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

When is the étale open topology a field topology?

Philip Dittmann Department of Mathematics, University of Manchester, Manchester M13 9PL, United Kingdom philip.dittmann@manchester.ac.uk Erik Walsberg University of California, Irvine ewalsber@uci.edu math.uci.edu/ ewalsber/  and  Jinhe Ye Institut de Mathématiques de Jussieu - Paris Rive Gauche
Current address:
Mathematical Institute, University of Oxford
jinhe.ye@imj-prg.fr, jinhe.ye@maths.ox.ac.uk sites.google.com/view/vincentye
Abstract.

We investigate the following question: Given a field KK, when is the étale open topology K\mathscr{E}_{K} induced by a field topology? On the positive side, when KK is the fraction field of a local domain RKR\neq K, using a weak form of resolution of singularities due to Gabber, we show that K\mathscr{E}_{K} agrees with the RR-adic topology when RR is quasi-excellent and henselian. Various pathologies appear when dropping the quasi-excellence assumption. For locally bounded field topologies, we introduce the notion of generalized t-henselianity (gt-henselianity) following Prestel and Ziegler. We establish the following: For a locally bounded field topology τ\uptau, the étale open topology is induced by τ\uptau if and only if τ\uptau is gt-henselian and some non-empty étale image is τ\uptau-bounded open. On the negative side, we obtain that for a pseudo-algebraically closed field KK, K\mathscr{E}_{K} is never induced by a field topology.

JY was partially supported by GeoMod AAPG2019 (ANR-DFG), Geometric and Combinatorial Configurations in Model Theory. The writing up of this material profited from discussions while PD and JY participated in the programme “Definability, Decidability and Computability in Number Theory, Part 2” hosted by the Mathematical Sciences Research Institute in Berkeley, California, and supported by the US National Science Foundation under Grant No. DMS-1928930. We would like to thank Will Johnson for spotting an error in the proof of Proposition 8.6 in a previous version.

1. Introduction

We continue the study of the étale open topology, initiated in [JTWY22] and continued in [WY23] and [JWY21]. Recall that the étale topology for a field KK, also called K\mathscr{E}_{K}, is given by a topology on the set of rational points V(K)V(K) for every KK-variety VV (a system of topologies in the terminology of [JTWY22]); concretely, the K\mathscr{E}_{K}-topology on V(K)V(K) is defined to have as a basis the collection of sets f(W(K))f(W(K)), where WW is another KK-variety and f:WVf\colon W\to V is an étale morphism.

The étale open topology is only interesting in the case of fields which are large in the sense of Pop (see [Pop14]) but not separably closed, since otherwise it degenerates to the discrete topology or the Zariski topology, respectively. Under this restriction, however, the abstract definition coincides with familiar topologies in many cases: Notably, over the fields ,,p\mathbb{C},\mathbb{R},\mathbb{Q}_{p} we recover on each variety the Zariski topology, resp. real topology, resp. pp-adic topology. In particular, for \mathbb{R} and p\mathbb{Q}_{p}, the étale open topology on every variety is induced by a Hausdorff non-discrete field topology on the ground field.

To generalize the phenomenon on \mathbb{R} or p\mathbb{Q}_{p}, consider a local domain RKR\subsetneq K with fraction field KK, and recall that the RR-adic topology on KK is the field topology with basis {aR+b:aK×,bK}\{aR+b\colon a\in K^{\times},b\in K\}. Like any other field topology, this induces a topology on V(K)V(K) for any KK-variety VV, which we also call the RR-adic topology. If RR is a (non-trivial) valuation ring, then the RR-adic topology is the usual valuation topology.

We now have the following facts relating K\mathscr{E}_{K} and RR-adic topologies.

Fact 1.1.

Let RKR\subsetneq K be a local domain with fraction field KK.

  1. (1)

    [JWY21, Theorem 1.2] If RR is henselian111We recall the definition of a henselian local ring below. For a valuation ring, this agrees with the usual notion of henselianity. then the RR-adic topology refines K\mathscr{E}_{K}.

  2. (2)

    [JTWY22, Theorem 6.15] If RR is a valuation ring and the Henselization of KK with respect to the corresponding valuation is not separably closed, then K\mathscr{E}_{K} refines the RR-adic topology.

  3. (3)

    [JWY21, Theorem 1.2] If RR is a regular (in the sense of commutative algebra) then K\mathscr{E}_{K} refines the RR-adic topology.

Here by the RR-adic topology refining K\mathscr{E}_{K} or vice versa we mean that the corresponding topologies on V(K)V(K) refine each other for every variety V/KV/K. (Note, however, that this is equivalent to merely saying that the same holds only on Kn=𝔸n(K)K^{n}=\mathbb{A}^{n}(K) for every nn, see Fact 2.3 below.) The present paper is motivated by the following natural questions:

Question 1.2.
  1. (1)

    When is the K\mathscr{E}_{K}-topology induced by a field topology?

  2. (2)

    When does the K\mathscr{E}_{K}-topology agree with the RR-adic topology for a local domain RKR\subsetneq K with fraction field KK?

We prove that the K\mathscr{E}_{K} is not induced by a field topology when KK is a pseudo-algebraically closed (PAC) field (Proposition 7.1 below), answering a question posed in [JTWY22, Section 8]. Since “most” algebraic extensions of \mathbb{Q} in a suitable sense are PAC, see [DF21, Proposition 1], this shows that the “generic” answer to Question 1.2(1) is negative.

In the other direction, we extend Fact 1.1 to quasi-excellent local domains, a wide class of non-pathological Noetherian domains:

Theorem 1.3 (Theorem 4.5).

If RKR\subsetneq K is a quasi-excellent local domain with fraction field KK, then the K\mathscr{E}_{K}-topology refines the RR-adic topology.

Together with Fact 1.1(1), we deduce:

Corollary 1.4.

If RKR\subsetneq K is quasi-excellent henselian local domain (e.g. RR a complete Noetherian local domain) with fraction field KK, then the RR-adic topology coincides with the K\mathscr{E}_{K}-topology.

In the case of a 11-dimensional Noetherian henselian local domain RR, we can even characterize precisely when the RR-adic topology coincides with the étale open topology on the fraction field, see Corollary 3.5.

In Sections 5 and 6 we give examples of pathologies that can arise when the quasi-excellence assumption is dropped, exhibiting at the same time interesting behaviour of the étale open topology under finite field extensions.

