _{n_m})=\gamma _{n_1+1}$ for all $m \geqslant 1$ , $n_1>\dots >n_m$ , and $q_1, \dots , q_m \in \mathbb {Q}^{\times }$ . Then by the proof of Lemma 4.5, $(\Gamma _0, \psi _0)$ is an H-asymptotic couple with gap $0$ that embeds into every nontrivial H-asymptotic couple with gap  $0$ . Hence, a gap-closure of $(\Gamma _0,\psi _0)$ is a gap-closure of $(\{0\},\psi )$ .

Next we establish quantifier elimination, first proving it in , the language $\mathcal L_{\operatorname {ac}}$ expanded by unary function symbols $\operatorname {div}_n$ interpreted as division by n with . As noted already, unlike in [Reference Aschenbrenner and van den Dries3, Reference Aschenbrenner, van den Dries and van der Hoeven6], we do not need the predicate for the $\Psi $ -set, since it is already quantifier-free definable without parameters.

Theorem 4.8. The theory of gap-closed H-asymptotic couples has quantifier elimination, and it is the model completion of the theory of H-asymptotic couples with gap $0$ . The theory of max-closed H-asymptotic couples has quantifier elimination, and it is the model completion of the theory of H-asymptotic couples $(\Gamma , \psi )$ with $\sup \Psi =0$ .

Proof. That these theories have quantifier elimination in $\mathcal L_{\operatorname {ac}, \operatorname {div}}$ follows from Lemmas 4.3, 4.4, and 4.5, and Corollary 4.7 by a standard quantifier elimination test. To see that they have quantifier elimination in $\mathcal L_{\operatorname {ac}}$ , recall how, for an H-asymptotic couple $(\Gamma , \psi )$ , $\psi $ extends uniquely to the divisible hull $\mathbb {Q}\Gamma $ of $\Gamma $ , preserving the property of having gap $0$ or max $0$ .

The model completion statements follow from quantifier elimination and Corollary 4.7.

Corollary 4.9. The theory of gap-closed H-asymptotic couples is complete and has a prime model. The theory of max-closed H-asymptotic couples is complete and has a prime model.

Proof. The trivial H-asymptotic couple (see the proof of Corollary 4.7), embeds into every H-asymptotic couple, yielding completeness of the two theories. It also has gap $0$ , so its gap-closure and its max-closure are the respective prime models.

As prime models of their respective complete theories, on standard model-theoretic grounds (see for instance [Reference Marker13, Corollary 4.2.16]) the gap-closure and the max-closure of the trivial H-asymptotic couple are unique up to isomorphism. By quantifier elimination and adding constants to the language, similar reasoning applies to the gap-closure and the max-closure of a countable H-asymptotic couple. However, we do not know in general if gap-closures and max-closures are unique up to isomorphism. One possibility is foreclosed: By adapting the arguments of [Reference Aschenbrenner1, Section 3], we can show that if an H-asymptotic couple $(\Gamma , \psi )$ with gap $0$ is not gap-closed, then a gap-closure of $(\Gamma , \psi )$ is not minimal in the sense that it has a proper gap-closed substructure containing $\Gamma $ . The same is true replacing “gap $0$ ” with “ $\sup \Psi =0$ ,” “gap-closed” with “max-closed,” and “gap-closure” with “max-closure.” This nonminimality result has no bearing on the rest of the paper, so we do not present details here. With more work, one could likely obtain further model-theoretic results on such H-asymptotic couples, along the lines of [Reference Aschenbrenner and van den Dries3, Reference Aschenbrenner, van den Dries and van der Hoeven6].

We do use Theorem 4.8 via the following corollary. For $n \geqslant 1$ , $\alpha _1, \dots , \alpha _n \in \Gamma $ , and $\gamma \in \Gamma $ , we define the function $\psi _{\alpha _1,\dots ,\alpha _n} \colon \Gamma _\infty \to \Gamma _\infty $ recursively by

Corollary 4.10. Let $(\Gamma , \psi )$ be a gap-closed H-asymptotic couple and let $(\Gamma ^{*}, \psi ^{*})$ be an H-asymptotic couple extending $(\Gamma , \psi )$ with gap $0$ . Suppose $n \geqslant 1$ , $\alpha _1, \dots , \alpha _n \in \Gamma $ , $q_1, \dots , q_n \in \mathbb {Q}$ , and $\gamma ^{*} \in \Gamma ^{*}$ are such that:

  1. (i) $\psi ^{*}_{\alpha _1, \dots , \alpha _n}(\gamma ^{*}) \neq \infty $ (so $\psi ^{*}_{\alpha _1, \dots , \alpha _i}(\gamma ^{*}) \neq \infty $ for $i=1,\dots ,n$ );

  2. (ii) $\gamma ^{*} + q_1 \psi ^{*}_{\alpha _1}(\gamma ^{*}) + \dots + q_n \psi ^{*}_{\alpha _1, \dots , \alpha _n}(\gamma ^{*}) \in \Gamma $ (in $\mathbb {Q}\Gamma ^{*}$ ).

Then $\gamma ^{*} \in \Gamma $ .

Proof. By Theorem 4.8, $(\Gamma , \psi )$ is an existentially closed H-asymptotic couple with gap $0$ , so we have $\gamma \in \Gamma $ with

$$\begin{align*}\gamma + q_1 \psi_{\alpha_1}(\gamma) + \dots + q_n \psi_{\alpha_1, \dots, \alpha_n}(\gamma)\ =\ \gamma^{*} + q_1 \psi^{*}_{\alpha_1}(\gamma^{*}) + \dots + q_n \psi^{*}_{\alpha_1, \dots, \alpha_n}(\gamma^{*}). \end{align*}$$

It remains to use [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.9.3] to obtain $\gamma ^{*}=\gamma \in \Gamma $ .

The same holds with “max-closed” replacing “gap-closed”and “ $\sup \Psi =0$ ” replacing “gap $0$ .”

4.2 A maximality theorem

The results and proofs of this section are adapted from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 16.1]. In the next lemma and its consequences, we use the quantifier elimination for gap-closed asymptotic couples from Section 4.1 to study extensions of certain asymptotic fields whose asymptotic couples are gap-closed. Note that if K is an H-asymptotic field with exponential integration and gap $0$ , then $\Psi = \Gamma ^<$ , so if additionally $\Gamma $ is divisible then $(\Gamma , \psi )$ is gap-closed.

Assumption. In this subsection, K has small derivation.

The next lemma follows from Corollary 4.10 in the same way that [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.1] follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 9.9.2], except with “ $\operatorname {d}$ -algebraically maximal” replacing “asymptotically $\operatorname {d}$ -algebraically maximal”: K is differential-algebraically maximal ( $\operatorname {d}$ -algebraically maximal for short) if it has no proper differentially algebraic (“ $\operatorname {d}$ -algebraic” for short) immediate extension with small derivation. And instead of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 14.0.2] (used implicitly in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.1]), we need [Reference Pynn-Coates14, Theorem 3.6]. We also need the notion of a pseudocauchy sequence (“pc-sequence” for short); for the definition and basic facts about pc-sequences, see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 2.2], and for pc-sequences in valued differential fields, see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Sections 4.4 and 6.9]. In an immediate extension L of K, every element of $L \setminus K$ is the pseudolimit of a pc-sequence in K that has no pseudolimit in K, called divergent in K, and divergent pc-sequences in K can be of $\operatorname {d}$ -algebraic or $\operatorname {d}$ -transcendental type over K. The exact definitions are not really important here, for which see the end of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 4.4]. But note that if the derivation on $\boldsymbol k$ is nontrivial, then K is $\operatorname {d}$ -algebraically maximal if and only if there is no divergent pc-sequence in K of $\operatorname {d}$ -algebraic type over K by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.9.3].

Lemma 4.11. Suppose that K is a $\operatorname {d}$ -henselian H-asymptotic field with exponential integration and gap $0$ whose value group is divisible. Let L be an H-asymptotic extension of K with gap $0$ and $\boldsymbol k_L = \boldsymbol k$ , and suppose that there is no $y \in L \setminus K$ such that $K\langle y \rangle $ is an immediate extension of K. Let $f \in L \setminus K$ . Then the vector space $\mathbb {Q}\Gamma _{K\langle f \rangle }/\Gamma $ is infinite dimensional.

Proof. First, we argue that there is no divergent pc-sequence in K with a pseudolimit in L. Towards a contradiction, suppose that $(a_\rho )$ is a divergent pc-sequence in K with pseudolimit $\ell \in L$ . Since K is $\operatorname {d}$ -henselian and asymptotic, it is $\operatorname {d}$ -algebraically maximal [Reference Pynn-Coates14, Theorem 3.6], so $(a_\rho )$ is not of $\operatorname {d}$ -algebraic type over K. Hence $(a_\rho )$ is of $\operatorname {d}$ -transcendental type over K, so $K\langle \ell \rangle $ is an immediate extension of K by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.9.1], a contradiction.

Thus for all $y \in L \setminus K$ , the set $v(y-K) \subseteq \Gamma _L$ has a maximum. Take $b_0 \in K$ with $v(f-b_0) = \max v(f-K)$ , so $v(f-b_0) \notin \Gamma $ since $\boldsymbol k_L = \boldsymbol k$ . Set . Then $f_1 \notin K$ , since otherwise, there would be $g \in K^{\times }$ with $(f-b_0)^{\dagger } = g^{\dagger }$ , so $v(f-b_0) = vg \in \Gamma $ , a contradiction. Setting , this leads to sequences $(f_n)$ in $L \setminus K$ and $(b_n)$ in K such that for all n:

  1. (1) $v(f_n-b_n) = \max v(f_n-K) \notin \Gamma $ ;

  2. (2) $f_{n+1} = (f_n-b_n)^{\dagger }$ .

To finish the proof of the lemma, show that

$$\begin{align*}v(f_0-b_0),\ v(f_1-b_1),\ \dots\ \text{are } \mathbb{Q}\text{-linearly independent over } \Gamma. \end{align*}$$

If $\Gamma = \{ 0 \}$ , then this follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.2.10(iv)]. Otherwise, the idea is that by taking for $n \geqslant 1$ an $a_n \in K^{\times }$ with $a_n^{\dagger }=b_n$ and setting , Corollary 4.10 (with one of the $v(f_n-b_n)$ in the role of $\gamma $ ) shows that a $\mathbb {Q}$ -linear dependence is impossible since each $v(f_n-b_n) \notin \Gamma $ .

The next result is a strengthening of [Reference Pynn-Coates14, Theorem 3.6] under additional hypotheses, including that the asymptotic couple of K is gap-closed. It follows from the Zariski–Abhyankar inequality by combining [Reference Pynn-Coates14, Theorem 3.6] and Lemma 4.11 just as [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 16.0.3] combines [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 14.0.2] and [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.1]. This result is used in Section 6 to prove the minimality of $\operatorname {d}$ -Hensel–Liouville closures.

Theorem 4.12. Suppose that K is a $\operatorname {d}$ -henselian H-asymptotic field with exponential integration and gap $0$ whose value group is divisible. Then K has no proper $\operatorname {d}$ -algebraic H-asymptotic extension with gap $0$ and the same residue field.

By quantifier elimination for max-closed H-asymptotic couples, the above result holds with “max $0$ ” replacing “gap $0$ ,” where we say an asymptotic field K has max $0$ if its asymptotic couple does.

We now summarize some facts about the asymptotic couple of the $K\langle f \rangle $ from Lemma 4.11 that we need in the next subsection.

Lemma 4.13. Let K, L, and f be as in Lemma 4.11, and let the sequences $(f_n)$ , $(b_n)$ , $(a_n)_{n \geqslant 1}$ , and $(\alpha _n)_{n \geqslant 1}$ be as in the proof of Lemma 4.11. Set . The asymptotic couple $(\Gamma _{K\langle f \rangle }, \psi )$ of $K\langle f \rangle $ has the following properties:

  1. (i) $(\beta _n)$ is $\mathbb {Q}$ -linearly independent over $\Gamma $ and $\Gamma _{K\langle f \rangle } = \Gamma \oplus \bigoplus _n \mathbb {Z}\beta _n$ (internal direct sum);

  2. (ii) $(\beta _n^{\dagger })$ is $\mathbb {Q}$ -linearly independent over $\Gamma $ , so $\beta _n^{\dagger } \notin \Gamma $ and $\beta _m^{\dagger } \neq \beta _n^{\dagger }$ for all $m \neq n$ ;

  3. (iii) $\psi (\Gamma _{K\langle f \rangle }^{\neq }) = \Psi \cup \{ \beta _n^{\dagger } : n \in \mathbb {N} \}$ ;

  4. (iv) $[\beta _n] \notin [\Gamma ]$ for all n, $[\beta _m] \neq [\beta _n]$ for all $m \neq n$ , and $[\Gamma _{K\langle f \rangle }] = [\Gamma ] \cup \{ [\beta _n] : n \in \mathbb {N} \}$ ;

  5. (v) if $\Gamma ^<$ is cofinal in $\Gamma _{K\langle f \rangle }^<$ , then $\beta _0^{\dagger } < \beta _1^{\dagger } < \beta _2^{\dagger } < \cdots $ .

Proof. The $\mathbb {Q}$ -linear independence of $(\beta _n)$ over $\Gamma $ follows from the proof of Lemma 4.11. Items (i)–(iv) are proved exactly as in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.2], without modification, but are sketched below. The only difference is (v): In [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.2], cofinality is established unconditionally. Here, Lemma 4.2 immediately gives (v).

The hardest part is (i), which involves showing that the element $P(f) \in K\langle f \rangle $ , where $P \in K\{Y\}^{\neq }$ , can be expressed as a polynomial in the “monomials” . Since $v(\mathfrak m_n) = \beta _n$ , item (i) then follows from the $\mathbb {Q}$ -linear independence of $(\beta _n)$ over $\Gamma $ . With that done, the $\mathbb {Q}$ -linear independence of $(\beta _n^{\dagger })$ over $\Gamma $ follows from the $\mathbb {Q}$ -linear independence of $(\beta _n)_{n \geqslant 1}$ over $\Gamma $ by noticing that $\beta _n^{\dagger } = \beta _{n+1}+\alpha _{n+2}$ , and the rest of (ii)–(iv) follows.

4.3 Further consequences in the ordered setting

Now we develop further the results of the previous subsection in the pre-H-field setting. If K, L, and f are as before but additionally pre-H-fields, and M is a pre-H-field with $g \in M$ realizing the same cut in K as f, our goal is to embed $K\langle f \rangle $ into M over K by sending f to g. This is accomplished in Proposition 4.15 and Lemma 4.16. The former handles the case that $\Gamma ^<$ is cofinal in $\Gamma _{K\langle f \rangle }^<$ , which follows from the previous subsection by making small adjustments to the arguments for analogous results in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 16.1]. The case that $\Gamma ^<$ is not cofinal in $\Gamma _{K\langle f \rangle }^<$ does not occur in that context, and here is handled in Lemma 4.16 by similar ideas.

In the next lemma, the reader may assume without harm that K, L, and f additionally satisfy the assumptions of Lemma 4.11, but the lemma is stated in more generality to clarify which assumptions are used. In particular, it does not assume that K and L have gap $0$ or that K is $\operatorname {d}$ -henselian. Small modifications to the proof of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.4] yield the following generalization.

Lemma 4.14. Let K be a pre-H-field with exponential integration and let L be a pre-H-field extension of K with $\boldsymbol k_L = \boldsymbol k$ . Suppose that $f \in L \setminus K$ and $b_0 \in K$ satisfy:

  1. (i) $v_L(f-K) \subseteq \Gamma _L$ has a maximum and $v_L(f-b_0) = \max v_L(f-K)$ ;

  2. (ii) $\Gamma ^<$ is cofinal in $\Gamma _{K\langle f \rangle }^<$ .

Suppose that M is a pre-H-field extension of K and $g \in M$ realizes the same cut in K as f. Then $v_M(g-b_0) = \max v_M(g-K) \notin \Gamma $ and $(g-b_0)^{\dagger }$ realizes the same cut in K as $(f-b_0)^{\dagger }$ .

Proof. Let $\alpha \in \Gamma $ and $b \in K$ . First we use the convexity of $\mathcal O_{K\langle f \rangle }$ and the cofinality assumption to show that

$$\begin{align*}v_L(f-b)<\alpha \iff v_M(g-b)<\alpha \qquad \text{and} \qquad v_L(f-b)>\alpha \iff v_M(g-b)>\alpha. \end{align*}$$

To see this, take $a \in K^>$ with $va = \alpha $ . Suppose that $v_L(f-b)<\alpha $ , so $|f-b|>a$ . Hence $|g-b|>a$ , and so $v_M(g-b)\leqslant \alpha $ . To get the strict inequality, use the cofinality assumption to take $\delta \in \Gamma $ with $v_L(f-b)<\delta <\alpha $ , and thus $v_M(g-b) \leqslant \delta < \alpha $ . One proves similarly that $v_L(f-b)>\alpha \implies v_M(g-b)>\alpha $ . Finally, consider the case that $v_L(f-b)=\alpha $ . This yields $f-b \sim ua$ for $u \in K$ with $u \asymp 1$ , since $\boldsymbol k = \boldsymbol k_L$ . Convexity yields $|u|a/2 < |f-b| < 2|u|a$ , so $|u|a/2 < |g-b| < 2|u|a$ , and thus $v_M(g-b)=va=\alpha $ , completing the proof of the displayed equivalences. Hence, $v_M(g-b_0) \notin \Gamma $ , since $v_L(f-b_0) \notin \Gamma $ . This in turn yields $v_M(g-b_0)=\max v_M(g-K)$ . It also follows that $(g-b_0)^{\dagger } \notin K$ , as otherwise $(g-b_0)^{\dagger } = b^{\dagger }$ for some $b \in K^{\times }$ , in which case $v_M(g-b_0)=vb \in \Gamma $ .

As written, the argument in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.4] that $(g-b_0)^{\dagger }$ realizes the same cut in K as $(f-b_0)^{\dagger }$ uses that K and L are H-fields, so we give the following modified version. First, we may assume that $f>b_0$ , so $g>b_0$ . Suppose towards a contradiction that we have $h \in K$ with $(f-b_0)^{\dagger }<h$ and $h<(g-b_0)^{\dagger }$ . Take ${\phi \in K^>}$ with $h=\phi ^{\dagger }$ and set . Then we have $s>0$ and $s^{\dagger }=(f-b_0)^{\dagger }-h<0$ . Since L is a pre-H-field, $v_L(s) \geqslant 0$ by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.5.2(i)], but since $v_L(f-b_0) \notin \Gamma $ , we have $v_L(s)>0$ . Thus $0<s<1$ . Similarly, $h<(g-b_0)^{\dagger }$ gives and $t^{\dagger }>0$ , so $v_M(t)<0$ and thus $t>1$ . Putting this together yields

$$\begin{align*}f\ =\ b_0+\phi s\ <\ b_0+\phi\ \qquad \text{and}\ \qquad b_0+\phi\ <\ b_0 + \phi t\ =\ g, \end{align*}$$

contradicting that f and g realize the same cut in K. The other case, that there is $h \in K$ with $(f-b_0)^{\dagger }>h$ and $h>(g-b_0)^{\dagger }$ , is symmetric.

The next embedding lemma follows from Lemma 4.14 in the same way that [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 16.1.5] follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 16.1.4], without modification. The idea is to use Lemma 4.14 to carry out the proof of Lemma 4.13 in M and thereby construct the desired embedding.

Proposition 4.15. Let K, L, and f be as in Lemma 4.11 and suppose that $\Gamma ^<$ is cofinal in $\Gamma _{K\langle f \rangle }^<$ . Additionally, suppose that K and L are pre-H-fields and L is a pre-H-field extension of K with gap $0$ . Suppose that M is a pre-H-field extension of K with gap $0$ and $g \in M$ realizes the same cut in K as f. Then there exists an embedding $K \langle f \rangle \to M$ over K with $f \mapsto g$ .

Above we assumed that $\Gamma ^<$ was cofinal in $\Gamma _{K\langle f \rangle }^<$ , and now we turn to the other case, which does not occur in the H-field setting of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 16.1].

Lemma 4.16. Let K be a pre-H-field and L be a pre-H-field extension of K with gap $0$ . Let $f \in L^>$ with $\Gamma ^< < v_L(f) < 0$ . Suppose that M is a pre-H-field extension of K with gap $0$ and $g \in M^>$ satisfies $\Gamma ^< < v_M(g) < 0$ . Then there is an embedding $K\langle f \rangle \to M$ over K with $f \mapsto g$ .

Proof. Set and , and let . Then

$$\begin{align*}[\Gamma^{\neq}]> [\beta_0]>[\beta_1]>[\beta_2]>\cdots>[0]. \end{align*}$$

In particular, $[\beta _n] \notin [\Gamma ]$ for all n and $\beta _0, \beta _1, \beta _2, \dots $ are $\mathbb {Q}$ -linearly independent over $\Gamma $ . Hence the vector space $\mathbb {Q}\Gamma _{K\langle f \rangle }/\Gamma $ is infinite dimensional, so f is $\operatorname {d}$ -transcendental over K. Since $f=f_0$ and $f_n' = f_n f_{n+1}$ , induction yields $F_n \in K[Y_0, \dots , Y_n]$ of (total) degree at most $n+1$ with $f^{(n)}=F_n(f_0, \dots , f_n)$ . Let $P \in K\{Y\}^{\neq }$ of order at most r. Then $P(f)=\sum _{\boldsymbol i \in I} a_{\boldsymbol i} f_0^{i_0} \dots f_r^{i_r}$ , where I is a nonempty finite set of indices $\boldsymbol i=(i_0, \dots , i_r) \in \mathbb {N}^{1+r}$ . In particular, $\Gamma _{K\langle f \rangle } = \Gamma \oplus \bigoplus _{n} \mathbb {Z}\beta _n$ .

Set , , and . The same argument yields that g is $\operatorname {d}$ -transcendental over K and $P(g)=\sum _{\boldsymbol i \in I} a_{\boldsymbol i} g_0^{i_0} \dots g_r^{i_r}$ , where I is the same set of indices as in $P(f)$ and $a_{\boldsymbol i}$ are the same coefficients. Hence $\Gamma _{K\langle g \rangle } = \Gamma \oplus \bigoplus _{n} \mathbb {Z}\beta _n^{*}$ . Thus we have an isomorphism of ordered abelian groups $j \colon \Gamma _{K\langle f \rangle } \to \Gamma _{K\langle g \rangle }$ with $\beta _n \mapsto \beta _n^{*}$ . By the expressions for $P(f)$ and $P(g)$ , we have $j\big (v_L(P(f))\big ) = v_M(P(g))$ , which yields a valued differential field embedding from $K\langle f \rangle \to M$ over K with $f \mapsto g$ . To see that this is an ordered valued differential field embedding, note that $f_n>0$ and $g_n>0$ for all n, so $P(f)>0 \iff P(g)>0$ .

5 Extending the constant field

The results of this short section are used mainly to strengthen the statement of Theorem 6.15 in the next section and play no role in the main results of Section 7. The term “residue constant closed” defined in the next paragraph, however, is used in some key lemmas of Section 6.

Assumption. In this section, K is asymptotic with small derivation.

Since $C \subseteq \mathcal O$ , C maps injectively into $\boldsymbol k$ under the residue field map, and hence into $C_{\boldsymbol k}$ . We say that K is residue constant closed if K is henselian and C maps onto $C_{\boldsymbol k}$ , that is, $\operatorname {\mathrm {res}}(C)=C_{\boldsymbol k}$ . We say that L is a residue constant closure of K if it is a residue constant closed H-asymptotic extension of K with small derivation that embeds uniquely over K into every residue constant closed H-asymptotic extension M of K with small derivation. Note that if K has a residue constant closure, then it is unique up to unique isomorphism over K.

Proposition 5.1. Suppose that K is pre- $\operatorname {d}$ -valued of H-type with $\sup \Psi =0$ . Then K has a residue constant closure that is an immediate extension of K.

Proof. Recall from Section 4.1 that $\sup \Psi =0 \iff (\Gamma ^>)'=\Gamma ^>$ . Also note that if L is an immediate asymptotic extension of K, then it is H-asymptotic, satisfies $\Psi _L=\Psi $ (in particular, $\sup \Psi _L=0$ ), and is pre- $\operatorname {d}$ -valued by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 10.1.17].

Build a tower of immediate asymptotic extensions of K as follows. Set . If $K_\lambda $ is not henselian, set , the henselization of $K_\lambda $ , which as an algebraic extension of $K_\lambda $ is asymptotic [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 9.5.3]. If $K_\lambda $ is residue constant closed, we are done, so suppose that $K_\lambda $ is henselian but not residue constant closed. Then we have $u \in K_\lambda $ with $u \asymp 1$ , $u'\prec 1$ , and . Let y be transcendental over $K_\lambda $ and equip with the unique derivation extending that of $K_\lambda $ such that $y'=u'$ . Since $\sup \Psi _{K_{\lambda }}=0$ , we have $v(u') \in (\Gamma ^>)'$ , so the set has no maximum [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.2.5(iii)]. Thus by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.2.4] we can equip $K_{\lambda +1}$ with the unique valuation making it an H-asymptotic extension of $K_\lambda $ with $y \not \asymp 1$ ; with this valuation, $y \prec 1$ and $K_{\lambda +1}$ is an immediate extension of $K_{\lambda }$ . If $\lambda $ is a limit ordinal, set . Since each extension is immediate, by Zorn’s lemma we may take a maximal such tower $(K_\lambda )_{\lambda \leqslant \mu }$ .

It is clear that $K_\mu $ is residue constant closed, and we show that it also has the desired universal property. Let M be an H-asymptotic extension of K with small derivation that is residue constant closed, and let $\lambda <\mu $ and $i \colon K_{\lambda } \to M$ be an embedding. It suffices by induction to extend i uniquely to an embedding $K_{\lambda +1} \to M$ . If $K_{\lambda +1}=K_{\lambda }^{\operatorname {\mathrm {h}}}$ , then this follows from the universal property of the henselization. Now suppose that $K_{\lambda +1} = K_\lambda (y)$ with y and u as above. Take the unique $c \in C_M$ with $c \sim i(u)$ and set . Then $z'=i(u)'$ and $z \prec 1$ , so by the remarks after [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.2.4], z is transcendental over $i(K_\lambda )$ , and thus mapping $y \mapsto z$ yields a differential field embedding $K_{\lambda +1} \to M$ extending i. By the uniqueness of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.2.4], this is a valued differential field embedding. Finally, if i extends to an embedding with $y \mapsto z_1 \in M$ , then $i(u)-z_1 \in C_M$ and $i(u)-z_1 \sim i(u) \sim c$ , so $z_1=z$ .

Note that if K is a pre-H-field with $\sup \Psi =0$ , then as an immediate extension of K any residue constant closure of K embeds uniquely, as an ordered valued differential field, over K into every residue constant closed pre-H-field extension of K with small derivation.

The following result is worth recording but not needed later. In it, we construe an algebraic extension of a residue constant closed K as a valued differential field extension of K by equipping it with the unique derivation and valuation extending those of K (uniqueness of the valuation follows from the henselianity of K); if K is additionally an ordered field, then by henselianity $\mathcal O$ is necessarily convex.

Lemma 5.2. Suppose that K is residue constant closed. An algebraic extension L of K remains residue constant closed.

Proof. First, recall that an algebraic extension of a henselian valued field is henselian, and note that L is asymptotic [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 9.5.3] and has small derivation [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 6.2.1]. Let $u \in L$ with $u \asymp 1$ and $u' \prec 1$ ; we need to show that there is $c \in C_{L}$ with $c \sim u$ . Since $\overline {u} \in C_{\operatorname {\mathrm {res}}(L)}$ is algebraic over $\operatorname {\mathrm {res}}(K)$ , it is algebraic over $C_{\operatorname {\mathrm {res}}(K)}$ . Take a $P \in C[X]$ such that $\overline {P} \in C_{\operatorname {\mathrm {res}}(K)}[X]$ is the minimum polynomial of $\overline {u}$ over $C_{\operatorname {\mathrm {res}}(K)}$ . Then $P = Q\cdot \prod _{i=1}^n (X-c_i)$ , where $c_i \in C_{L}$ for $i=1,\dots ,n$ and $Q \in C_{L}[X]$ has no roots in $C_{L}$ . Moreover, Q has no roots in L, so henselianity yields i with $1 \leqslant i \leqslant n$ and $\overline {c_i}=\overline {u}$ .

6 Differential-Hensel–Liouville closures

Assumption. In this section, K is a pre-H-field.

In this section we construct differential-Hensel–Liouville closures (Theorem 6.16) in analogy with the Newton–Liouville closures of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 14.5] and prove that they are unique (Corollary 6.18). To do this, we first construct extensions that are real closed, have exponential integration, and satisfy an embedding property (Corollary 6.11, Lemma 6.14), in analogy with the Liouville closures of [Reference Aschenbrenner and van den Dries4]; however, we use [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 10.6] as a reference and also adapt some preliminaries from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Sections 10.4–10.6]. Combining this with the residue constant closures from the previous section, we record in Theorem 6.15 an improvement that is not needed later.

Suppose momentarily that K has gap $0$ . In several results below, it is assumed that $\boldsymbol k$ has exponential integration. It is worth noting that if K has exponential integration, then so does $\boldsymbol k$ . Hence, the assumption that $\boldsymbol k$ has exponential integration is necessary for K to have an extension with exponential integration and the same ordered differential residue field.

6.1 Adjoining exponential integrals

Suppose that $s \in K \setminus (K^{\times })^{\dagger }$ and f is transcendental over K. We give $K(f)$ the unique derivation extending that of K with $f^{\dagger } = s$ . In the first lemma, K need only be an ordered differential field.

Lemma 6.1. If K is real closed and $K(f)$ can be ordered making it an ordered field extension of K, then $C_{K(f)} = C$ .

Proof. Suppose towards a contradiction that we have $y \in K^{\times }$ and $m \geqslant 1$ with $y^{\dagger } = ms$ . Then by taking a $z \in K^{\times }$ with $z^m=y$ , after arranging $y>0$ if necessary, we have $z^{\dagger }=s$ , a contradiction. Thus [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 4.6.11] (with $R=K$ , $r=s$ , and $x=f$ ) and [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 4.6.12] (with $R=K[f, f^{-1}]$ ) yield that $C_{K(f)}$ is algebraic over C, so $C_{K(f)} = C$ .

In the next two lemmas, K is just a valued differential field, and need not be ordered. The first is based on [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.4.2].

Lemma 6.2. Suppose that K has small derivation and $\boldsymbol k = (\boldsymbol k^{\times })^{\dagger }$ . Let $K(f)$ have a valuation that makes it an extension of K with $\Gamma _{K(f)} = \Gamma $ and . Then $s-a^{\dagger } \prec 1$ for some $a \in K^{\times }$ .

Proof. Since $vf \in \Gamma $ , there is $b \in K^{\times }$ with . Then $s-b^{\dagger } = g^{\dagger } \asymp g' \preccurlyeq 1$ . If $s-b^{\dagger } \prec 1$ , set . If $s-b^{\dagger } \asymp 1$ , since $\boldsymbol k = (\boldsymbol k^{\times })^{\dagger }$ , we have $u \asymp 1$ in $K^{\times }$ with $s-b^{\dagger } \sim u^{\dagger }$ . Then set .

The last part of the previous argument also yields the following useful fact.

Lemma 6.3. Suppose that K has small derivation and $\boldsymbol k = (\boldsymbol k^{\times })^{\dagger }$ . If $s-a^{\dagger } \succcurlyeq 1$ for all $a \in K^{\times }$ , then $s-a^{\dagger } \succ 1$ for all $a \in K^{\times }$ .

Now we return to the situation that K is a pre-H-field.