Finally, to study Question 1.2 in much greater generality, we borrow the model-theoretic tools of [PZ78]. This allows to obtain comprehensive answers at least up to replacing the field KK by a suitable elementary extension.

In this vein, it had previously been shown [JTWY22, Theorem B] that the K\mathscr{E}_{K}-topology for KK not separably closed is induced by a so-called V-topology on KK if and only if KK is a so-called t-henselian field, i.e. if and only if some elementary extension KKK^{\ast}\succ K carries a henselian valuation.

We study a notion of gt-henselian field topologies, a natural generalization of the notion of a t-henselian field topology from [PZ78]. In fact, this notion agrees with a different notion of henselianity for rings suggested (but hardly studied) in the literature, see Remark 8.2. When K\mathscr{E}_{K} is induced by a field topology, that topology must necessarily be gt-henselian (Lemma 8.14). We then obtain the following answer to Question 1.2 with the restriction to locally bounded field topologies:

Theorem 1.5 (Proposition 8.16).

Suppose that τ\uptau is a locally bounded field topology on KK. Then τ\uptau induces the K\mathscr{E}_{K}-topology if and only if τ\uptau is gt-henselian and some nonempty étale image in KK is τ\uptau-bounded.

Theorem 1.6.

The K\mathscr{E}_{K}-topology is induced by a locally bounded field topology if and only if there exists an elementary extension KKK^{\ast}\succ K and a henselian local domain RKR\subsetneq K^{\ast} with fraction field KK^{\ast} such that the RR-adic topology induces K\mathscr{E}_{K^{\ast}}.

It remains open whether the K\mathscr{E}_{K}-topology can ever be induced by a field topology which is not locally bounded.

2. Conventions and background

Throughout, KK is a field and Char(K)\operatorname{Char}(K) its characteristic.

2.1. Scheme theory

A KK-variety is a separated KK-scheme of finite type, not necessarily irreducible or reduced. (This is the convention of [Poo17, Definition 2.1.1].) Throughout 𝔸n\mathbb{A}^{n} is nn-dimensional affine space over KK, i.e. 𝔸n=SpecK[X1,,Xn]\mathbb{A}^{n}=\operatorname{Spec}K[X_{1},\ldots,X_{n}]. We let V(K)V(K) be the set of KK-points of a KK-variety VV. Given a scheme WW we let 𝒪W\mathcal{O}_{W} be the structure sheaf of WW, 𝒪W,p\mathcal{O}_{W,p} be the local ring of WW at pWp\in W, and let 𝒪p=𝒪W,p\mathcal{O}_{p}=\mathcal{O}_{W,p} when WW is clear.

2.2. The étale open topology

We gather basic facts on the K\mathscr{E}_{K}-topology from [JTWY22]. For a KK-variety WW, the étale open topology on the set of rational points W(K)W(K) is the topology with basis the collection of étale images UW(K)U\subseteq W(K), i.e. the sets U=f(V(K))U=f(V(K)) where f:VWf\colon V\to W is an étale morphism. We also write K\mathscr{E}_{K} for the étale open topology on W(K)W(K) for any WW, with WW generally clear from context.

Fact 2.1.

Suppose that VWV\to W is a morphism of KK-varieties. Then:

  1. (1)

    the induced map V(K)W(K)V(K)\to W(K) is K\mathscr{E}_{K}-continuous.

  2. (2)

    if VWV\to W is étale then the induced map V(K)W(K)V(K)\to W(K) is K\mathscr{E}_{K}-open.

  3. (3)

    the map KKK\to K, xαx+βx\mapsto\alpha x+\beta is an K\mathscr{E}_{K}-homeomorphism for any αK×,βK\alpha\in K^{\times},\beta\in K.

  4. (4)

    if nn is prime to Char(K)\operatorname{Char}(K) then {αn:αK×}\{\alpha^{n}:\alpha\in K^{\times}\} is an étale open subset of KK.

Proof.

(1), (2) is [JTWY22, Lemma 5.2, 5.3], respectively. (3) follows from (1). Let 𝔾m=SpecK[X,X1]\mathbb{G}_{m}=\operatorname{Spec}K[X,X^{-1}] be the scheme-theoretic multiplicative group over KK. Then (4) follows from (2) as the morphism 𝔾m𝔾m\mathbb{G}_{m}\to\mathbb{G}_{m}, XXnX\mapsto X^{n} is étale when nn is prime to Char(K)\operatorname{Char}(K). ∎

Suppose that LL is an extension of KK and VV is a KK-variety. We let VL=V×SpecKSpecLV_{L}=V\times_{\operatorname{Spec}K}\operatorname{Spec}L be the base change of VV. Recall that VL(L)V_{L}(L) is canonically identified with V(L)V(L), so we canonically equip V(L)V(L) with the L\mathscr{E}_{L}-topology. Fact 2.2 below is [JTWY22, Theorem 5.8].

Fact 2.2.

Suppose that LL is an algebraic extension of KK and VV is a KK-variety. Then the K\mathscr{E}_{K}-topology on V(K)V(K) refines the topology induced on V(K)V(K) by the L\mathscr{E}_{L}-topology on V(L)=VL(L)V(L)=V_{L}(L), i.e. if OV(L)O\subseteq V(L) is L\mathscr{E}_{L}-open then OV(K)O\cap V(K) is K\mathscr{E}_{K}-open.

2.3. Ring topologies and field topologies

Our general reference for ring topologies and field topologies is [PZ78], and we follow its conventions. In particular, ring topologies are always taken to be Hausdorff and not discrete.

We have the following basic fact about comparisons between the étale open topology and a field topology, proven in [JTWY22, Lemma 4.8, Lemma 4.2].

Fact 2.3.

Suppose that τ\uptau is a field topology on KK. If the τ\uptau-topology on each Kn=𝔸n(K)K^{n}=\mathbb{A}^{n}(K) refines the K\mathscr{E}_{K}-topology, then the τ\uptau-topology on V(K)V(K) refines the K\mathscr{E}_{K}-topology for any KK-variety VV. If the K\mathscr{E}_{K}-topology on KK refines τ\uptau, then the K\mathscr{E}_{K}-topology on V(K)V(K) refines the τ\uptau-topology for any KK-variety VV.