Lemma 6.4 [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.5.18]

Suppose that K is henselian and $vs \in (\Gamma ^>)'$ . Then there is a unique valuation on $K(f)$ making it an H-asymptotic extension of K with $f \sim 1$ . With this valuation, $K(f)$ is an immediate extension of K, so there is a unique ordering of $K(f)$ making it a pre-H-field extension of K.

Here is a pre-H-field version of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.5.20] with the same proof; the assumption there that K is an H-field (rather than a pre-H-field) is only used at the end of the proof in showing that L too is an H-field. Recall that $\Psi ^{\downarrow }$ is the downward closure of $\Psi $ in $\Gamma $ .

Lemma 6.5. Suppose that K is real closed, $s<0$ , and $v(s-a^{\dagger }) \in \Psi ^{\downarrow }$ for all ${a \in K^{\times }}$ . Then there is a unique pair of a field ordering and a valuation on making it a pre-H-field extension of K with $f>0$ . Moreover, we have:

  1. (i) $vf \notin \Gamma $ , $\Gamma _L = \Gamma \oplus \mathbb {Z} vf$ , $f \prec 1$ ;

  2. (ii) $\Psi $ is cofinal in ;

  3. (iii) a gap in K remains a gap in L;

  4. (iv) if L has a gap not in $\Gamma $ , then $[\Gamma _L]=[\Gamma ]$ ;

  5. (v) $\boldsymbol k_L = \boldsymbol k$ .

6.2 Exponential integration closures

Let E be a differential field. We call a differential field extension F of E an exponential integration extension of E (expint-extension for short) if $C_F$ is algebraic over $C_E$ and for every $a \in F$ there are $t_1, \dots , t_n \in F^{\times }$ with $a \in E(t_1, \dots , t_n)$ such that for $i=1,\dots ,n$ , either $t_i$ is algebraic over $E(t_1,\dots ,t_{i-1})$ or $t_i^{\dagger } \in E(t_1,\dots ,t_{i-1})$ . In particular, any expint-extension is $\operatorname {d}$ -algebraic. The following is routine.

Lemma 6.6. Let $E \subseteq F \subseteq M$ be a chain of differential field extensions.

  1. (i) If M is an expint-extension of E, then M is an expint-extension of F;

  2. (ii) If M is an expint-extension of F and F is an expint-extension of E, then M is an expint-extension of E.

Lemma 6.7. If F is an expint-extension of E, then $|F|=|E|$ .

Proof. Supposing that F is an expint-extension of E, define an increasing sequence of differential subfields of F by setting and $E_{n+1}$ to be the algebraic closure of $E_n$ in F when n is even and when n is odd. Clearly, $F = \bigcup _n E_n$ , so it remains to check by induction that $|E_n|=|E|$ for all n.

Now suppose that E is an ordered differential field. We call E exponential integration closed (expint-closed for short) if it is real closed and has exponential integration. We call an ordered differential field extension F of E an exponential integration closure (expint-closure for short) of E if it is an expint-extension of E that is expint-closed. Note that in this latter definition, “closure” does not include a universal property, unlike for instance the gap-closures of Section 4.1.

The next observation has the same proof as [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.6.9].

Lemma 6.8. If E is expint-closed, then E has no proper expint-extension with the same constants as E.

Assumption. For the rest of this subsection, suppose that K has gap $0$ .

From this assumption it follows that $(\Gamma ^>)'=\Gamma ^>$ and $\Psi ^{\downarrow } = \Gamma ^<$ (see Section 4.1).

Below, we construe the real closure $K^{\operatorname {rc}}$ of K as an ordered valued differential field by equipping it with the unique derivation extending that of K and the unique valuation extending that of K whose valuation ring is convex (see for example [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 3.5.18]). Then $\Gamma _{K^{\operatorname {rc}}}$ is the divisible hull $\mathbb {Q}\Gamma $ of $\Gamma $ , $\boldsymbol k_{K^{\operatorname {rc}}}$ is the real closure of $\boldsymbol k$ , and $C_{K^{\operatorname {rc}}}$ is the real closure of C. Since K is a pre-H-field, so is $K^{\operatorname {rc}}$ by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 10.5.4], and it also has gap $0$ (see Section 4.1 for details on the extensions of the asymptotic couples).

Definition. We call a strictly increasing chain $(K_\lambda )_{\lambda \leqslant \mu }$ of pre-H-fields with gap $0$ an expint-tower on K if:

  1. (1) $K_0=K$ ;

  2. (2) if $\lambda $ is a limit ordinal, then $K_{\lambda } = \bigcup _{\rho <\lambda } K_\rho $ ;

  3. (3) if $\lambda <\lambda +1\leqslant \mu $ , then either:

    1. (a) $K_\lambda $ is not real closed and $K_{\lambda +1}$ is the real closure of $K_\lambda $ ; or

    2. (b) $K_\lambda $ is real closed and $K_{\lambda +1}=K_{\lambda }(y_\lambda )$ with $y_\lambda \notin K_\lambda $ satisfying either:

      1. (b1) $y_\lambda ^{\dagger } = s_\lambda \in K_\lambda $ with $y_\lambda \sim 1$ , $s_\lambda \prec 1$ , and $s_\lambda \neq a^{\dagger }$ for all $a \in K_\lambda ^{\times }$ ; or

      2. (b2) $y_\lambda ^{\dagger } = s_\lambda \in K_\lambda $ with $s_\lambda <0$ , $y_\lambda>0$ , and $s_\lambda -a^{\dagger } \succ 1$ for all $a \in K_\lambda ^{\times }$ .

Given such a tower, we call $K_{\mu }$ its top and set and for $\lambda \leqslant \mu $ .

Lemma 6.9. Let $(K_\lambda )_{\lambda \leqslant \mu }$ be an expint-tower on K. Then:

  1. (i) $K_\mu $ is an expint-extension of K;

  2. (ii) $C_\mu $ is the real closure of C if $\mu>0$ ;

  3. (iii) $\boldsymbol k_\mu $ is the real closure of $\boldsymbol k$ if $\mu>0$ ;

  4. (iv) $|K_\lambda |=|K|$ , and hence $\mu <|K|^+$ .

Proof. For (i), go by induction on $\lambda \leqslant \mu $ . The main thing to check is the condition on the constant fields. If $\lambda =0$ or $\lambda $ is a limit ordinal, this is clear. If $K_{\lambda +1}$ is the real closure of $K_\lambda $ , then $C_{\lambda +1}$ is the real closure of $C_{\lambda }$ . If $K_\lambda $ is real closed and $K_{\lambda +1}$ is as in (b) above, then $C_{\lambda +1}=C_{\lambda }$ by Lemma 6.1.

For (ii), $C_1$ is the real closure of C, and then $C_\lambda =C_1$ for all $\lambda \geqslant 1$ as in the proof of (i).

For (iii), $\boldsymbol k_1$ is the real closure of $\boldsymbol k$ , and then $\boldsymbol k_\lambda = \boldsymbol k_1$ for all $\lambda \geqslant 1$ ; in case (a) this is obvious, in case (b1) Lemma 6.4 applies, and in case (b2) Lemma 6.5 applies.

Finally, (iv) follows from (i) and Lemma 6.7.

Lemma 6.10. Let L be the top of a maximal expint-tower on K such that $\boldsymbol k_L$ has exponential integration. Then L is an expint-closure of K.

Proof. Suppose that L is not expint-closed. If L is not real closed, then its real closure is a proper pre-H-field extension of L with gap $0$ , so we could extend the expint-tower. We are left with the case that L is real closed and we have $s \in L \setminus (L^{\times })^{\dagger }$ . In particular, L is henselian. We may assume that $s<0$ . Take f transcendental over L with $f^{\dagger } = s$ .

First suppose that $s-a^{\dagger } \prec 1$ for some $a \in L^{\times }$ . Then taking such an a and replacing f and s by $f/a$ and $s-a^{\dagger }$ , we arrange that $s \prec 1$ . Giving $L(f)$ the valuation and ordering from Lemma 6.4 makes it a pre-H-field extension of L with gap $0$ of type (b1).

Now suppose that $s-a^{\dagger } \succcurlyeq 1$ for all $a \in L^{\times }$ . By Lemma 6.3, $s-a^{\dagger } \succ 1$ for all $a \in L^{\times }$ . Then giving $L(f)$ the ordering and valuation from Lemma 6.5 makes it a pre-H-field extension of L with gap $0$ of type (b2).

Thus L is expint-closed, and hence an expint-closure of K by Lemma 6.9(i).

Recall from Section 5 the term “residue constant closed.” In a few arguments below, this property roughly plays the same role as “ $\operatorname {d}$ -valued” does in the construction of Liouville closures in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 10.6].

Corollary 6.11. Suppose that $\boldsymbol k$ is expint-closed. Then K has an expint-closure that is a pre-H-field extension of K with gap $0$ . If K is residue constant closed, then K has a residue constant closed expint-closure that is a pre-H-field extension of K with gap $0$ .

Proof. By Lemma 6.9(iv), Zorn gives a maximal expint-tower $(K_\lambda )_{\lambda \leqslant \mu }$ on K. Then $\boldsymbol k_\mu = \boldsymbol k$ by Lemma 6.9(iii), so $C_{\boldsymbol k_\mu }=C_{\boldsymbol k}$ , and $C_{\mu }=C$ by Lemma 6.9(ii). Hence $K_\mu $ is an expint-closure of K by Lemma 6.10. If K is residue constant closed, then so is $K_\mu $ since it is real closed, so henselian, and $C=C_{\mu }$ maps onto $C_{\boldsymbol k} = C_{\boldsymbol k_\mu }$ .

Lemma 6.12. Let M be a residue constant closed, expint-closed pre-H-field extension of K with gap $0$ . Suppose that $(K_\lambda )_{\lambda \leqslant \mu }$ is an expint-tower on K in M (i.e., each $K_\lambda $ is a pre-H-subfield of M) and maximal in M (i.e., it cannot be extended to an expint-tower $(K_\lambda )_{\lambda \leqslant \mu +1}$ on K in M) such that $\boldsymbol k_\mu $ has exponential integration. Then $(K_\lambda )_{\lambda \leqslant \mu }$ is a maximal expint-tower on K.

Proof. Since M is real closed, $K_\mu $ must be real closed by maximality in M. So supposing $(K_\lambda )_{\lambda \leqslant \mu }$ is not a maximal expint-tower on K, we have $s_\mu \in K_\mu $ such that $s_\mu \neq a^{\dagger }$ for all $a \in K_\mu ^{\times }$ ; we may assume that $s_\mu <0$ . Since M is expint-closed, we have $y_\mu \in M$ with $y_\mu ^{\dagger }=s_\mu $ ; we may assume that $y_\mu>0$ .

First suppose that $s_\mu -a^{\dagger } \succcurlyeq 1$ for all $a \in K_\mu ^{\times }$ , so actually $s_\mu -a^{\dagger } \succ 1$ for all $a \in K_\mu ^{\times }$ by Lemma 6.3. Thus setting yields an extension of $(K_\lambda )_{\lambda \leqslant \mu }$ in M of type (b2).

Now suppose that $s_\mu -a^{\dagger } \prec 1$ for some $a \in K_\mu ^{\times }$ . Taking such an a and replacing $s_\mu $ and $y_\mu $ by $s_\mu -a^{\dagger }$ and $y_\mu /a$ , we may assume that $s_\mu \prec 1$ . Since M has gap $0$ , we have $y_\mu \asymp 1$ and so $y_\mu ' \asymp s_\mu \prec 1$ . That is, $\overline {y_\mu } \in C_{\operatorname {\mathrm {res}}(M)}$ , so we have $c \in C_M$ with $y_\mu \sim c$ . Replacing $y_\mu $ by $y_\mu /c$ , we obtain the desired extension of $(K_\lambda )_{\lambda \leqslant \mu }$ in M of type (b1).

This comment is not used later, but in the above lemma, we can replace the assumption that M is residue constant closed (so $C_{\operatorname {\mathrm {res}}(M)} = \operatorname {\mathrm {res}}(C_{M})$ ) with $C_{\operatorname {\mathrm {res}}(M)} = C_{\operatorname {\mathrm {res}}(K)}$ . In the final argument, instead of $c \in C_M$ we have $u \in K$ with $u \asymp 1$ and $u' \prec 1$ , so we also replace $s_\mu $ with $s_\mu -u^{\dagger }$ .

Corollary 6.13. Suppose that L is an expint-closed pre-H-field extension of K.

  1. (i) If L is an expint-closure of K, then no proper differential subfield of L containing K is expint-closed.

  2. (ii) Suppose that $\boldsymbol k$ is expint-closed, and that L has gap $0$ and is residue constant closed. If no proper differential subfield of L containing K is expint-closed, then L is an expint-closure of K.

Proof. For (i), if L is an expint-closure of K, then no proper differential subfield of L containing K is expint-closed by Lemmas 6.6 and 6.8.

For (ii), suppose that no proper differential subfield of L containing K is expint-closed. Take an expint-tower $(K_\lambda )_{\lambda \leqslant \mu }$ on K in L that is maximal in L. Since $\boldsymbol k$ is real closed, $\boldsymbol k_\mu =\boldsymbol k$ , and hence $\boldsymbol k_\mu $ has exponential integration. Then $(K_\lambda )_{\lambda \leqslant \mu }$ is a maximal expint-tower on K by Lemma 6.12. By Lemma 6.10, $K_\mu $ is expint-closed and hence equal to L.

Lemma 6.14. Let $(K_\lambda )_{\lambda \leqslant \mu }$ be an expint-tower on K. Then any embedding of K into a residue constant closed, expint-closed pre-H-field extension M of K with gap $0$ extends to an embedding of $K_\mu $ . If K is residue constant closed and $\boldsymbol k$ is expint-closed, then any two residue constant closed expint-closures of K that are pre-H-field extensions of K with gap $0$ are isomorphic over K.

Proof. Let M be a residue constant closed, expint-closed pre-H-field with gap $0$ . We prove that for $\lambda <\mu $ any embedding $K_\lambda \to M$ extends to an embedding ${K_{\lambda +1} \to M}$ , which yields the result by induction. Suppose that $i \colon K_\lambda \to M$ is an embedding. If $K_{\lambda +1}$ is the real closure of $K_\lambda $ , then we may extend i to $K_{\lambda +1}$ .

So suppose that $K_\lambda $ is real closed and we have $s_\lambda \in K_{\lambda }$ and $y_\lambda \in K_{\lambda +1}\setminus K_{\lambda }$ with $K_{\lambda +1}=K_{\lambda }(y_\lambda )$ , $y_\lambda ^{\dagger } = s_\lambda $ , $y_\lambda \sim 1$ , $s_\lambda \prec 1$ , and $s_\lambda \neq a^{\dagger }$ for all $a \in K_\lambda ^{\times }$ . Take $z \in M$ with $z^{\dagger } = i(s_\lambda )$ . Hence $z \asymp 1$ and $\overline {z} \in C_{\operatorname {\mathrm {res}}(M)}$ , so we have $c \in C_M$ with $z \sim c$ . By the uniqueness of Lemma 6.4, we may extend i to an embedding of $K_\lambda (y_\lambda )$ into M sending $y_\lambda $ to $z/c$ .

Now suppose that $K_\lambda $ is real closed and we have $s_\lambda \in K_\lambda $ and $y_\lambda \in K_{\lambda +1}\setminus K_\lambda $ with $K_{\lambda +1}=K_{\lambda }(y_\lambda )$ , $y_\lambda ^{\dagger } = s_\lambda $ , $s_\lambda <0$ , $y_\lambda>0$ , and $s_\lambda -a^{\dagger } \succ 1$ for all $a \in K^{\times }_\lambda $ . Take $z \in M$ with $z^{\dagger } = i(s_\lambda )$ ; we may assume that $z>0$ . Then by the uniqueness of Lemma 6.5, we can extend i to an embedding of $K_\lambda (y_\lambda )$ into M sending $y_\lambda $ to z.

The second statement follows from the first by Corollary 6.13(i) and the proof of Corollary 6.11.

Combining these results with Proposition 5.1 yields the following theorem, although for the construction of differential-Hensel–Liouville closures in Theorem 6.16, Corollary 6.11 and Lemma 6.14 already suffice.

Theorem 6.15. Suppose that $\boldsymbol k$ is expint-closed. Then K has a pre-H-field extension L with gap $0$ such that:

  1. (i) L is a residue constant closed, expint-closed extension of K;

  2. (ii) L embeds over K into any residue constant closed, expint-closed pre-H-field extension of K with gap $0$ ;

  3. (iii) L has no proper differential subfield containing K that is residue constant closed and expint-closed.

Proof. By Proposition 5.1, let $K_0$ be the residue constant closure of K. Taking the top of a maximal expint-tower on $K_0$ yields an expint-closure L of $K_0$ that is residue constant closed as in Corollary 6.11. For (ii), let M be a pre-H-field extension of K with gap $0$ that is residue constant closed and expint-closed. Then $K_0$ embeds uniquely into M over K, so by Lemma 6.14 we can extend this to an embedding of L. For (iii), suppose that $L_0 \supseteq K$ is a differential subfield of L that is residue constant closed and expint-closed. Then $L_0 \supseteq K_0$ , and hence $L_0=L$ by Corollary 6.13(i).

Let L be as above. Then any pre-H-field extension of K with gap $0$ satisfying (i) and (ii) is isomorphic to L over K by (iii). Also, L is a Liouville extensionFootnote 2 of K, since $K_0$ is a Liouville extension of K by construction and expint-extensions are Liouville extensions.

6.3 Differential-Hensel–Liouville closures

Assumption. We continue to assume in this subsection that the pre-H-field K has gap $0$ .

Definition. We call K differential-Hensel–Liouville closed (slightly shorter: $\operatorname {d}$ -Hensel–Liouville closed) if it is $\operatorname {d}$ -henselian and expint-closed. We call a pre-H-field extension L of K a differential-Hensel–Liouville closure (slightly shorter: $\operatorname {d}$ -Hensel–Liouville closure) of K if it is $\operatorname {d}$ -Hensel–Liouville closed and embeds over K into every $\operatorname {d}$ -Hensel–Liouville closed pre-H-field extension of K.

Note that if K is $\operatorname {d}$ -henselian, then K is also closed under integration in the sense that is surjective [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 7.1.8], hence the use of “Liouville” in the terms just defined.

To build $\operatorname {d}$ -Hensel–Liouville closures, we use the fact that if F is an asymptotic valued differential field with small derivation and linearly surjective differential residue field, then it has a (unique) differential-henselization ( $\operatorname {d}$ -henselization for short) $F^{\operatorname {dh}}$ by [Reference Pynn-Coates14, Theorem 3.7]. For such F, $F^{\operatorname {dh}}$ is an immediate asymptotic $\operatorname {d}$ -algebraic extension of F that is $\operatorname {d}$ -henselian and embeds over F into every $\operatorname {d}$ -henselian asymptotic extension of F; if F is a pre-H-field, then $F^{\operatorname {dh}}$ is too and embeds (as an ordered valued differential field) into every $\operatorname {d}$ -henselian pre-H-field extension of F.

Theorem 6.16. Suppose that $\boldsymbol k$ is expint-closed and linearly surjective. Then K has a $\operatorname {d}$ -Hensel–Liouville closure $K^{\operatorname {dhl}}$ .

Proof. We use below that any $\operatorname {d}$ -henselian asymptotic field is residue constant closed by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.4.10]. Define a sequence of pre-H-field extensions of K with gap $0$ as follows. Set . For $n \geqslant 1$ , if n is odd, let $K_n$ be the $\operatorname {d}$ -henselization of $K_{n-1}$ , and if n is even, let $K_n$ be the expint-closure of $K_{n-1}$ from Corollary 6.11. Note that $\boldsymbol k_{K_n} = \boldsymbol k$ for all n. We set and show that $K^{\operatorname {dhl}}$ is a $\operatorname {d}$ -Hensel–Liouville closure of K.

Let L be a pre-H-field extension of K that is $\operatorname {d}$ -henselian and expint-closed. We show by induction on n that we can extend any embedding $K_n \to L$ to an embedding $K_{n+1} \to L$ , so suppose that we have an embedding $i \colon K_n \to L$ . If n is even, then $K_{n+1}$ is the $\operatorname {d}$ -henselization of $K_n$ , so we may extend i to an embedding $K_{n+1} \to L$ . If n is odd, then $K_n$ is $\operatorname {d}$ -henselian and $K_{n+1}$ is the expint-closure of $K_n$ , so we can extend i to an embedding $K_{n+1} \to L$ by Lemma 6.14.

Note that $K^{\operatorname {dhl}}$ is a $\operatorname {d}$ -algebraic extension of K with the same residue field. In the next two results, adapted from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 16.2], we show that $K^{\operatorname {dhl}}$ is the unique, up to isomorphism over K, $\operatorname {d}$ -Hensel–Liouville closure of K.

Lemma 6.17. Suppose that $\boldsymbol k$ is expint-closed and linearly surjective. Let ${i \colon K^{\operatorname {dhl}} \to L}$ be an embedding into a pre-H-field L with gap $0$ such that $\operatorname {\mathrm {res}}\big (i(K^{\operatorname {dhl}})\big )=\operatorname {\mathrm {res}}(L)$ . Then

Proof. We have $i(K^{\operatorname {dhl}}) \subseteq i(K)^{\operatorname {dalg}}$ since $K^{\operatorname {dhl}}$ is a $\operatorname {d}$ -algebraic extension of K. For the other direction, note that $i(K^{\operatorname {dhl}})$ is a $\operatorname {d}$ -henselian, expint-closed pre-H-subfield of $i(K)^{\operatorname {dalg}}$ , so $i(K^{\operatorname {dhl}})=i(K)^{\operatorname {dalg}}$ by Theorem 4.12.

Hence for K as in the lemma above, any $\operatorname {d}$ -algebraic extension of K that is a $\operatorname {d}$ -henselian, expint-closed pre-H-field with the same residue field as K is isomorphic to $K^{\operatorname {dhl}}$ over K, and is thus a $\operatorname {d}$ -Hensel–Liouville closure of K.

Corollary 6.18. Suppose that $\boldsymbol k$ is expint-closed and linearly surjective. Then $K^{\operatorname {dhl}}$ has no proper differential subfield containing K that is $\operatorname {d}$ -Hensel–Liouville closed. Thus any $\operatorname {d}$ -Hensel–Liouville closure of K is isomorphic to $K^{\operatorname {dhl}}$ over K.

Proof. If $L \supseteq K$ is a $\operatorname {d}$ -Hensel–Liouville closed differential subfield of $K^{\operatorname {dhl}}$ , then we have an embedding $i \colon K^{\operatorname {dhl}} \to L$ over K. Viewing this as an embedding into $K^{\operatorname {dhl}}$ , by Lemma 6.17 we have $K^{\operatorname {dhl}} = i(K^{\operatorname {dhl}})$ , so $K^{\operatorname {dhl}}=L$ .

7 Main results

Assumption. In this section, K is a pre-H-field with gap $0$ .

7.1 Two-sorted results

We first combine earlier results to establish the key embedding lemma. For an ordered set S we denote the cofinality of S by $\operatorname {\mathrm {cf}}(S)$ . In Case 2 of the next lemma, recall from Section 4.2 the brief discussion of pc-sequences and the notion “ $\operatorname {d}$ -algebraically maximal,” along with its connection to pc-sequences.

Lemma 7.1. Suppose that K is $\operatorname {d}$ -Hensel–Liouville closed, and let E be a pre-H-subfield of K with $\operatorname {\mathrm {res}}(E) = \operatorname {\mathrm {res}}(K)$ . Let L be a $\operatorname {d}$ -Hensel–Liouville closed pre-H-field such that L is $|K|^+$ -saturated as an ordered set and $\operatorname {\mathrm {cf}}(\Gamma _L^<)>|\Gamma |$ . Then any embedding $E \to L$ can be extended to an embedding $K \to L$ .

Proof. Let $i \colon E \to L$ be an embedding. We may assume that $E \neq K$ . It suffices to show that i can be extended to an embedding $F \to L$ for some pre-H-subfield F of K properly containing E.

First, suppose that $\Gamma _E^<$ is not cofinal in $\Gamma ^<$ and let $f \in K^>$ with $\Gamma _E^< < vf < 0$ . By the cofinality assumption on $\Gamma _L^<$ , take $g \in L^>$ with $\Gamma _{i(E)}^< < v_L(g) < 0$ . Then we extend i to an embedding $E \langle f \rangle \to L$ sending $f \mapsto g$ by Lemma 4.16.

Now suppose that $\Gamma _E^<$ is cofinal in $\Gamma ^<$ and consider the following three cases.

Case 1: E is not $\operatorname {d}$ -Hensel–Liouville closed. From the assumptions on K, we get that $\operatorname {\mathrm {res}}(K)$ is expint-closed and linearly surjective. Since $\operatorname {\mathrm {res}}(E) = \operatorname {\mathrm {res}}(K)$ , we may extend i to an embedding of the $\operatorname {d}$ -Hensel–Liouville closure of E into L by Theorem 6.16.

Case 2: E is $\operatorname {d}$ -Hensel–Liouville closed and $E\langle y \rangle $ is an immediate extension of E for some $y \in K \setminus E$ . Take such a y and let $(a_\rho )$ be a divergent pc-sequence in K with $a_\rho \rightsquigarrow y$ . Since E is $\operatorname {d}$ -henselian, it is $\operatorname {d}$ -algebraically maximal [Reference Pynn-Coates14, Theorem 3.6], and so $(a_\rho )$ is of $\operatorname {d}$ -transcendental type over E. By the saturation assumption on L and [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 2.4.2], we have $z \in L$ with $i(a_\rho ) \rightsquigarrow z$ . Then [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 6.9.1] yields a valued differential field embedding $E\langle y \rangle \to L$ sending $y \mapsto z$ . Since $E\langle y \rangle $ is an immediate extension of E, this is also an ordered field embedding.

Case 3: E is $\operatorname {d}$ -Hensel–Liouville closed and there is no $y \in K \setminus E$ with $E\langle y \rangle $ an immediate extension of E. Take any $f \in K \setminus E$ . By saturation, take $g \in L$ such that for all $a \in E$ , we have

$$\begin{align*}a<f \implies i(a)<g \qquad \text{and} \qquad f<a \implies g<i(a). \end{align*}$$

Then we can extend i to an embedding $E\langle f \rangle \to L$ with $f \mapsto g$ by Proposition 4.15.

We now suspend the convention that $\boldsymbol k$ is the ordered differential residue field of K, and instead let $\boldsymbol k$ be an ordered differential field such that we have a surjective ordered differential ring homomorphism $\pi \colon \mathcal O \to \boldsymbol k$ ; note that $\pi $ has kernel , so it induces an ordered differential field isomorphism $\operatorname {\mathrm {res}}(K) \to \boldsymbol k$ . Thus we continue to think of $\boldsymbol k$ as the ordered differential residue field of K, but below possibly equipped with additional structure. We consider two-sorted structures of the form $\boldsymbol K = (K, \boldsymbol k; \pi _2)$ in the language $\mathcal L_2$ , where the language on the sort K is , the language on the sort $\boldsymbol k$ is , and $\pi _2 \colon K^2 \to \boldsymbol k$ is defined by:

  1. (1) $\pi _2(a, b) = \pi (ab^{-1})$ for $a \in K$ and $b \in K^{\times }$ with $a \preccurlyeq b$ ;

  2. (2) $\pi _2(a, b) = 0$ for $a,b \in K$ with $a \succ b$ ;

  3. (3) $\pi _2(0, 0) = 0$ .

Note that the dominance relation $\preccurlyeq $ can be axiomatized as a binary relation on K, without reference to v or $\Gamma $ (see for example [Reference Aschenbrenner, van den Dries and van der Hoeven5, Definition 3.1.1] and successive remarks). Also, the only reason for including a binary residue map here instead of the unary residue map is so we can avoid including multiplicative inversion in the language.

Fix an $\mathcal L_{\operatorname {\mathrm {res}}}$ -theory $T_{\operatorname {\mathrm {res}}}$ extending the theory of ordered differential fields that are linearly surjective and expint-closed, and let $T^{\operatorname {dhl}}$ be the $\mathcal L_2$ -theory whose models are structures $\boldsymbol K = (K, \boldsymbol k; \pi _2)$ as above such that K is $\operatorname {d}$ -Hensel–Liouville closed, $\Gamma \neq \{0\}$ , and $\boldsymbol k \models T_{\operatorname {\mathrm {res}}}$ .

7.1.1 Equivalence theorem

Let $\boldsymbol K = (K, \boldsymbol k; \pi _2)$ and $\boldsymbol K^{*} = (K^{*}, \boldsymbol k^{*}; \pi _2^{*})$ be models of $T^{\operatorname {dhl}}$ . We aim to construct a back-and-forth system from $\boldsymbol K$ to $\boldsymbol K^{*}$ (when $\boldsymbol K$ and $\boldsymbol K^{*}$ are sufficiently saturated). To that end, a good substructure of $\boldsymbol K$ is an $\mathcal L_2$ -substructure $\boldsymbol E = (E, \boldsymbol k_{\boldsymbol E})$ of $\boldsymbol K$ such that E and $\boldsymbol k_E$ are fields, where we drop the $\pi _2$ from the notation. Note that then E is a pre-H-subfield of K with gap $0$ and also that the restriction $\pi _2 \colon E^2 \to \boldsymbol k_E$ need not be surjective. Let $\boldsymbol E$ and $\boldsymbol E^{*}$ be good substructures of $\boldsymbol K$ and $\boldsymbol K^{*}$ , respectively. We say that a map $\boldsymbol f \colon \boldsymbol E \to \boldsymbol E^{*}$ is a good map if $\boldsymbol f = (f, f_{\operatorname {r}})$ is an $\mathcal L_2$ -isomorphism such that $f_{\operatorname {r}} \colon \boldsymbol k_{\boldsymbol E} \to \boldsymbol k_{\boldsymbol E^{*}}$ is moreover elementary as a partial map from $\boldsymbol k$ to $\boldsymbol k^{*}$ . In particular, $f \colon E \to E^{*}$ is a pre-H-field isomorphism and $f_{r} \colon \boldsymbol{k}_{\boldsymbol{E}} \to \boldsymbol{k}_{\boldsymbol{E}^*}$ is an $\mathcal L_{\operatorname {\mathrm {res}}}$ -isomorphism.

Theorem 7.2 (Equivalence theorem)

Every good map $\boldsymbol E \to \boldsymbol E^{*}$ is elementary as a partial map from $\boldsymbol K$ to $\boldsymbol K^{*}$ .

Proof. Let $\boldsymbol f = (f, f_{\operatorname {r}})$ be a good map from $\boldsymbol E$ to $\boldsymbol E^{*}$ . Let $\kappa $ be a cardinal of uncountable cofinality such that $\max \{|E|, |\boldsymbol k_{\boldsymbol E}|, |\mathcal L_{\operatorname {\mathrm {res}}}|\}<\kappa $ and $2^\lambda < \kappa $ for every cardinal $\lambda <\kappa $ . By passing to elementary extensions, we may suppose that $\boldsymbol K$ and $\boldsymbol K^{*}$ are $\kappa $ -saturated. We say a good substructure $(E_1, \boldsymbol k_1)$ of $\boldsymbol K$ is small if $\max \{|E_1|, |\boldsymbol k_1|\}<\kappa $ ; note that if $(E_1, \boldsymbol k_1)$ is small and $E_2 \subseteq K$ is a pre-H-field extension of $E_1$ with $\Gamma _{E_2}=\Gamma _{E_1}$ and $\pi _2(E_2, E_2) \subseteq \boldsymbol k_1$ , then $(E_2, \boldsymbol k_1)$ is small. To establish the theorem, it suffices to show that the set of good maps with small domain is a back-and-forth system from $\boldsymbol K$ to $\boldsymbol K^{*}$ .