Let RR be a domain with fraction field K=Frac(R)K=\operatorname{Frac}(R), and assume RKR\neq K. The RR-adic topology on KK is the topology with basis {aR+b:aK×,bK}\{aR+b\colon a\in K^{\times},b\in K\}. This is a ring topology. (Compare [PZ78, Example 1.2], although the name RR-adic topology is not used there.) We are chiefly but not exclusively interested in the situation where RR is local.

We let J(R)J(R) be the Jacobson radical of RR. It is the intersection of all maximal ideals of RR, or equivalently J(R)={xR:1+xRR×}J(R)=\{x\in R\colon 1+xR\subseteq R^{\times}\}.

Fact 2.4.

The RR-adic topology on KK is a field topology if and only if J(R){0}J(R)\neq\{0\}.

Proof.

The right to left implication is [Joh20, Proposition 3.1]. We prove the left to right implication. Suppose that the RR-adic topology is a field topology. Then inversion gives a continuous map K×K×K^{\times}\to K^{\times}. Hence there is nonzero αR\alpha\in R such that (1+αR)1R(1+\alpha R)^{-1}\subseteq R. Thus (1+αR)R×(1+\alpha R)\subseteq R^{\times} and αJ(R)\alpha\in J(R). ∎

Fact 2.5 follows from Fact 2.3, 2.4 and the definitions. We leave the details to the reader.

Fact 2.5.

Suppose that RR has nonzero Jacobson radical (so the RR-adic topology is a field topology.) The following are equivalent:

  1. (1)

    The K\mathscr{E}_{K}-topology on V(K)V(K) refines the RR-adic topology for any KK-variety VV.

  2. (2)

    RR contains a nonempty K\mathscr{E}_{K}-open subset of KK.

Given a ring topology τ\uptau on KK, a set BKB\subseteq K is called bounded if for every neighbourhood UU of 0 there exists a neighbourhood VV of 0 such that VBUV\cdot B\subseteq U. The topology τ\uptau is locally bounded if there exists a bounded neighbourhood of 0.

Fact 2.6.

Let τ\uptau be a ring topology on KK and SS an open subring of KK. Then K=Frac(S)K=\operatorname{Frac}(S).

Proof.

Suppose that αK\alpha\in K and αFrac(S)\alpha\notin\operatorname{Frac}(S). Note that SS is a neighbourhood of zero. Then SαS={0}S\cap\alpha S=\{0\}, hence τ\uptau is discrete, contradiction. ∎

Fact 2.7.

Let τ\uptau be a ring topology on KK and SS a bounded open subring of KK. Then τ\uptau is the SS-adic topology.

Therefore the RR-adic topologies are exactly the ring topologies which admit bounded open subrings.

Proof.

Since {αS+β:αK×,βK}\{\alpha S+\beta:\alpha\in K^{\times},\beta\in K\} is a basis for the SS-adic topology and SS is open in τ\uptau, τ\uptau is finer than the SS-adic topology. By boundedness of SS and non-discreteness of the τ\uptau-topology, for every τ\uptau-open neighbourhood UU of 0 there exists an αK×\alpha\in K^{\times} with αSU\alpha S\subseteq U. This implies that the SS-adic topology refines τ\uptau. ∎

2.4. Commutative algebra

Let RR be local with maximal ideal 𝔪\mathfrak{m}. Then RR is henselian if for any fR[X]f\in R[X] and αR\alpha\in R with f(α)0f(α)(mod𝔪)f(\alpha)\equiv 0\not\equiv f^{\prime}(\alpha)\pmod{\mathfrak{m}} there is αR\alpha^{*}\in R such that f(α)=0f(\alpha^{*})=0 and αα(mod𝔪)\alpha^{*}\equiv\alpha\pmod{\mathfrak{m}}.

Fact 2.8.

The following are equivalent for a local domain RR with maximal ideal 𝔪\mathfrak{m}.

  1. (1)

    RR is henselian,

  2. (2)

    If a0,,an1𝔪a_{0},\ldots,a_{n-1}\in\mathfrak{m} then Xn+1Xn+an1Xn1++a1X+a0X^{n+1}-X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in 𝔪+1\mathfrak{m}+1.

  3. (3)

    If a0,,an1𝔪a_{0},\ldots,a_{n-1}\in\mathfrak{m} then Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in 𝔪1\mathfrak{m}-1.

Proof.

(1)\Leftrightarrow(2) is in [Gab92, Proposition 1]. (2)\Leftrightarrow(3) follows by considering the substitution Y=XY=-X. ∎

We gather some more intricate notions from commutative algebra, for use in Sections 3, 4 and 5. Let SS be a ring. We let dimS\dim S be the Krull dimension of SS. If SS is local then SS is regular if SS is Noetherian and the maximal ideal of SS is generated by dimS\dim S elements. This is a notion of non-singularity. A locally Noetherian scheme is defined to be regular if all its stalks are regular local rings, and a Noetherian ring RR is defined to be regular if SpecR\operatorname{Spec}R is regular, i.e. if all localizations of RR at prime ideals are regular local rings.

Fact 2.9.

Suppose that RR is a one-dimensional Noetherian domain, KK is the fraction field of RR, and SS is the integral closure of RR in KK. Then SS is a regular ring.

Proof.

By Krull-Akizuki [Sta20, Tag 00PG] SS is Noetherian, and by [Sta20, Tag 00OK] SS is one-dimensional. A one-dimensional Noetherian normal domain is a Dedekind domain, hence regular [Sta20, Tag 034X]. ∎

Let RR be a domain with fraction field KK and SS the integral closure of RR in KK. Then RR is normal if R=SR=S, RR is NN-1 if SS is a finite RR-module, and RR is Japanese (or NN-2) if the integral closure of RR in any finite field extension of KK is a finite RR-module. Non-Japanese Noetherian rings are viewed as pathologies.

We now discuss quasi-excellent rings, a class of Noetherian rings, and the related slightly more restrictive class of excellent rings. The definitions in full generality are somewhat technical, so we omit them. We direct the readers to [Sta20, Tag 07QT, 07GH, 07P7, 00NL] for the definitions and to [Rot97] for a friendlier introduction, as well as [ILO14, Exposé I] for a comprehensive overview. The class of excellent rings excludes certain pathologies that can arise for general Noetherian rings, but nevertheless includes virtually all “naturally occurring” Noetherian rings.