Given $a \in K$ , we need to extend $\boldsymbol f$ to a good map with small domain $(F, \boldsymbol k_{\boldsymbol F})$ such that $a \in F$ . First, we note two basic procedures for extending $\boldsymbol k_{\boldsymbol E}$ and $\pi _2(E, E)$ :

  1. (1) Given $d \in \boldsymbol k$ , arranging that $d \in \boldsymbol k_{\boldsymbol E}$ : By the saturation assumption, we can extend $f_{\operatorname {r}}$ to a partial elementary map with d in its domain. Without changing f or E, this yields an extension of $\boldsymbol f$ to a good map with small domain.

  2. (2) Given $d \in \boldsymbol k_{\boldsymbol E} \setminus \pi _2(E, E)$ , arranging that $d \in \pi _2(E, E)$ : By Lemmas 3.2 and 3.3, we have $b \in K$ such that $\pi _2(E\langle b \rangle , E\langle b \rangle ) = \pi _2(E, E)\langle d \rangle $ and $\Gamma _{E\langle b \rangle } = \Gamma _E$ , and a good map $(g, f_{\operatorname {r}})$ extending $\boldsymbol f$ with small domain $(E\langle b \rangle , \boldsymbol k_{\boldsymbol E})$ .

In this proof only, we call an ordered differential field closed if it is real closed, linearly surjective, and has exponential integration. Consider $E \langle a \rangle $ . By (1), we extend $f_{\operatorname {r}}$ to a partial elementary map $f_{1, \operatorname {r}}$ with domain $\boldsymbol k_1$ such that $\boldsymbol k_1$ is closed, $\pi _2(E \langle a \rangle , E \langle a \rangle ) \subseteq \boldsymbol k_1$ , and $|\boldsymbol k_1|<\kappa $ . By (2), we extend f to $f_1$ so that $\boldsymbol f_1 = (f_1, f_{1, \operatorname {r}})$ is a good map with small domain $(E_1, \boldsymbol k_1)$ satisfying $\pi _2(E_1, E_1) = \boldsymbol k_1$ . In the same way, we extend $\boldsymbol f_1$ to a good map $\boldsymbol f_2 = (f_2, f_{2, \operatorname {r}})$ with small domain $(E_2, \boldsymbol k_2)$ such that $\boldsymbol k_2$ is closed and $\pi _2(E_1 \langle a \rangle , E_1 \langle a \rangle ) \subseteq \boldsymbol k_2 = \pi _2(E_2, E_2)$ . Iterating this procedure and taking unions yields a good map $\boldsymbol f_\omega = (f_\omega , f_{\omega , \operatorname {r}})$ with small domain $\boldsymbol E_\omega = (E_\omega , \boldsymbol k_\omega )$ such that $\boldsymbol k_\omega $ is closed and $\pi _2(E_\omega \langle a \rangle , E_\omega \langle a \rangle ) = \boldsymbol k_\omega = \pi _2(E_\omega , E_\omega )$ .

Now take the $\operatorname {d}$ -Hensel–Liouville closure $F \subseteq K$ of $E_\omega \langle a \rangle $ by Theorem 6.16. By construction, $\pi _2(F,F)=\boldsymbol k_\omega $ and $F = \bigcup _n F_n$ is a countable increasing union of pre-H-fields $F_n$ such that $F_0 = E_\omega \langle a \rangle $ and for all n, either $|F_{n+1}|=|F_n|$ or $F_{n+1}$ is an immediate extension of $F_n$ . It follows that $(F, \boldsymbol k_\omega )$ is small. Thus we may apply Lemma 7.1 to extend $f_\omega $ to $f_{\omega +1}$ so that $(f_{\omega +1}, f_{\omega , \operatorname {r}})$ is a good map with small domain $(F, \boldsymbol k_\omega )$ .

Corollary 7.3. We have $\boldsymbol K \equiv \boldsymbol K^{*}$ if and only if $\boldsymbol k \equiv \boldsymbol k^{*}$ .

Proof. The left-to-right direction is trivial, so suppose that $\boldsymbol k \equiv \boldsymbol k^{*}$ . We may assume that $\mathcal L_{\operatorname {\mathrm {res}}}$ is an expansion of by relation symbols, so then we can identify $\mathbb {Q}$ with an $\mathcal L_{\operatorname {\mathrm {res}}}$ -substructure of $\boldsymbol k$ and an $\mathcal L_{\operatorname {\mathrm {res}}}$ -substructure of $\boldsymbol k^{*}$ , respectively, and by assumption, these are isomorphic. Consider $(\mathbb {Q}, \mathbb {Q}; \pi _2)$ : the first $\mathbb {Q}$ is endowed with its usual ordered ring structure, the trivial derivation, and the trivial dominance relation $\preccurlyeq $ ; the second $\mathbb {Q}$ is an $\mathcal L_{\operatorname {\mathrm {res}}}$ -structure as above; and $\pi _2 \colon \mathbb {Q}^2 \to \mathbb {Q}$ is defined by, for $q_1, q_2 \in \mathbb {Q}$ , $\pi _2(q_1, q_2) = q_1q_2^{-1}$ if $q_2 \neq 0$ and ${\pi _2(q_1, 0) = 0}$ . This structure embeds into both $\boldsymbol K$ and $\boldsymbol K^{*}$ , inducing an obvious good map between good substructures, which is then elementary as a partial map from $\boldsymbol K$ to $\boldsymbol K^{*}$ by Theorem 7.2. Hence $\boldsymbol K \equiv \boldsymbol K^{*}$ .

Corollary 7.4. Let $\boldsymbol E = (E, \boldsymbol k_E; \pi _2) \subseteq \boldsymbol K$ with $\boldsymbol E \models T^{\operatorname {dhl}}$ and $\boldsymbol k_E \preccurlyeq \boldsymbol k$ . Then ${\boldsymbol E \preccurlyeq \boldsymbol K}$ .

Proof. View the identity map on $(E, \boldsymbol k_E)$ as a map from $\boldsymbol E$ to $\boldsymbol K$ and note that it is good.

In particular, if $T_{\operatorname {\mathrm {res}}}$ is model complete, then so is $T^{\operatorname {dhl}}$ . We now prove relative quantifier elimination, using much more of the strength of the Equivalence theorem than the previous corollaries did.

7.1.2 Relative quantifier elimination

Let $x = (x_1, \dots , x_m)$ be a tuple of distinct variables of sort K and $y = (y_1, \dots , y_n)$ be a tuple of distinct variables of sort $\boldsymbol k$ . We call an $\mathcal L_2$ -formula $\psi (x, y)$ special if $\psi (x, y)$ is

$$\begin{align*}\psi_{\operatorname{r}}\big(\pi_2(P_1(x), Q_1(x)), \dots, \pi_2(P_\ell(x), Q_\ell(x)), y\big), \end{align*}$$

for some $\ell \in \mathbb {N}$ , $\mathcal L_{\operatorname {\mathrm {res}}}$ -formula $\psi _{\operatorname {r}}(u_1, \dots , u_\ell, y)$ , and $P_1, Q_1, \dots , P_\ell, Q_\ell \in \mathbb {Z}\{X_1, \dots , X_m\}$ .

Theorem 7.5. Let $\phi (x, y)$ be an $\mathcal L_2$ -formula with x and y as above. Then $\phi (x, y)$ is $T^{\operatorname {dhl}}$ -equivalent to

(*) $$ \begin{align} \big(\theta_1(x) \wedge \psi_1(x,y)\big) \vee \dots \vee \big(\theta_N(x) \wedge \psi_N(x,y)\big) \end{align} $$

for some $N \in \mathbb {N}$ , quantifier-free $\mathcal L$ -formulas $\theta _1(x)$ , …, $\theta _N(x)$ , and special formulas $\psi _1(x,y)$ , …, $\psi _N(x,y)$ .

Proof. Let $\Theta (x, y)$ be the set of $\mathcal L_2$ -formulas displayed in (*). Then $\Theta (x, y)$ is obviously closed under disjunction and also closed under negation, up to logical equivalence. It suffices to show that every $(x,y)$ -type (consistent with $T^{\operatorname {dhl}}$ ) is determined by its intersection with $\Theta (x, y)$ . Below, $\theta (x)$ ranges over quantifier-free $\mathcal L$ -formulas and $\psi (x, y)$ ranges over special formulas. For a model $\boldsymbol K = (K, \boldsymbol k; \pi _2)$ of $T^{\operatorname {dhl}}$ and $a \in K^{m}$ and $d \in \boldsymbol k^{n}$ , we set

and

Let $\boldsymbol K = (K, \boldsymbol k; \pi _2)$ and $\boldsymbol K^{*} = (K^{*}, \boldsymbol k^{*}; \pi _2^{*})$ be models of $T^{\operatorname {dhl}}$ . Let $a \in K^{m}$ , $a^{*} \in (K^{*})^{m}$ , $d \in \boldsymbol k^{n}$ , and $d^{*} \in (\boldsymbol k^{*})^{n}$ , and assume that $\operatorname {\mathrm {qftp}}^K(a) = \operatorname {\mathrm {qftp}}^{K^{*}}(a^{*})$ and $\operatorname {\mathrm {tp}}_{\operatorname {r}}^{\boldsymbol K}(a, d) = \operatorname {\mathrm {tp}}_{\operatorname {r}}^{\boldsymbol K^{*}}(a^{*}, d^{*})$ . We need to show that $\operatorname {\mathrm {tp}}^{\boldsymbol K}(a, d) = \operatorname {\mathrm {tp}}^{\boldsymbol K^{*}}(a^{*}, d^{*})$ .

Let be the differential subfield of K generated by a, construed as a pre-H-subfield of K, and let $\boldsymbol k_{\boldsymbol E}$ be the $\mathcal L_{\operatorname {\mathrm {res}}}$ -substructure and subfield of $\boldsymbol k$ generated by $\pi _2(E,E)$ and d. Then $\boldsymbol E = (E, \boldsymbol k_{\boldsymbol E})$ is a good substructure of $\boldsymbol K$ . Defining $E^{*}$ and $\boldsymbol k_{\boldsymbol E^{*}}$ likewise yields a good substructure $\boldsymbol E^{*} = (E^{*}, \boldsymbol k_{\boldsymbol E^{*}})$ of $\boldsymbol K^{*}$ . Since we have $\operatorname {\mathrm {qftp}}^{K}(a) = \operatorname {\mathrm {qftp}}^{K^{*}}(a^{*})$ and $\operatorname {\mathrm {tp}}_{\operatorname {r}}^{\boldsymbol K}(a, d) = \operatorname {\mathrm {tp}}_{\operatorname {r}}^{\boldsymbol K^{*}}(a^{*}, d^{*})$ , the natural map $\boldsymbol E \to \boldsymbol E^{*}$ sending $a \mapsto a^{*}$ and $d \mapsto d^{*}$ is good, and hence $\operatorname {\mathrm {tp}}^{\boldsymbol K}(a, d) = \operatorname {\mathrm {tp}}^{\boldsymbol K^{*}}(a^{*}, d^{*})$ by Theorem 7.2.

Note that the outer $\mathcal L_{\operatorname {\mathrm {res}}}$ -formula in a special formula may have quantifiers, so the previous theorem eliminates quantifiers down to quantifiers over the sort $\boldsymbol k$ . In particular, if $T_{\operatorname {\mathrm {res}}}$ has quantifier elimination, then so does $T^{\operatorname {dhl}}$ . Also, $\boldsymbol k$ is purely stably embedded in $\boldsymbol K$ in the following sense.

Corollary 7.6. Let $\boldsymbol K = (K, \boldsymbol k; \pi _2) \models T^{\operatorname {dhl}}$ . Every subset of $\boldsymbol k^n$ definable in $\boldsymbol K$ (with parameters from $\boldsymbol K$ ) is definable in the $\mathcal L_{\operatorname {\mathrm {res}}}$ -structure $\boldsymbol k$ (with parameters from $\boldsymbol k$ ).

Let $T^{\prime }_{\operatorname {\mathrm {res}}} \subseteq T_{\operatorname {\mathrm {res}}}$ and let T be the $\mathcal L_2$ -theory of structures $(K, \boldsymbol k; \pi _2)$ such that K is a pre-H-field with gap $0$ , $\boldsymbol k \models T^{\prime }_{\operatorname {\mathrm {res}}}$ , and we no longer require $\pi \colon \mathcal O \to \boldsymbol k$ to be surjective.

Corollary 7.7. If $T_{\operatorname {\mathrm {res}}}$ is the model companion of $T^{\prime }_{\operatorname {\mathrm {res}}}$ , then $T^{\operatorname {dhl}}$ is the model companion of T. If $T_{\operatorname {\mathrm {res}}}$ additionally has quantifier elimination, then $T^{\operatorname {dhl}}$ is the model completion of T.

Proof. Suppose that every model of $T^{\prime }_{\operatorname {\mathrm {res}}}$ can be extended to a model of $T_{\operatorname {\mathrm {res}}}$ . By Corollary 7.4 and Theorem 7.5, it remains to show that every model of T can be extended to a model of $T^{\operatorname {dhl}}$ . Let $(K, \boldsymbol k; \pi _2) \models T$ . First extend $\boldsymbol k$ to $\boldsymbol k^{*} \models T_{\operatorname {\mathrm {res}}}$ and apply Corollary 3.4 to obtain a pre-H-field extension $K^{*}$ of K with gap $0$ and a surjective ordered differential ring homomorphism $\pi ^{*} \colon \mathcal O_{K^{*}} \to \boldsymbol k^{*}$ . Now by Theorem 6.16 extend $K^{*}$ to its $\operatorname {d}$ -Hensel–Liouville closure $(K^{*})^{\operatorname {dhl}}$ and $\pi ^{*}$ to $\pi ^{*} \colon \mathcal O_{(K^{*})^{\operatorname {dhl}}} \to \boldsymbol k^{*}$ , and note that $\big ((K^{*})^{\operatorname {dhl}}, \boldsymbol k^{*}; \pi _2^{*}\big ) \models T^{\operatorname {dhl}}$ .

The NIP transfer principle in Proposition 7.8 below follows from Theorem 7.5 and an analogous transfer principle from [Reference Bélair and Bousquet8, Reference Delon10] by the standard trick of “forgetting” the derivation. In those papers, it is shown that the three-sorted theory of henselian valued fields of equicharacteristic $0$ has NIP if the residue field does. More explicitly, let $(K, \boldsymbol k, \Gamma; \pi , v)$ be a valued field of equicharacteristic $0$ construed as a three-sorted structure. That is, K is a field in the language of rings, $\Gamma $ is an ordered abelian group, and $\boldsymbol k$ is a field of characteristic $0$ in the language of rings such that the surjective map $v \colon K^{\times } \to \Gamma $ makes K a henselian valued field with residue field (isomorphic to) $\boldsymbol k$ and residue field map $\pi \colon \mathcal O \to \boldsymbol k$ . The relevant NIP transfer principle from [Reference Bélair and Bousquet8, Reference Delon10] is that if $\boldsymbol k$ has NIP, then so does $(K, \boldsymbol k, \Gamma; \pi , v)$ . Moreover, although not explicitly stated there, the proof goes through when $\boldsymbol k$ has extra structure. It is not relevant here, but one can also expand $(K, \boldsymbol k, \Gamma; \pi , v)$ by an angular component map, as done in [Reference Bélair and Bousquet8].

What is the difference with our setup? We have $\boldsymbol K = (K, \boldsymbol k; \pi _2) \models T^{\operatorname {dhl}}$ , a two-sorted structure in which K is additionally equipped with a derivation, an ordering, and a dominance relation, and we have a binary rather than unary version of the residue field map. Although $\boldsymbol k$ too is equipped with additional structure, this is taken care of already, as mentioned above. Let $K_{\mathrm {ring}}$ be the reduct of K to the language of rings. Then the ordering, the dominance relation, and $\pi _2$ are definable without parameters in $(K_{\mathrm {ring}}, \boldsymbol k, \Gamma; \pi , v)$ : Since K is real closed, its ordering is definable in $K_{\mathrm {ring}}$ , and its dominance relation is definable from v and $\Gamma $ . Finally, $\pi _2$ is definable from $\pi $ . Thus, we must only deal with the derivation, as done in detail in the proposition.

Proposition 7.8. Let $\boldsymbol K = (K, \boldsymbol k; \pi _2) \models T^{\operatorname {dhl}}$ . If the $\mathcal L_{\operatorname {\mathrm {res}}}$ -structure $\boldsymbol k$ has NIP, then so does $\boldsymbol K$ .

Proof. Suppose that $\boldsymbol k$ has NIP. Recall that a boolean combination of formulas that have NIP has NIP, so by Theorem 7.5 it suffices to show that every quantifier-free $\mathcal L$ -formula has NIP and every special formula has NIP.

First, consider a special formula $\psi (x; y, z)$ , where y is an m-tuple of distinct variables of sort K, z is an n-tuple of distinct variables of sort $\boldsymbol k$ , and x is a single variable distinct from the variables in y and z, and suppose towards a contradiction that $\psi (x; y, z)$ is independent in $\boldsymbol K$ . As a reminder, this means that the family defined by $\psi (x; y, z)$ has infinite VC-dimension, where, by convention, the semicolon indicates that x is the object variable and $y,z$ are the parameter variables. That is, for every $N \in \mathbb {N}$ with $N \geqslant 1$ , there exist $a_1, \dots , a_N$ of the same sort as x such that, for all $I \subseteq \{1, \dots , N\}$ , there exist $b_I \in K^{m}$ and $d_I \in \boldsymbol k^{n}$ with:

$$\begin{align*}\boldsymbol K \models \psi(a_i; b_I, d_I)\ \iff\ i \in I. \end{align*}$$

First suppose that the variable x is of sort $\boldsymbol k$ . As a special formula, $\psi (x; y, z)$ is thus

$$\begin{align*}\psi_{\operatorname{r}}\big(x; \pi_2(P_1(y), Q_1(y)), \dots, \pi_2(P_\ell(y), Q_\ell(y)), z\big),\end{align*}$$

where $\ell \in \mathbb {N}$ , $P_1, Q_1, \dots , P_\ell, Q_\ell \in \mathbb {Z}\{Y_1, \dots , Y_m\}$ , and $\psi _{\operatorname {r}}(x; u_1, \dots , u_\ell, z)$ is an $\mathcal L_{\operatorname {\mathrm {res}}}$ -formula. Then $\psi _{\operatorname {r}}(x; u_1, \dots , u_\ell, z)$ itself is independent in $\boldsymbol k$ . To see this, let $N \geqslant 1$ and $I \subseteq \{1, \dots , N\}$ , and take $a_1, \dots , a_N \in \boldsymbol k$ , $b_I \in K^{m}$ , and $d_I \in \boldsymbol k^{n}$ as above. Then letting

we have

$$ \begin{align*} \boldsymbol k \models \psi_{\operatorname{r}}(a_i; e_I, d_I)\ &\iff\ \boldsymbol K \models \psi_{\operatorname{r}}\big(a_i; \pi_2(P_1(b_I), Q_1(b_I)), \dots, \pi_2(P_\ell(b_I), Q_\ell(b_I)), d_I\big)\\ &\iff\ i \in I. \end{align*} $$

Now suppose that x is of sort K, so then $\psi (x; y, z)$ is

$$\begin{align*}\psi_{\operatorname{r}}\big(\pi_2(P_1(x, y), Q_1(x, y)), \dots, \pi_2(P_\ell(x, y), Q_\ell(x, y)), z\big), \end{align*}$$

where $\ell \in \mathbb {N}$ , $P_1, Q_1, \dots , P_\ell, Q_\ell \in \mathbb {Z}\{X, Y_1, \dots , Y_m\}$ , and $\psi _{\operatorname {r}}(u_1, \dots , u_\ell, z)$ is an $\mathcal L_{\operatorname {\mathrm {res}}}$ -formula. For notational simplicity, we assume that $\ell=m=n=1$ and replace $P_1$ by P, $Q_1$ by Q, and $Y_1$ by Y. It is no longer obvious how to extract an independent $\mathcal L_{\operatorname {\mathrm {res}}}$ -formula from $\psi $ now that x and y both appear in the term $\pi _2(P(x,y),Q(x,y))$ , so for this case we instead appeal to the results of [Reference Bélair and Bousquet8, Reference Delon10] (this works too for the previous case, but the goal in separating the cases is to illustrate the reason for using [Reference Bélair and Bousquet8, Reference Delon10]). Now comes the “forgetting” of the derivation. Explicitly, take r and $p, q \in \mathbb {Z}[X_0, \dots , X_r, Y_{0}, \dots , Y_{r}]$ such that

$$\begin{align*}P(a, b) = p(a, a', \dots, a^{(r)}, b, b', \dots, b^{(r)})\quad \text{and}\quad Q(a, b) = q(a, a', \dots, a^{(r)}, b, b', \dots, b^{(r)})\end{align*}$$

for all $a, b \in K$ . Then for all $a,b \in K$ and $d \in \boldsymbol k$ , we have $\boldsymbol K \models \psi (a; b, d)$ if and only if

$$\begin{align*}(K_{\mathrm{ring}}, \boldsymbol k; \pi_2) \models \psi_{\operatorname{r}}\big(\pi_2(p(a, \dots, a^{(r)}, b, \dots, b^{(r)}), q(a, \dots, a^{(r)}, b, \dots, b^{(r)})), d\big), \end{align*}$$

where $K_{\mathrm {ring}}$ is the reduct of K to the language of rings (no change is made here to the sort $\boldsymbol k$ , which in particular retains its derivation). Letting $\varphi (x_0, \dots , x_r; y_0, \dots , y_r, z)$ be

$$\begin{align*}\psi_{\operatorname{r}}\big(\pi_2(p(x_0, \dots, x_r, y_0, \dots, y_r), q(x_0, \dots, x_r, y_0, \dots, y_r)), z\big), \end{align*}$$

$\varphi $ is independent in $(K_{\mathrm {ring}}, \boldsymbol k; \pi _2)$ , but since $\pi _2$ is definable in $(K_{\mathrm {ring}}, \boldsymbol k, \Gamma; \pi , v)$ , this yields an independent formula in $(K_{\mathrm {ring}}, \boldsymbol k, \Gamma; \pi , v)$ , contradicting [Reference Bélair and Bousquet8, Reference Delon10].

A similar argument shows that every quantifier-free $\mathcal L$ -formula has NIP; alternatively, reduce to the fact that RCVF, the theory of real closed fields with a nontrivial valuation whose valuation ring is convex, has NIP, bypassing [Reference Bélair and Bousquet8, Reference Delon10].

7.2 One-sorted results

Recall that the theory of closed ordered differential fields is the model completion of the theory of ordered differential fields, where no interaction is assumed between the ordering and the derivation. This theory has quantifier elimination and is complete [Reference Singer22]. We have the following results for $T^{\operatorname {dhl}}_{\operatorname {codf}}$ , the one-sorted theory of $\operatorname {d}$ -Hensel–Liouville closed pre-H-fields with nontrivial value group and closed ordered differential residue field.

Theorem 7.9. The theory $T^{\operatorname {dhl}}_{\operatorname {codf}}$ :

  1. (i) has quantifier elimination;

  2. (ii) is the model completion of the theory of pre-H-fields with gap $0$ ;

  3. (iii) is complete, and hence decidable;

  4. (iv) is distal, and hence has NIP;

  5. (v) is locally o-minimal.

Proof. Item (i) follows from a standard quantifier elimination test (see for example [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary B.11.9]), using quantifier elimination for closed ordered differential fields, Corollary 3.4, and Lemma 7.1.

For item (ii), argue as in Corollary 7.7 that every pre-H-field with gap $0$ extends to a model of $T^{\operatorname {dhl}}_{\operatorname {codf}}$ , using the definition of closed ordered differential fields to check that they are expint-closed and linearly surjective.

The completeness in item (iii) follows from (i) using that the ordered ring $(\mathbb {Z}; +, -,\cdot ,0,1,\leqslant )$ expanded by the trivial derivation and dominance relation embeds into every pre-H-field (or use Corollary 7.3). Decidability then follows, since $T^{\operatorname {dhl}}_{\operatorname {codf}}$ has a recursive axiomatization.

Item (iv) follows by [Reference Aschenbrenner, Chernikov, Gehret and Ziegler2, Proposition 7.1] from (i) and the fact that RCVF, the theory of real closed fields with a nontrivial valuation whose valuation ring is convex, is distal; in applying [Reference Aschenbrenner, Chernikov, Gehret and Ziegler2, Proposition 7.1], note that a derivation satisfies its assumptions, as shown implicitly in the proof of Proposition 7.8.

Item (v) is similar to [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 16.6.8], which is the statement of local o-minimality of $T^{\operatorname {nl}}$ (strictly speaking, it is equivalent to local o-minimality by taking fractional linear transformations). By (i) and compactness, it suffices to show that for $L \succcurlyeq K \models T^{\operatorname {dhl}}_{\operatorname {codf}}$ with $f,g>K$ in L, there is an isomorphism $K\langle f \rangle \to K\langle g \rangle $ over K, which is done as in Lemma 4.16. The only difference is that now $vf<\Gamma $ instead of $\Gamma ^< < vf < 0$ , so use that $\Psi =\Gamma ^<$ to get $vf_n<\Gamma $ for all n, where $f_n$ is as in the proof of Lemma 4.16.

8 Examples

In this section, we provide two examples of pre-H-fields with gap $0$ . The first, outlined in the introduction, is a model of $T^{\operatorname {dhl}}$ , with $T_{\operatorname {\mathrm {res}}} = T^{\operatorname {nl}}_{\operatorname {small}}$ , that is transexponential and whose residue field is exponentially bounded; we have continued the study of such structures in [Reference Pynn-Coates15, Reference Pynn-Coates16]. The second shows that the assumption of exponential integration is necessary in Corollary 7.3 and also that $T^{\operatorname {dhl}}$ is satisfiable for any $T_{\operatorname {\mathrm {res}}}$ extending the theory of ordered differential fields that are real closed, linearly surjective, and have exponential integration.

8.1 A transexponential pre-H-field

This subsection elaborates on [Reference Aschenbrenner, van den Dries and van der Hoeven5, Example 10.1.7]. Let $\mathbb {T}^{*}$ be an $\aleph _0$ -saturated elementary extension of $\mathbb {T}$ . Enlarging the valuation ring $\mathcal O_{\mathbb {T}^{*}}$ of $\mathbb {T}^{*}$ to $\dot {\mathcal O}_{\mathbb {T}^{*}} = \{ f \in \mathbb {T}^{*} : |f| \leqslant \exp ^n (x)\ \text {for some}\ n \geqslant 1 \}$ , where $\exp ^n$ denotes the nth iterate of the exponential function, yields a pre-H-field $(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ that is $\operatorname {d}$ -henselian, real closed, and has exponential integration. The saturation ensures that $\dot {\mathcal O}_{\mathbb {T}^{*}}$ is a proper subring of K, i.e., K contains a transexponential element. Moreover, the residue field of $(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ is elementarily equivalent to $\mathbb {T}$ as an ordered valued differential field. To explain this, we first review the theory of $\mathbb {T}$ .

In the language the theory of $\mathbb {T}$ is model complete and axiomatized by the theory $T^{\operatorname {nl}}_{\operatorname {small}}$ of newtonian, $\unicode{x3c9} $ -free, Liouville closed H-fields with small derivation. An asymptotic field K is differential-valued ( $\operatorname {d}$ -valued for short) if ; $\operatorname {d}$ -valued fields are pre- $\operatorname {d}$ -valued and an H-fields is a ${d}$ -valued pre-H-fields. An H-field K is Liouville closed if it is real closed, has exponential integration, and has integration in the sense that is surjective. For more on H-fields and related notions, see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Chapter 10]. The property of $\unicode{x3c9} $ -freeness is crucial to studying $\mathbb {T}$ but incidental here, so we refer the reader to [Reference Aschenbrenner, van den Dries and van der Hoeven5, Section 11.7]. Likewise, we do not define newtonianity, a technical cousin of $\operatorname {d}$ -henselianity, and instead refer the reader to [Reference Aschenbrenner, van den Dries and van der Hoeven5, Chapter 14].

To describe $(\mathbb {T}^{*}, \dot {\mathcal O}_{\mathbb {T}^{*}})$ and its residue field, it is convenient to work with value groups instead of valuation rings via the notions of coarsening and specialization. Let K be pre-H-field with small derivation and let $\Delta $ be a nontrivial proper convex subgroup of $\Gamma $ such that $\psi (\Delta ^{\neq }) \subseteq \Delta $ and $\psi (\Gamma \setminus \Delta ) \subseteq \Gamma \setminus \Delta $ . The coarsening of K by $\Delta $ is the ordered differential field K with the valuation

$$ \begin{align*} v_\Delta \colon K^{\times} &\to \Gamma/\Delta\\ a &\mapsto va + \Delta, \end{align*} $$

denoted by $K_\Delta $ . Its valuation ring is

with maximal ideal

Then $K_\Delta $ is asymptotic with gap $0$ by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 9.2.26 and Lemma 9.2.24], so it is pre- $\operatorname {d}$ -valued. Moreover, $K_\Delta $ is a pre-H-field: $\dot {\mathcal O}$ remains convex since and $K_{\Delta }$ still satisfies (PH3) since $\dot {\mathcal O}\supseteq \mathcal O$ .

Setting for $a \in \dot {\mathcal O}$ , we equip the ordered differential residue field of $K_\Delta $ with the valuation

$$ \begin{align*} v \colon \dot{K}^{\times} &\to \Delta\\ \dot{a} &\mapsto va, \end{align*} $$

making $\dot K$ an ordered valued differential field with small derivation called the specialization of K to $\Delta $ . Its valuation ring is with maximal ideal . Clearly, $\mathcal O_{\dot K}$ is convex. The map $\mathcal O \to \mathcal O_{\dot K}$ given by $a \mapsto \dot {a}$ induces an isomorphism $\operatorname {\mathrm {res}}(K) \cong \operatorname {\mathrm {res}}(\dot {K})$ of ordered differential fields. By [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.1.8], $\dot K$ is pre- $\operatorname {d}$ -valued and if K is $\operatorname {d}$ -valued, then so is $\dot {K}$ with $C_{\dot K}=C$ , where we identify C with a subfield of $C_{\dot K}$ via $a \mapsto \dot a$ . To see that $\dot K$ is a pre-H-field, it remains to check (PH3), so let $a \in \dot {\mathcal O}$ with $\dot a> \mathcal O_{\dot {K}}$ . Then $va \in \Delta ^{<}$ and so $v(a')=va+\psi (va) \in \Delta $ . Also, $a'>0$ , so , and thus $\dot {a}'>0$ . If K is an H-field, then so is $\dot {K}$ , since H-fields are exactly the $\operatorname {d}$ -valued pre-H-fields.

Suppose now that $K=\mathbb {T}^{*}$ and set , a nontrivial proper convex subgroup of $\Gamma $ with $\psi (\Delta ^{\neq }) \subseteq \Delta $ and $\psi (\Gamma \setminus \Delta ) \subseteq \Gamma \setminus \Delta $ , as in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Example 10.1.7]. It is an exercise to check that $\dot {\mathcal O}=\dot {\mathcal O}_{\mathbb {T}^{*}}$ as defined above. To summarize the preceding discussion, $K_\Delta $ is a pre-H-field with gap $0$ and $\dot {K}$ is an H-field with $C_{\dot {K}}=C$ , an elementary extension of $\mathbb {R}$ . As the differential field structure is unchanged, K is still real closed and has exponential integration. And since K is newtonian, $K_\Delta $ is $\operatorname {d}$ -henselian by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 14.1.2]; this also uses that is the unique element of $\Gamma ^{\neq }$ with $\psi (1)=1$ (see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.2.15]), so automatically $1 \in \Delta $ .