We give a definition of quasi-excellence for local rings. Suppose that LL is a field and RR is an LL-algebra. Then RR is geometrically regular if RLLalgR\otimes_{L}L^{\mathrm{alg}} is regular, where LalgL^{\mathrm{alg}} is an algebraic closure of LL. Regularity implies geometric regularity when LL is perfect. A morphism RSR\to S of Noetherian rings is regular if RSR\to S is flat and SRFrac(R/𝔭)S\otimes_{R}\operatorname{Frac}(R/\mathfrak{p}) is geometrically regular over Frac(R/𝔭)\operatorname{Frac}(R/\mathfrak{p}) for every prime ideal 𝔭\mathfrak{p} in RR. In scheme-theoretic language RSR\to S is regular if it is flat and every scheme-theoretic fiber of SpecSSpecR\operatorname{Spec}S\to\operatorname{Spec}R is geometrically regular.

Fact 2.10.

Let SS be a Noetherian local ring.

  1. (1)

    SS is quasi-excellent if and only if SS^S\to\widehat{S} is regular, where S^\widehat{S} is the completion.

  2. (2)

    If SS is either normal or henselian, then SS is quasi-excellent if and only if it is excellent.

See the discussion in [Mat80, Section 34] or [ILO14, Exposé I, Proposition 5.5.1 (ii)] for Fact 2.10(1) and [HRW04, Corollary 2.3] or [Sta20, Tag 0C2F] for Fact 2.10(2). We may take regularity of SS^S\to\widehat{S} to be the definition of quasi-excellence for Noetherian local rings. Note that complete local Noetherian rings are trivially excellent by this definition. We collect some general facts.

Fact 2.11.
  1. (1)

    The class of normal rings is closed under localizations.

  2. (2)

    The class of quasi-excellent rings is closed under finite extensions, localizations, and quotients.

  3. (3)

    Complete local rings are excellent.

  4. (4)

    Quasi-excellent rings are Japanese.

  5. (5)

    The class of henselian local rings is closed under quotients.

  6. (6)

    The Henselization of a quasi-excellent local ring is quasi-excellent.

  7. (7)

    If RR is NN-1 and Char(K)=0\operatorname{Char}(K)=0 then RR is Japanese.

Proof.

(1) is [Sta20, Tag 00GY] and (2) is [Sta20, Tag 07QU]. (3) follows from Fact 2.10. (4) is [Sta20, Tag 07QV]. (5) follows easily from the definitions. (6) is [Gro67, Corollaire 18.7.6]. (7) is [Sta20, Tag 032M]. ∎

Remark 2.12.

We now give some examples of excellent (in particular quasi-excellent) henselian local rings, most of which arise as local rings in various kinds of tame spaces. Let LL be a field.

  1. (1)

    Henselizations of localizations of finitely generated LL-algebras are excellent. In particular the local ring

    L[[t1,,tn]]alg={pL[[t1,,tn]]:p algebraic over L(t1,,tn)}L[[t_{1},\ldots,t_{n}]]_{\mathrm{alg}}=\{p\in L[[t_{1},\dotsc,t_{n}]]\colon p\text{ algebraic over }L(t_{1},\dotsc,t_{n})\}

    is excellent. (This is the Henselization of the localization of L[t1,,tn]L[t_{1},\dotsc,t_{n}] at the maximal ideal (t1,,tk)(t_{1},\dotsc,t_{k}): henselianity of L[[t1,,tn]]algL[[t_{1},\dotsc,t_{n}]]_{\mathrm{alg}} follows immediately from henselianity of L[[t1,,tn]]L[[t_{1},\dotsc,t_{n}]], and conversely the Henselization of L[t1,,tn]L[t_{1},\dotsc,t_{n}] at (t1,,tn)(t_{1},\dotsc,t_{n}) is algebraically closed in the completion [Nag75, Corollary 44.3].) When LL is real closed L[[t1,,tn]]algL[[t_{1},\ldots,t_{n}]]_{\mathrm{alg}} is the ring of germs of nn-variable Nash functions at the origin [BCR98, Corollary 8.1.6].

  2. (2)

    Complete Noetherian local rings, such as L[[t1,,tk]]L[[t_{1},\dotsc,t_{k}]] and its quotients, are excellent.

  3. (3)

    If LL is complete with respect to a norm the ring of covergent power series L{t1,,tn}L\{t_{1},\ldots,t_{n}\} in nn-variables is an excellent local ring. (See [Nag75, Theorem 45.5] for henselianity, [Mat80, (34.B)] for excellence in the case of L=L=\mathbb{R} or L=L=\mathbb{C}, and [Duc09, Théorème 2.13] for excellence in the non-archimedean case.) Quotients of {t1,,tn}\mathbb{C}\{t_{1},\ldots,t_{n}\} arise as local rings of complex analytic varieties and when LL is non-archimedean quotients of L{t1,,tn}L\{t_{1},\ldots,t_{n}\} arise as local rings of Berkovich spaces, see [Duc09].

Fact 2.13.

Let RR be a Noetherian henselian local domain. Then the integral closure SS of RR in a finite extension LL of its fraction field KK is also a henselian local domain.

Proof.

The integral closure SS is a direct limit of domains which are finite over RR. Any domain finite over RR is itself a henselian and local by the characterization [Sta20, Tag 04GG (10)] of henselianity, and the class of henselian local domains is closed under direct limits. ∎

2.5. Resolution of Singularities

A resolution of singularities of a reduced Noetherian scheme WW is given by a regular scheme VV and a proper birational morphism VWV\to W. A resolution of singularities of a Noetherian ring RR is a resolution of singularities of SpecR\operatorname{Spec}R.

Fact 2.14 is related to the fact that a one-dimensional reduced KK-variety admits a resolution of singularities.

Fact 2.14.

Suppose that RR is a one-dimensional Noetherian domain and let SS be the integral closure of RR in K=Frac(R)K=\operatorname{Frac}(R). Then the following are equivalent:

  1. (1)

    SpecR\operatorname{Spec}R admits a resolution of singularities.

  2. (2)

    RR is NN-1 (i.e. SS is a finite RR-module).

  3. (3)

    the natural morphism π:SpecSSpecR\uppi\colon\operatorname{Spec}S\to\operatorname{Spec}R is a resolution of singularities for SpecR\operatorname{Spec}R.

Recall that if TT is a finite extension of RR in KK then SpecTSpecR\operatorname{Spec}T\to\operatorname{Spec}R is birational.

Proof.

By Fact 2.9 SS is regular. If (2)(2) holds then π\uppi is finite, hence proper and birational. So (2) implies (3). Clearly (3) implies (1). See [Cut04, Section 2.4, p. 11, last paragraph before Exercise 2.15] for a proof that (1) implies (2). ∎

Fact 2.15 is a famous theorem of Hironaka [Hir64] (cited as in [Tem13, 1.2 (i)]).

Fact 2.15.

Suppose that RR is a quasi-excellent local domain of residue characteristic zero. Then any reduced scheme of finite type over RR admits a resolution of singularities. In particular RR admits a resolution of singularities.

Fact 2.15 in positive residue characteristic is of course an open conjecture [Gro65, 7.9.6]. We use a weaker form of resolution of singularities due to Gabber. Suppose that RR is a Noetherian domain. An altered local uniformization222The terminology is borrowed from [Tem13, 1.2 (iv)]. of RR consists of regular integral schemes V1,,VnV_{1},\dotsc,V_{n} and generically finite dominant morphisms ViSpecRV_{i}\to\operatorname{Spec}R of finite type such that every valuation ring 𝒪\mathcal{O} containing RR can be prolonged to a valuation ring 𝒪\mathcal{O}^{*} centered on some ViV_{i}, i.e. there exists a commutative diagram as follows:

Spec𝒪\textstyle{\operatorname{Spec}\mathcal{O}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Vi\textstyle{V_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec𝒪\textstyle{\operatorname{Spec}\mathcal{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SpecR\textstyle{\operatorname{Spec}R}

The valuative criterion for properness implies that a resolution of singularities is an altered local uniformization.

Theorem 2.16 (Gabber).

A quasi-excellent domain admits an altered local uniformization.

Proof.

By [ILO14, Exposé VII, Théorème 1.1], there are regular integral schemes V1,,VnV_{1},\ldots,V_{n} and finite type morphisms πi:ViSpecR\uppi_{i}\colon V_{i}\to\operatorname{Spec}R such that π1,,πn\uppi_{1},\ldots,\uppi_{n} are a covering family in the Grothendieck topology of alterations [ILO14, Exposé II, 2.3.3]. In particular, each πi\uppi_{i} is dominant and generically finite. The prolongation property for valuation rings follows from [ILO14, Exposé IV, Théorème 4.2.1]. ∎

There are non-quasi-excellent Noetherian local domains which admit an altered local uniformization. For instance, this is trivially the case for regular local rings which are not quasi-excellent, see for example [Mat80, Chapter 13 (34.B)].

3. The one dimensional case

Let RKR\subsetneq K be a domain with fraction field KK.

Lemma 3.1.

Suppose that RR^{*} is a domain with Frac(R)=K\operatorname{Frac}(R^{*})=K. The following are equivalent:

  1. (1)

    The RR-adic topology on KK refines the RR^{*}-adic topology,

  2. (2)

    RR^{*} is RR-adically open,

  3. (3)

    αRR\alpha R\subseteq R^{*} for some αK×\alpha\in K^{\times},

  4. (4)

    RR is bounded in the RR^{*}-adic topology.

Hence the RR-adic topology agrees with the RR^{*}-adic topology if and only if there are α,βK×\alpha,\beta\in K^{\times} such that αRR\alpha R\subseteq R^{*} and βRR\beta R^{*}\subseteq R, i.e. if R,RR,R^{*} is RR^{*}-,RR-adically bounded, respectively.

Lemma 3.1 follows easily from the definitions, so we leave it to the reader.

Lemma 3.2.

Suppose that RR is Noetherian and SS is a subring of KK containing RR. Then the following are equivalent:

  1. (1)

    SS is a finite RR-module,

  2. (2)

    the RR-adic and SS-adic topologies on KK agree.

Proof.

Suppose (2). By Lemma 3.1 there is αK×\alpha\in K^{\times} with αSR\alpha S\subseteq R. As RR is Noetherian αS\alpha S is a finite RR-module, so SS is a finite RR-module. Suppose (1). By Lemma 3.1 it is enough to show αSR\alpha S\subseteq R for some αK×\alpha\in K^{\times}. We have S=β1R++βnRS=\beta_{1}R+\ldots+\beta_{n}R for β1,,βnK\beta_{1},\ldots,\beta_{n}\in K. Fix αK\alpha\in K with αβiR\alpha\beta_{i}\in R for all ii. Then αS=(αβ1)R++(αβn)RR\alpha S=(\alpha\beta_{1})R+\ldots+(\alpha\beta_{n})R\subseteq R. ∎

Lemma 3.3 is immediate from Lemma 3.2.

Lemma 3.3.

Suppose that RR is Noetherian and SS is the integral closure of RR in KK. Then the following are equivalent:

  1. (1)

    RR is NN-1,

  2. (2)

    the RR-adic and SS-adic topologies on KK agree.

Proposition 3.4 is a partial converse to our theorem that if RR is an excellent henselian local domain then the RR-adic and K\mathscr{E}_{K}-topologies agree. (Recall that an excellent ring is NN-1.)

Proposition 3.4.

Suppose that RR is a Noetherian, henselian, and local, and the K\mathscr{E}_{K}-topology agrees with the RR-adic topology. Then RR is NN-11.

Proof.

Let SS be the integral closure of RR in KK. As RSR\subseteq S, the RR-adic topology refines the SS-adic topology. By Fact 2.13 SS is a henselian local ring. By Fact 1.1(1) the SS-adic topology on KK refines the K\mathscr{E}_{K}-topology. Hence the SS-adic topology agrees with the RR-adic topology. Apply Lemma 3.3. ∎

Corollary 3.5.

Suppose that RR is one-dimensional, Noetherian, and henselian local. Then the following are equivalent:

  1. (1)

    the K\mathscr{E}_{K}-topology agrees with the RR-adic topology.

  2. (2)

    RR is NN-1.

  3. (3)

    SpecR\operatorname{Spec}R admits a resolution of singularities.

Proof.