We now proceed to show that $\dot K \models T^{\operatorname {nl}}_{\operatorname {small}}$ . First, since K is $\unicode{x3c9} $ -free, so is $\dot K$ by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 11.7.24]. Since K is real closed, so is $\dot K$ , and since K has integration, so does $\dot K$ by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 9.4.13]. To see that $\dot K$ has exponential integration, let and take $y \in K^{\times }$ with $y^{\dagger }=a$ . We may have $vy=0 \in \Delta $ . If $vy\neq 0$ , from $\psi (vy)=v(y^{\dagger })=va \in \Delta $ , we get $vy \in \Delta $ . In either case, $\dot {y} \in \dot {K}^{\times }$ and $\dot {y}^{\dagger } = \dot a$ . Hence, $\dot K$ is Liouville closed.

It remains to see that $\dot K$ is newtonian. To explain this, for a valued differential field F and $a \in F^{\times }$ , we let $F^{a}$ denote the valued field F equipped with the derivation ; if F is newtonian, then so is $F^{a}$ . Let be such that $\psi (\delta ) \geqslant v\phi \in \Delta $ for some $\delta \in \Delta ^{\neq }$ . Then by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 14.1.4], it suffices to show that $(\dot {K}^{\dot \phi }, v_{\dot \phi }^{\flat })$ , the coarsening of $\dot {K}^{\dot \phi }$ by , is $\operatorname {d}$ -henselian. First, note that $\Delta _{\dot \phi }^{\flat } \neq \emptyset $ , since $\psi(\Delta^{\neq})$ has no maximum, and that , since for any $\gamma \in \Gamma _\phi ^{\flat }$ we have $\psi (\gamma ) \in \Delta $ and thus $\gamma \in \Delta $ . It follows that $(\dot {K}^{\dot \phi }, v_{\dot \phi }^{\flat })$ is isomorphic to the specialization of $(K^{\phi }, v_\phi ^{\flat })$ to $\Delta /\Gamma _\phi ^{\flat }$ , where $(K^{\phi }, v_\phi ^{\flat })$ denotes the coarsening of $K^{\phi }$ by $\Gamma _\phi ^{\flat }$ . Since K is newtonian, so is $K^{\phi }$ . Hence, $(K^{\phi }, v_\phi ^{\flat })$ is $\operatorname {d}$ -henselian by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 14.1.2] again; in applying this lemma, note that the new $1_{\phi }>0$ is the unique element of $\Gamma ^{\neq }$ with $\psi (1_{\phi })-v\phi = 1_{\phi }$ , so again $1_{\phi } \in \Gamma _\phi ^{\flat }$ . Thus the specialization of $(K^{\phi }, v_\phi ^{\flat })$ to $\Delta /\Gamma _\phi ^{\flat }$ is also $\operatorname {d}$ -henselian by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 7.1.6].

Putting this together, we have established:

Proposition 8.1. Let be an $\aleph _0$ -saturated elementary extension of $\mathbb {T}$ and set Then the coarsening $K_{\Delta }$ of K by $\Delta $ (equivalently, K equipped with the valuation ring $\dot {\mathcal O} = \{ f \in \mathbb {T}^{*} : |f| \leqslant \exp ^n (x)\ \text {for some}\ n \geqslant 1 \}$ ) is a pre-H-field that is $\operatorname {d}$ -henselian, real closed, and has exponential integration. The residue field $\dot {K}$ of $K_{\Delta }$ is elementarily equivalent to $\mathbb {T}$ as an ordered valued differential field.

In summary, we have thus decomposed $\mathbb {T}^{*}$ into a transexponential pre-H-field with gap $0$ , $K_{\Delta }$ , and an exponentially bounded model of $T^{\operatorname {nl}}_{\operatorname {small}}$ , $\dot {K}$ . Combining this with the results of the previous section and facts about $T^{\operatorname {nl}}_{\operatorname {small}}$ from [Reference Aschenbrenner, van den Dries and van der Hoeven5], we obtain:

Corollary 8.2. Let and $T_{\operatorname {\mathrm {res}}} = T^{\operatorname {nl}}_{\operatorname {small}}$ . Then $(K_\Delta , \dot {K}; \pi _2) \models T^{\operatorname {dhl}}$ , where $\pi _2$ is the binary version of the residue map $K \to \dot {K}$ , and $\dot {K}$ is purely stably embedded in $(K_\Delta , \dot {K}; \pi _2)$ in the sense of Corollary 7.6. Also, $T^{\operatorname {dhl}}$ :

  1. (i) is complete;

  2. (ii) is model complete, and moreover is the model companion of the theory T from Corollary 7.7 with $T^{\prime }_{\operatorname {\mathrm {res}}}$ being the theory of H-fields with small derivation;

  3. (iii) has NIP;

  4. (iv) has quantifier elimination if $\mathcal L_{\operatorname {\mathrm {res}}}$ and $T^{\operatorname {nl}}_{\operatorname {small}}$ are expanded by a function symbol for multiplicative inversion and two unary predicates and as in [Reference Aschenbrenner, van den Dries and van der Hoeven5, Chapter 16].

Proof. Note that $T_{\operatorname {\mathrm {res}}}$ is required to expand the theory of linearly surjective, expint-closed ordered differential fields, and indeed newtonian fields are linearly surjective by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 14.2.2]. Then the completeness of $T^{\operatorname {dhl}}$ follows from the completeness of $T^{\operatorname {nl}}_{\operatorname {small}}$ [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 16.6.3] by Corollary 7.3. Likewise, the model completeness and model companion statements follow from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Corollary 16.2.5] and Corollary 7.4, NIP follows from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Proposition 16.6.6] and Proposition 7.8, and quantifier elimination from [Reference Aschenbrenner, van den Dries and van der Hoeven5, Theorem 16.0.1] and Theorem 7.5.

8.2 Exponential integration is necessary

Unlike the AKE theorems for monotone fields, Corollary 7.3 includes the assumption of closure under exponential integration. Corollary 8.6 shows that the theorem fails when this assumption is dropped by exhibiting two $\operatorname {d}$ -henselian, real closed pre-H-fields that are not elementarily equivalent but have isomorphic ordered differential residue fields. More precisely, one of these pre-H-fields has exponential integration but the other does not, and, indeed, the issue is pervasive in the sense that it works for any ordered differential residue field that is linearly surjective and expint-closed. Additionally, the asymptotic couples of these pre-H-fields are elementarily equivalent, which shows that some assumption on logarithmic derivatives is necessary not just in Corollary 7.3, but even in any possible three-sorted extension of it.

This lemma is a variant of [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.4.6]. We now set for $A \subseteq K$ , and make use of the background material on asymptotic couples from Section 4.1.

Lemma 8.3. Suppose that K is pre- $\operatorname {d}$ -valued of H-type with gap $0$ . Let $s \in K$ such that $vs \in \Gamma ^< \setminus \Psi $ . Take a transcendental over K with $a^{\dagger } = s$ . Then the differential field can be equipped with a unique valuation making $K_a$ a pre- $\operatorname {d}$ -valued extension of K of H-type such that $a \succ 1$ . With this valuation, $K_a$ has gap $0$ and satisfies:

  1. (i) $\operatorname {\mathrm {res}}(K_a)=\operatorname {\mathrm {res}}(K)$ ;

  2. (ii) for every $f \in K_a^{\succ 1}$ , either $f^{\dagger } \sim g^{\dagger }$ for some $g \in K^{\succ 1}$ or $f^{\dagger } \sim es$ for some $e \in \mathbb {Q}^{>}$ ;

  3. (iii) the asymptotic couple $(\Gamma _a, \psi _a)$ of $K_a$ satisfies $\Gamma _a = \Gamma \oplus \mathbb {Q}\alpha $ , with ${\alpha = v(a)<0}$ and $[\alpha ] \notin [\Gamma ]$ , and $\psi _a(\gamma + e\alpha ) = \min \{\psi (\gamma ), v(es)\}$ for all $\gamma \in \Gamma $ and $e \in \mathbb {Q}$ ;

  4. (iv) if K is a pre-H-field with $s>0$ , then $K_a$ can be equipped with a unique ordering making $K_a$ an ordered field extension of K such that the valuation ring $\mathcal O_a$ of $K_a$ is convex; with this ordering, $K_a$ is a pre-H-field.

Proof. We first explain how we adjoin a new element $\alpha $ to the asymptotic couple $(\Gamma , \psi )$ , which later will satisfy $\alpha =v(a)$ . Let $D_1=\{\gamma \in \Gamma ^< : \psi (\gamma )<vs\}$ and $D_2=\{\gamma \in \Gamma ^{\leqslant } : \psi (\gamma )>vs\}$ , and take $\alpha $ such that $D_1<\alpha <D_2$ and $[D_1]>[\alpha ]>[D_2]$ . Recall that since $(\Gamma , \psi )$ has gap $0$ , for any $\gamma \in \Gamma ^<$ , we have $\gamma <\psi (\gamma )<0$ . In particular, $vs \in D_1$ and $[\alpha ]>[vs]$ . Set . Let $\gamma \in \Gamma $ and $e \in \mathbb {Q}$ , and define $\gamma + e\alpha>0$ if either $[\gamma ]>[e\alpha ]$ and $\gamma>0$ or $[e\alpha ]>[\gamma ]$ and $e<0$ ; if $e \neq 0$ , then $\gamma + e\alpha>0$ if and only if either $[\gamma ]\in [D_1]$ and $\gamma>0$ or $[\gamma ]\in [D_2]$ and $e<0$ . This makes $\Gamma _a$ an ordered abelian group extending $\Gamma $ such that $\alpha <0$ . Now we extend $v \colon K^{\times } \to \Gamma $ to the unique valuation $v \colon K_a^{\times } \to \Gamma _a$ satisfying $v(a)=\alpha $ . More explicitly, every element of $K_a^{\times }$ is of the form $p/q$ for some $p=\sum _{e \in \mathbb {Q}} p_e a^e$ and $q=\sum _{e \in \mathbb {Q}} q_e a^e$ , where all $p_e, q_e \in K$ with only finitely many of them nonzero and at least one $p_e$ and one $q_e$ are nonzero. For such p and q, we set

and it is routine to check that this map is a valuation. If w is any valuation on $K_a$ making it a pre- $\operatorname {d}$ -valued extension of K of H-type such that $wa<0$ , then $D_1<wa<D_2$ and $[D_1]>[wa]>[D_2]$ , and it follows that $w = v$ after identifying their value groups via $wa \mapsto \alpha $ .

For every $f \in K_a^{\times }$ , there are $g \in K^{\times }$ and a unique $e \in \mathbb {Q}$ such that $f \sim ga^e$ . Item (i) follows. Now we use [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.1.19] with $T = K^{\times }a^{\mathbb {Q}}$ and $L = K_a$ to show that $K_a$ is pre- $\operatorname {d}$ -valued. Let $t=ga^e \in T$ , where $g \in K^{\times }$ and $e \in \mathbb {Q}$ , and suppose that $t\prec 1$ . It suffices to show that $t^{\dagger } \succ 1$ and $t' \prec 1$ . We may assume that $e \neq 0$ . Then $v(g^{\dagger }) \neq vs=v(es)$ , so $v(t^{\dagger })=v(g^{\dagger }+es) = \min \{v(g^{\dagger }), vs\}<0$ . If $[vg]<[\alpha ]$ , then $e<0$ and $v(g^{\dagger })>vs$ , and so

$$\begin{align*}v(t')\ =\ v(t^{\dagger})+vt\ =\ vs+e\alpha+vg\>\ 0. \end{align*}$$

If $[vg]>[\alpha ]$ , then similarly $v(t') = vg'+e\alpha> 0$ . Hence $K_a$ is pre- $\operatorname {d}$ -valued with gap $0$ . To prove (ii), let $f \in K_a^{\succ 1}$ and take $g \in K^{\times }$ and $e \in \mathbb {Q}$ with $f \sim ga^e$ . Then $f^{\dagger } \sim g^{\dagger } + es$ , so likewise either $f^{\dagger } \sim g^{\dagger }$ and $vg<0$ or $f^{\dagger } \sim es$ and $e>0$ . This also finishes the proof of (iii).

Now we show that $(\Gamma _a, \psi _a)$ is of H-type. Let $\gamma _1, \gamma _2 \in \Gamma $ and $e_1, e_2 \in \mathbb {Q}$ and suppose that $\gamma _1+e_1\alpha <\gamma _2+e_2\alpha <0$ ; we need to show that $\psi (\gamma _1+e_1\alpha ) \leqslant \psi (\gamma _2+e_2\alpha )$ . If $[\gamma _i]>[e_i\alpha ]$ for $i=1,2$ , then $[\gamma _1] \geqslant [\gamma _2]$ and $\psi (\gamma _1+e_1\alpha )=\psi (\gamma _1) \leqslant \psi (\gamma _2)=\psi (\gamma _2+e_2\alpha )$ . If $[\gamma _i]<[e_i\alpha ]$ for $i=1,2$ , then $\psi (\gamma _1+e_1\alpha )=vs=\psi (\gamma _2+e_2\alpha )$ . If $[\gamma _1]>[e_1\alpha ]$ and $[\gamma _2]<[e_2\alpha ]$ , then $\psi (\gamma _1+e_1\alpha )=\psi (\gamma _1)<vs=\psi (\gamma _2+e_2\alpha )$ . Finally, if $[\gamma _1]<[e_1\alpha ]$ and $[\gamma _2]>[e_2\alpha ]$ , then $[e_1\alpha ]>[\gamma _2]$ and $\psi (\gamma _1+e_1\alpha )=vs<\psi (\gamma _2)=\psi (\gamma _2+e_2\alpha )$ (in this case, $e_2=0$ ).

To show (iv), suppose that K is a pre-H-field and $s>0$ . Let $f \in K_a^{\times }$ , and take $g \in K^{\times }$ and $e \in \mathbb {Q}$ with $f \sim ga^e$ . For $K_a$ to be an ordered field extension of K, we must have $a^e>0$ for all $e \in \mathbb {Q}$ , and hence there is at most one way to define the ordering: $f>0$ if and only if $g>0$ . First, this is independent of the choice of g: if $f \sim g_1a^{e}$ with $g_1 \in K^{\times }$ , then $g_1 \sim g$ , so $g_1>0$ if and only if $g>0$ . To see that this makes $K_a$ an ordered field, let $f_1, f_2 \in K_a^{\times }$ with $f_1, f_2>0$ . For $i=1,2$ , take $g_i \in K^{\times }$ and $e_i \in \mathbb {Q}$ such that $f_i \sim g_ia^{e_i}$ , so $g_i>0$ . Then $f_1f_2 \sim g_1g_2a^{e_1+e_2}>0$ . Similarly, $f^2>0$ for all $f \in K_a^{\times }$ . If $f_1\prec f_2$ , then $f_1+f_2 \sim f_2> 0$ ; the case $f_1 \succ f_2$ is symmetric. Suppose that $f_1 \asymp f_2$ , so $e_1=e_2$ and $g_1 \asymp g_2$ . Then since $g_1, g_2>0$ , we have $g_1+g_2 \asymp g_1$ and thus $f_1+f_2 \sim (g_1+g_2)a^{e_1}>0$ .

To see that $\mathcal O_a$ is convex with respect to this ordering, we show that its maximal ideal satisfies : If , then $1 \pm f \sim 1a^0>0$ . To show that $K_a$ is a pre-H-field, it remains to check that $f^{\dagger }>0$ whenever $f \in K_a$ satisfies $f>\mathcal O_a$ , which follows from the proof of (ii) and the fact that K is a pre-H-field.

Proposition 8.4. Let $\boldsymbol k$ be an ordered differential field that is linearly surjective and real closed. Then there is a $\operatorname {d}$ -henselian, real closed pre-H-field K such that $\Psi = \Gamma ^<$ and $\operatorname {\mathrm {res}}(K) \cong \boldsymbol k$ , but $(K^{\times })^{\dagger } \neq K$ .