The equivalence of (2) and (3) is Fact 2.14. Proposition 3.4 shows that (1) implies (2). Suppose that RR is NN-1 and let SS be the integral closure of RR in KK. Then SS is itself local by Fact 2.13. By Lemma 3.3 it is enough to show that the K\mathscr{E}_{K}-topology agrees with the SS-adic topology. By Facts 2.13 and 1.1(1) the SS-adic topology refines the K\mathscr{E}_{K}-topology. By Facts 2.9 and 1.1(2) the K\mathscr{E}_{K}-topology refines the SS-adic topology. ∎

4. The quasi-excellent case

Let again RKR\subsetneq K be a domain with fraction field KK. The central result of this section is the following theorem, which generalizes Fact 1.1(3).

Theorem 4.1.

Suppose that RR is local, normal, and Noetherian. If RR admits an altered local uniformization, then the étale open topology over KK refines the RR-adic topology.

We gather some lemmas. Fact 4.2 is a slight generalization given in [JWY21, Lemma 4.3] of a result of Jensen and Lenz- ing [JL89, pg 52,55].

Fact 4.2.

Suppose that RR is a regular local domain with maximal ideal 𝔪\mathfrak{m} and dimR2\dim R\geq 2.

  1. (1)

    If Char(R/𝔪)2\operatorname{Char}(R/\mathfrak{m})\neq 2 and α,βK\alpha,\beta\in K satisfy 1+α4=β21+\alpha^{4}=\beta^{2} then αR\alpha\in R or 1/αR1/\alpha\in R.

  2. (2)

    If Char(R/𝔪)=2\operatorname{Char}(R/\mathfrak{m})=2 and α,βK\alpha,\beta\in K satisfy 1+α3=β31+\alpha^{3}=\beta^{3} then αR\alpha\in R or 1/αR1/\alpha\in R.

Lemma 4.3.

Suppose that RR is a local domain, O1,,OkO_{1},\ldots,O_{k} are discrete valuation subrings of KK with maximal ideals 𝔪1,,𝔪k\mathfrak{m}_{1},\ldots,\mathfrak{m}_{k}, respectively. Let 𝔞=𝔪1𝔪k\mathfrak{a}=\mathfrak{m}_{1}\cap\ldots\cap\mathfrak{m}_{k}. Suppose that one of the following holds:

  1. (1)

    Char(K)2\operatorname{Char}(K)\neq 2 and if α𝔞,βK\alpha\in\mathfrak{a},\beta\in K satisfy 1+α4=β21+\alpha^{4}=\beta^{2} then αR\alpha\in R,

  2. (2)

    Char(K)3\operatorname{Char}(K)\neq 3 and if α𝔞,βK\alpha\in\mathfrak{a},\beta\in K satisfy 1+α3=β31+\alpha^{3}=\beta^{3} then αR\alpha\in R,

Then the étale open topology over KK refines the RR-adic topology.

Proof.

Let us assume that Char(K)2\operatorname{Char}(K)\neq 2. By Fact 2.5 it is enough to show that RR has interior in the K\mathscr{E}_{K}-topology. Each 𝔪i\mathfrak{m}_{i} is K\mathscr{E}_{K}-open by Fact 1.1(2). (The condition there that the henselization of KK with respect to OiO_{i} is not separably closed holds since OiO_{i} is discrete; cf. also [JTWY22, Corollary 6.17].) Hence 𝔞\mathfrak{a} is K\mathscr{E}_{K}-open. Let B={β2:βK×}B=\{\beta^{2}:\beta\in K^{\times}\} and f:KKf\colon K\to K be given by f(α)=1+α4f(\alpha)=1+\alpha^{4}. By Fact 2.1 f1(B)f^{-1}(B) is K\mathscr{E}_{K}-open. Note that f1(B)𝔞f^{-1}(B)\cap\mathfrak{a} is contained in RR by (1)(1) in the assumption. Finally f1(B)𝔞f^{-1}(B)\cap\mathfrak{a} is nonempty as f1(B)f^{-1}(B) and each 𝔪i\mathfrak{m}_{i} is an K\mathscr{E}_{K}-neighbourhood of zero. The argument for Char(K)3\operatorname{Char}(K)\neq 3 is analogous. ∎

Lemma 4.4.

Suppose that RR is normal, local, and Noetherian, and αKR\alpha\in K\setminus R. Then there exists a valuation ring 𝒪\mathcal{O} of KK dominating RR with α𝒪\alpha\not\in\mathcal{O}.

Proof.

By normality there is a height one prime ideal 𝔭\mathfrak{p} in RR such that α𝒪𝔭\alpha\notin\mathcal{O}_{\mathfrak{p}}, see [Mat80, Chapter 7 (17.H) Theorem 38]. Then 𝒪𝔭\mathcal{O}_{\mathfrak{p}} is normal by Fact 2.11(1), and so 𝒪𝔭\mathcal{O}_{\mathfrak{p}} is a DVR since it is a one-dimensional normal local domain [Sta20, Tag 00PD]. By Chevalley’s extension theorem there is a valuation ring 𝒪\mathcal{O}^{*} of Frac(𝒪/𝔭)\operatorname{Frac}({\mathcal{O}/\mathfrak{p}}) dominating the local ring R/𝔭R/\mathfrak{p}. Let 𝒪\mathcal{O} be the valuation ring corresponding to the composition of the places associated to 𝒪𝔭\mathcal{O}_{\mathfrak{p}} and 𝒪\mathcal{O}^{*}. Then α𝒪\alpha\not\in\mathcal{O} since 𝒪𝒪𝔭\mathcal{O}\subseteq\mathcal{O}_{\mathfrak{p}}, and by construction 𝒪\mathcal{O} dominates RR. ∎

We now prove Theorem 4.1.

Proof.