Proof. We construct increasing sequences of pre-H-fields with gap $0$ $(K_n)$ and sets $(\mathfrak M_n)$ satisfying, for all n:

  1. (1) $K_n = \boldsymbol k(\mathfrak M_{n+1})$ ;

  2. (2) $\mathfrak M_{n+1}$ is a divisible monomial group for $K_n$ ;

  3. (3) $\mathfrak m>0$ for all $\mathfrak m \in \mathfrak M_n$ ;

  4. (4) $\mathfrak M_n^{\succ 1} \subseteq (\mathfrak M_{n+1}^{\succ 1})^{\dagger }$ ;

  5. (5) for every $f \in K_n^{\succ 1}$ , there are $\mathfrak m \in \mathfrak M_n^{\succ 1}$ and $e \in \mathbb {Q}^{>}$ such that $f^{\dagger } \sim e\mathfrak m$ .

In particular, each $K_n$ has ordered differential residue field naturally isomorphic to $\boldsymbol k$ .

Let $(b_i)_{i \in \mathbb {Q}}$ be algebraically independent over $\boldsymbol k$ and set . Let k range over and $\boldsymbol i = (i_1, \dots , i_k)$ and $\boldsymbol e = (e_1, \dots , e_k)$ over $\mathbb {Q}^{k}$ with $i_1<\dots <i_k$ . We set and define and . Set with $(\beta _i)_{i \in \mathbb {Q}}$ $\mathbb {Q}$ -linearly independent, and order $\Gamma _0$ lexicographically with each $\beta _i<0$ . That is, if $e_1 \neq 0$ , define $e_1\beta _{i_1} + \dots + e_k\beta _{i_k}>0$ if $e_1<0$ , so $[e_1\beta _{i_1} + \dots + e_k\beta _{i_k}] = [\beta _{i_1}]$ , and $[\beta _i]> [\beta _j]$ if and only if $i<j$ , for $i, j \in \mathbb {Q}$ .

We equip $K_0$ with the unique valuation $v \colon K_0^{\times } \to \Gamma _0$ satisfying $v(u)=0$ and $v(b_i)=\beta _i$ for all $u \in \boldsymbol k^{\times }$ and $i \in \mathbb {Q}$ . To do so more explicitly, note that every element $y \in K_0^{\times }$ has the form $y=(\sum _{\boldsymbol e} f_{\boldsymbol e}b_{\boldsymbol i}^{\boldsymbol e})/(\sum _{\boldsymbol e} g_{\boldsymbol e}b_{\boldsymbol i}^{\boldsymbol e})$ for some $\boldsymbol i$ and some $f_{\boldsymbol e}, g_{\boldsymbol e} \in \boldsymbol k$ such that only finitely many $f_{\boldsymbol e}$ and $g_{\boldsymbol e}$ are nonzero and at least one $f_{\boldsymbol e}$ and one $g_{\boldsymbol e}$ are nonzero. Define $v \colon K_0^{\times } \to \Gamma _0$ by , for $u \in \boldsymbol k^{\times }$ , and, for $y \in K_0^{\times }$ as above

It is routine to check that this map is a valuation. Hence $\mathfrak M_1$ is a divisible monomial group for $K_0$ . Moreover, for all $f \in K_0^{\times }$ with $f \not \asymp 1$ , there exist unique $u \in \boldsymbol k^{\times }$ , $\boldsymbol i$ , and $\boldsymbol e \in (\mathbb {Q}^{\times })^k$ such that $f \sim ub_{\boldsymbol i}^{\boldsymbol e}$ ; for all $f \in K_0^{\times }$ with $f \asymp 1$ , there exists a unique $u \in \boldsymbol k^{\times }$ such that $f \sim u$ .

We extend the derivation on $\boldsymbol k$ to $K_0$ so that $(b_i^e)^{\dagger } = eb_{i+1}$ for all $i, e \in \mathbb {Q}$ . It is easy to verify that for all , if $t \prec 1$ , then $t' \prec 1$ and $t^{\dagger } \succ 1$ . Then by [Reference Aschenbrenner, van den Dries and van der Hoeven5, Lemma 10.1.19] (with $K=\boldsymbol k$ and $L=K_0$ ), $K_0$ is pre- $\operatorname {d}$ -valued and has gap $0$ . Also, for every $f \in K_0^{\succ 1}$ , there exist $i, e \in \mathbb {Q}$ with $e> 0$ and $f^{\dagger } \sim eb_i$ . Hence the asymptotic couple of $K_0$ is $(\Gamma _0, \psi _0)$ , where $\psi _0$ is defined by $\psi _0(e_1\beta _{i_1} + \dots + e_k\beta _{i_k}) = \beta _{i_1+1}$ if $e_1 \neq 0$ , and it satisfies .

To extend the ordering on $\boldsymbol k$ to $K_0$ , let $f \in K_0^{\times }$ and take $u \in \boldsymbol k^{\times }$ and $\mathfrak m \in \mathfrak M_1$ with $f \sim u\mathfrak m$ . Then define $f>0$ if $u>0$ . The proof that this makes $K_0$ a pre-H-field is similar to the corresponding part of the proof of Lemma 8.3.

To construct $K_1$ and $\mathfrak M_2$ , enumerate $(\mathfrak M_1 \setminus \mathfrak M_0)^{\succ 1}$ by $(\mathfrak m_n)$ and set and . By (5), $v\mathfrak m_n \in \Gamma _{0}^< \setminus \Psi _{0}$ for all n. Hence by repeated applications of Lemma 8.3, we construct increasing sequences of pre-H-fields with gap $0$ $(L_{n})$ , sets $(\mathfrak N_{n})$ , and elements $(\mathfrak n_n)$ such that, for all n:

  1. (a) $L_{n} = \boldsymbol k(\mathfrak N_{n})$ ;

  2. (b) $\mathfrak N_{n}$ is a divisible monomial group for $L_{n}$ ;

  3. (c) $\mathfrak N_{n+1} = \mathfrak N_{n}\mathfrak n_n^{\mathbb {Q}}$ ;

  4. (d) $\mathfrak n_n^{\dagger } = \mathfrak m_n$ , $\mathfrak n_n \succ 1$ , and $\mathfrak n_n^e>0$ for all $e \in \mathbb {Q}$ ;

  5. (e) for every $f \in L_{n+1}^{\succ 1}$ , either $f^{\dagger } \sim g^{\dagger }$ for some $g \in L_{n}^{\succ 1}$ or $f^{\dagger } \sim e\mathfrak m_n$ for some $e \in \mathbb {Q}^{>}$ ;

  6. (f) $v\mathfrak m_m \notin \Psi _{L_{n}}$ for all $m \geqslant n$ .

Now we set and . Iterating this procedure yields the desired sequences $(K_n)$ and $(\mathfrak M_n)$ . Set , where . Then K is a pre-H-field with gap $0$ such that:

  1. (i) K has ordered differential residue field (isomorphic to) $\boldsymbol k$ ;

  2. (ii) $\Gamma $ is nontrivial and divisible, and satisfies $\Gamma ^< = \Psi $ ;

  3. (iii) for every $f \in K^{\succ 1}$ , there are $\mathfrak m \in \mathfrak M^{\succ 1}$ and $e \in \mathbb {Q}^{>}$ such that $f^{\dagger } \sim e\mathfrak m$ .

We have not yet needed the assumptions on $\boldsymbol k$ . Using that $\boldsymbol k$ is linearly surjective, we pass to the $\operatorname {d}$ -henselization of K, an immediate pre-H-field extension of K, thereby arranging that K is $\operatorname {d}$ -henselian while preserving (i)–(iii). Then $K$ is real closed, since $\boldsymbol{k} $ is real closed, $\Gamma $ is divisible, and K is henselian. Finally, by (iii) we have $u\mathfrak m \notin (K^{\times })^{\dagger }$ for all $u \in \boldsymbol k \setminus \mathbb {Q}$ and $\mathfrak m \in \mathfrak M^{\succ 1}$ .

By taking a $\boldsymbol k$ as above that additionally has exponential integration, we obtain:

Corollary 8.5. There exists a $\operatorname {d}$ -henselian, real closed pre-H-field K such that $\Psi = \Gamma ^<$ and $(\operatorname {\mathrm {res}}(K)^{\times })^{\dagger } = \operatorname {\mathrm {res}}(K)$ , but $(K^{\times })^{\dagger } \neq K$ .

Corollary 8.6. There exist $\operatorname {d}$ -henselian, real closed pre-H-fields $K_1$ and $K_2$ such that $\operatorname {\mathrm {res}}(K_1) \cong \operatorname {\mathrm {res}}(K_2)$ (as ordered differential fields) and $(\Gamma _1, \psi _1) \equiv (\Gamma _2, \psi _2)$ , but $(K_1^{\times })^{\dagger } \neq K_1$ and $(K_2^{\times })^{\dagger } = K_2$ ; in particular, $K_1 \not \equiv K_2$ .

Proof. Let $\boldsymbol k$ be an ordered differential field that is linearly surjective and expint-closed. Then Proposition 8.4 yields a $\operatorname {d}$ -henselian, real closed pre-H-field $K_1$ with ordered differential residue field isomorphic to $\boldsymbol k$ , nontrivial divisible asymptotic couple $(\Gamma _1, \psi _1)$ satisfying $\Gamma _1^< = \Psi _1$ , and $(K_1^{\times })^{\dagger } \neq K_1$ . Now let $K_2$ be the $\operatorname {d}$ -Hensel–Liouville closure of $K_1$ by Theorem 6.16. Then $\operatorname {\mathrm {res}}(K_2)=\operatorname {\mathrm {res}}(K_1)$ and the nontrivial divisible asymptotic couple $(\Gamma _2, \psi _2)$ of $K_2$ satisfies $\Gamma _2^< = \Psi _2$ . Hence $(\Gamma _1, \psi _1) \equiv (\Gamma _2, \psi _2)$ by Corollary 4.9.

Acknowledgments

Thanks are due to Lou van den Dries for helpful discussions and for comments on an earlier draft of this paper, to Anton Bernshteyn and Chris Miller for suggestions on an earlier draft of this paper, and to Matthias Aschenbrenner for comments related to Proposition 7.8. I am also grateful to the anonymous referee for carefully reading the paper and suggesting many improvements. Some of this research appeared in an earlier form in the author’s PhD thesis [Reference Pynn-Coates17].

Funding

This research was funded in whole or in part by the Austrian Science Fund (FWF) 10.55776/ESP450. For open access purposes, the author has applied a CC BY public copyright licence to any author accepted manuscript version arising from this submission. This material is based upon work supported by the National Science Foundation under Grant No. DMS-2154086. Presentation of some of this research at DART X was supported by NSF Grant No. DMS-1952694.

Footnotes

1 Instead, $\mathbb {T}$ satisfies a strong eventual form of $\operatorname {d}$ -henselianity called newtonianity; see [Reference Aschenbrenner, van den Dries and van der Hoeven5, Chapter 14].

2 The definition is similar to that of expint-extensions in Section 6.2 except that we also allow ${t_i' \in E(t_1, \dots , t_{i-1})}$ .

References

REFERENCES

Aschenbrenner, M., Some remarks about asymptotic couples , Valuation Theory and its Applications (Saskatoon, SK, 1999) (F.-V. Kuhlmann, S. Kuhlmann, and M. Marshall, editors), vol. 2, Fields Institute Communications, 33, American Mathematical Society, Providence, 2003, pp. 718.Google Scholar
Aschenbrenner, M., Chernikov, A., Gehret, A., and Ziegler, M., Distality in valued fields and related structures . Transactions of the American Mathematical Society, vol. 375 (2022), no. 7, pp. 46414710.CrossRefGoogle Scholar
Aschenbrenner, M. and van den Dries, L., Closed asymptotic couples . Journal of Algebra, vol. 225 (2000), no. 1, pp. 309358. https://doi.org/10.1006/jabr.1999.8128.CrossRefGoogle Scholar
Aschenbrenner, M. and van den Dries, L. $H$ -fields and their Liouville extensions . Mathematische Zeitschrift, vol. 242 (2002), no. 3, pp. 543588. https://doi.org/10.1007/s002090000358.CrossRefGoogle Scholar
Aschenbrenner, M., van den Dries, L., and van der Hoeven, J., Asymptotic Differential Algebra and Model Theory of Transseries, vol. 195, Annals of Mathematics Studies, Princeton University Press, Princeton, 2017, pp. xxi+849. https://doi.org/10.1515/9781400885411.Google Scholar
Aschenbrenner, M., van den Dries, L., and van der Hoeven, J., Revisiting closed asymptotic couples . Proceedings of the Edinburgh Mathematical Society (2), vol. 65 (2022), no. 2, pp. 530555. https://doi.org/10.1017/S0013091522000219.CrossRefGoogle Scholar
Ax, J. and Kochen, S., Diophantine problems over local fields: III: Decidable fields . Annals of Mathematics (2), vol. 83 (1966), pp. 437456. https://doi.org/10.2307/1970476.CrossRefGoogle Scholar
Bélair, L. and Bousquet, M., Types dans les corps valués . Comptes Rendus de l’Académie des Sciences, vol. 323 (1996), no. 8, pp. 841844.Google Scholar
Boshernitzan, M., Hardy fields and existence of transexponential functions . Aequationes Mathematicae, vol. 30 (1986), nos. 2–3, pp. 258280. https://doi.org/10.1007/BF02189932.CrossRefGoogle Scholar
Delon, F., Types sur $C((X))$ , Study Group on Stable Theories (Bruno Poizat) Second year: 1978/79 (French), Secrétariat Math., Paris, 1981. Exp. No. 5, 29.Google Scholar
Ershov, Y. L., On the elementary theory of maximal normed fields . Doklady Akademii Nauk SSSR, vol. 165 (1965), pp. 2123.Google Scholar
Hakobyan, T., An Ax-Kochen-Ershov theorem for monotone differential-Henselian fields . The Journal of Symbolic Logic, vol. 83 (2018), no. 2, pp. 804816. https://doi.org/10.1017/jsl.2017.64.CrossRefGoogle Scholar
Marker, D., Model Theory: An Introduction, vol. 217, Graduate Texts in Mathematics, Springer-Verlag, New York, 2002, pp. viii+342.Google Scholar
Pynn-Coates, N., Differential-henselianity and maximality of asymptotic valued differential fields . Pacific Journal of Mathematics, vol. 308 (2020), no. 1, pp. 161205. https://doi.org/10.2140/pjm.2020.308.CrossRefGoogle Scholar
Pynn-Coates, N., Tame pairs of transseries fields, Preprint. 2024. https://arxiv.org/abs/2408.07033v1.Google Scholar
Pynn-Coates, N., “Dimension and topology in transserial tame pairs”, 2025, preprint at url: https://arxiv.org/abs/2508.16415.Google Scholar
Pynn-Coates, N., “On asymptotic valued differential fields with small derivation”, PhD thesis, University of Illinois at Urbana-Champaign, 2020.Google Scholar
Rosenlicht, M., On the value group of a differential valuation . American Journal of Mathematics, vol. 101 (1979), no. 1, pp. 258266. https://doi.org/10.2307/2373949.CrossRefGoogle Scholar
Rosenlicht, M., Differential valuations . Pacific Journal of Mathematics, vol. 86 (1980), no. 1, pp. 301319. url: https://projecteuclid.org/euclid.pjm/1102780625.CrossRefGoogle Scholar
Rosenlicht, M., On the value group of a differential valuation. II . American Journal of Mathematics, vol. 103 (1981), no. 5, pp. 977996. https://doi.org/10.2307/2374255.CrossRefGoogle Scholar
Scanlon, T., A model complete theory of valued $D$ -fields . Journal of Symbolic Logic, vol. 65 (2000), no. 4, pp. 17581784. https://doi.org/10.2307/2695074.CrossRefGoogle Scholar
Singer, M. F., The model theory of ordered differential fields . Journal of Symbolic Logic, vol. 43 (1978), no. 1, pp. 8291. https://doi.org/10.2307/2271951.CrossRefGoogle Scholar
van den Dries, L., Macintyre, A., and Marker, D., Logarithmic-exponential power series . Journal of the London Mathematical Society (2), vol. 56 (1997), no. 3, pp. 417434. https://doi.org/10.1112/S0024610797005437.CrossRefGoogle Scholar