Let Π=(XiπiSpecR:i{1,,n})\Pi=(X_{i}\xrightarrow[]{\uppi_{i}}\operatorname{Spec}R:i\in\{1,\ldots,n\}) be an altered local uniformization of RR. We make some definitions and constructions for arbitrary i{1,,n}i\in\{1,\ldots,n\}. Recall that XiX_{i} is integral and let KiK_{i} be the function field of XiX_{i}. Let 𝔪\mathfrak{m} be the maximal ideal of RR. Then 𝔪\mathfrak{m} is the closed point of SpecR\operatorname{Spec}R, hence πi1(𝔪)\uppi^{-1}_{i}(\mathfrak{m}) is a proper closed subset of XiX_{i} by dominance of πi\uppi_{i}. The set πi1(𝔪)\uppi^{-1}_{i}(\mathfrak{m}) has only finitely many irreducible components, each of which is contained in an irreducible codimension one subset of XiX_{i}. Let AiA_{i} be a finite set of codimension one points in XiX_{i} such that every point in πi1(𝔪)\uppi^{-1}_{i}(\mathfrak{m}) is in the closure of some pAip\in A_{i}. By regularity 𝒪Xi,p\mathcal{O}_{X_{i},p} is a DVR for every pAip\in A_{i}. Let 𝒪i,p=𝒪Xi,pK\mathcal{O}_{i,p}=\mathcal{O}_{X_{i},p}\cap K for every i{1,,n}i\in\{1,\ldots,n\} and pAip\in A_{i}. Since the extension Ki/KK_{i}/K is finite, 𝒪i,p\mathcal{O}_{i,p} is a (non-trivial) DVR. Let 𝔪i,p\mathfrak{m}_{i,p} be the maximal ideal of each 𝒪i,p\mathcal{O}_{i,p}.

Now suppose first that Char(R/𝔪)2\operatorname{Char}(R/\mathfrak{m})\neq 2, hence Char(K)2\operatorname{Char}(K)\neq 2. Suppose that α,βK\alpha,\beta\in K satisfy 1+α4=β21+\alpha^{4}=\beta^{2} and α\alpha is in 1inpAi𝔪i,p\bigcap_{1\leq i\leq n}\bigcap_{p\in A_{i}}\mathfrak{m}_{i,p}. By Lemma 4.3 it is enough to show that αR\alpha\in R. We suppose towards a contradiction that αR\alpha\not\in R. By Lemma 4.4, there exists a valuation subring 𝒪\mathcal{O} of KK dominating RR with α𝒪\alpha\not\in\mathcal{O}. By the defining property of altered local uniformizations, there is i{1,,n}i\in\{1,\ldots,n\}, pXip\in X_{i}, and a valuation subring 𝒪\mathcal{O}^{*} of KiK_{i} such that 𝒪\mathcal{O}^{*} prolongs 𝒪\mathcal{O} and 𝒪\mathcal{O}^{*} dominates 𝒪Xi,p\mathcal{O}_{X_{i},p}. Thus α𝒪Xi,p\alpha\notin\mathcal{O}_{X_{i},p}. Fact 4.2 shows that 1/α𝒪Xi,p1/\alpha\in\mathcal{O}_{X_{i},p} when dim𝒪Xi,p2\dim\mathcal{O}_{X_{i},p}\geq 2; in fact the same holds if dim𝒪Xi,p=1\dim\mathcal{O}_{X_{i},p}=1, since then 𝒪Xi,p\mathcal{O}_{X_{i},p} is a valuation ring. Since πi(p)=𝔪\uppi_{i}(p)=\mathfrak{m} as 𝒪\mathcal{O}^{*} dominates RR, by construction we can take qAiq\in A_{i} such that 𝒪Xi,p𝒪Xi,q\mathcal{O}_{X_{i},p}\subseteq\mathcal{O}_{X_{i},q}. Then 1/α𝒪Xi,qK=𝒪i,q1/\alpha\in\mathcal{O}_{X_{i},q}\cap K=\mathcal{O}_{i,q}, which is a contradiction as α𝔪i,q\alpha\in\mathfrak{m}_{i,q}.

Finally, suppose that Char(R/𝔪)=2\operatorname{Char}(R/\mathfrak{m})=2, hence Char(K)3\operatorname{Char}(K)\neq 3. Follow the same argument as above, replacing 1+X4=Y21+X^{4}=Y^{2} with 1+X3=Y31+X^{3}=Y^{3}, and apply the second case of Lemma 4.3. ∎

Theorem 4.5.

If RR is quasi-excellent local then the K\mathscr{E}_{K}-topology refines the RR-adic topology.

Proof.

By Fact 2.11(4) RR is NN-1, so the integral closure SS of RR in KK is finite over RR. By Lemma 3.3 it is enough to show that the K\mathscr{E}_{K}-topology refines the SS-adic topology. It is enough to show that SS is K\mathscr{E}_{K}-open. Since SS is finite over the local ring RR, SS has only finitely many maximal ideals 𝔪1,,𝔪k\mathfrak{m}_{1},\ldots,\mathfrak{m}_{k}. Let SiS_{i} be the localization of SS at 𝔪i\mathfrak{m}_{i} for each i{1,,k}i\in\{1,\ldots,k\}. Then S=S1SkS=S_{1}\cap\ldots\cap S_{k}, so it is enough to fix i{1,,k}i\in\{1,\ldots,k\} and show that SiS_{i} is K\mathscr{E}_{K}-open. Note that SiS_{i} is a localization of a finite extension of the quasi-excellent ring RR and SiS_{i} is a localization of the normal ring SS. By Fact 2.11 SiS_{i} is quasi-excellent and normal. Theorem 4.1 (which applies by Theorem 2.16) shows that SiS_{i} is K\mathscr{E}_{K}-open. ∎

Remark 4.6.

In [JL89, Theorem 3.35], the henselian case of Fact 4.2 is used to prove that any henselian regular local domain is first-order definable in its fraction field. Whether the same holds for a henselian quasi-excellent local domain RR remains open.

We only obtain the weaker statement that the RR-adic topology is definable in the fraction field KK, i.e. there is a definable family of sets forming a basis for the RR-adic topology: Indeed, we have shown in Lemma 4.3 that there is an étale image UK=𝔸K1(K)\emptyset\neq U\subseteq K=\mathbb{A}^{1}_{K}(K) contained in RR. Since UU is definable and open, the family {aU+b:aK×,bK}\{aU+b\colon a\in K^{\times},b\in K\} is a definable basis for the RR-adic topology.

Essentially the same argument shows that whenever K\mathscr{E}_{K} is induced by a locally bounded field topology τ\uptau (a situation which we shall study later in some detail), the topology τ\uptau is definable.

5. Behaviour of K\mathscr{E}_{K} under field extension

Suppose that L/KL/K is a finite field extension and let [L:K]=d[L:K]=d. We briefly describe the extension ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) of the K\mathscr{E}_{K}-topology to LL, see [JTWY22, Section 4.5] for details. After fixing a KK-basis for LL we may identify each LnL^{n} with KdnK^{dn}. We declare the ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-topology on LnL^{n} to be the K\mathscr{E}_{K}-topology on KdnK^{dn}. This topology does not depend on the choice of the KK-basis. More generally, given a quasi-projective LL-variety VV the ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-topology on V(L)V(L) is the K\mathscr{E}_{K}-topology on the KK-points of the Weil restriction of VV (this set is canonically identified with V(L)V(L).) Any variety is Zariski-locally quasi-projective, so we can define the ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-topology on the KK-points of an arbitrary KK-variety in a natural way. For example Ext/()\operatorname{Ext}_{\mathbb{C}/\mathbb{R}}(\mathscr{E}_{\mathbb{R}}) is the usual complex analytic topology over \mathbb{C}.

Endowing all V(L)V(L) with the ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-topology gives a well-behaved system of topologies in the sense of [JTWY22, Definition 1.2], see the following consequence of [JTWY22, Proposition-Definition 4.17].

Fact 5.1.

Suppose that L/KL/K is finite and VWV\to W is a morphism of LL-varieties. Then V(L)W(L)V(L)\to W(L) is ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-continuous. In particular LL,xαx+βL\to L,x\mapsto\alpha x+\beta is an ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-homeomorphism for any αL×\alpha\in L^{\times}, βL\beta\in L.

By [JTWY22, Proposition 5.7] ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) refines L\mathscr{E}_{L} for any finite L/KL/K. We would like to know when this refinement is strict.

Up to now we knew two examples. If KK is real closed and L=K(1)L=K(\sqrt{-1}) then L\mathscr{E}_{L} is the Zariski topology and K\mathscr{E}_{K} is the order topology, hence ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) strictly refines L\mathscr{E}_{L}. Recall that the following are equivalent by [JTWY22, Theorem C.1]:

  1. (1)

    LL is large,

  2. (2)

    the L\mathscr{E}_{L}-topology on LL is not discrete,

  3. (3)

    the L\mathscr{E}_{L}-topology on V(L)V(L) is not discrete when VV is an LL-variety with V(L)V(L) infinite.

By [Sri19] there are non-large fields with large finite extensions. If LL is large and KK is not then the K\mathscr{E}_{K}-topology on KdK^{d} is discrete, hence the ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-topology on LL is discrete, and the L\mathscr{E}_{L}-topology on LL is not discrete.

We give a third example where ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) strictly refines L\mathscr{E}_{L}. This is also the first example where both K,L\mathscr{E}_{K},\mathscr{E}_{L} are non-discrete field topologies.

Theorem 5.2.

Let RR be a henselian regular local domain and LL a finite extension of the fraction field KK of RR such that the integral closure SS of RR in LL is not a finite RR-module. Then ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) strictly refines L\mathscr{E}_{L}.

Note that any RR as in the theorem is by definition not Japanese and hence not quasi-excellent. Since regular local rings are normal, Fact 2.11(7) shows that the theorem is only ever applicable in positive characteristic.

Before proving the theorem, we give an important special case in the language of valued fields. See [Kuh11, Example 3.5] for an example of this situation.

Corollary 5.3.

Let vv be a henselian discrete valuation on a field KK, L/KL/K a finite extension and vv^{\prime} the unique prolongation of vv to LL. If (L,v)/(K,v)(L,v^{\prime})/(K,v) is a defect extension, i.e. e(v/v)f(v/v)[L:K]e(v^{\prime}/v)f(v^{\prime}/v)\lneq[L:K] where ee and ff are the relative ramification index and inertia degree, then ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) strictly refines L\mathscr{E}_{L}.

Proof.

Let RR and SS be the valuation rings of vv and vv^{\prime}, respectively. Both are discrete valuation rings, in particular regular local domains. Furthermore, SS is the integral closure of RR in LL [Bou06, Chap. V, §8, no 3, Remarque], and the defect condition implies that SS is not a finite RR-module [Bou06, Chap. V, §8, no 5, Théorème 2]. Hence Theorem 5.2 is applicable. ∎

Proof of Theorem 5.2.

Let b1,,bdLb_{1},\dotsc,b_{d}\in L be a KK-basis of LL. By scaling with a suitable element of RR, we may assume that for all ii, biSb_{i}\in S. Hence R:=R[b1,,bd]R^{\prime}:=R[b_{1},\dotsc,b_{d}] is a finite RR-module. The fraction field of RR^{\prime} is LL, and SS is the normalization of RR^{\prime}. By assumption, SS is not a finite RR^{\prime}-module, and thus by Lemma 3.2 the RR^{\prime}-adic topology on LL strictly refines the SS-adic topology. The SS-adic topology in turn refines L\mathscr{E}_{L} (not necessarily strictly) by Fact 1.1(1), since SS is henselian local by Fact 2.13.

Under the identification of LL with KdK^{d} given by the basis b1,,bdb_{1},\dotsc,b_{d}, the subgroup b1R++bdRRLb_{1}R+\dotsb+b_{d}R\subseteq R^{\prime}\subseteq L is identified with RdKdR^{d}\subseteq K^{d}, which is open in the product topology of dd copies of the RR-adic topology on KK. Since the RR-adic topology on KK coincides with the K\mathscr{E}_{K}-topology by Fact 1.1, this means that b1R++bdRb_{1}R+\dotsb+b_{d}R and hence RR^{\prime} are ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K})-open. By Fact 5.1 it follows that ExtL/K(K)\operatorname{Ext}_{L/K}(\mathscr{E}_{K}) refines the RR^{\prime}-adic topology on LL, which strictly refines L\mathscr{E}_{L}. ∎

6. A large collection of incomparable topologies on p\mathbb{Q}_{p}

Fix a prime pp. In this section we produce 2202^{2^{\aleph_{0}}}-many henselian local subrings RpR\subsetneq\mathbb{Q}_{p} with fraction field p\mathbb{Q}_{p} such that the corresponding RR-adic topologies are pairwise incomparable. This is interesting in light of Corollary 1.4, which shows that this behaviour cannot occur for quasi-excellent RR, since in this case the RR-adic topology induces the étale open topology. It is also in contrast to F. K. Schmidt’s theorem [EP05, Theorem 4.4.1], which shows that any two henselian valuation rings on a field which is not separably closed induce the same topology (compare also Fact 1.1(1, 2)).

Our approach is based on [JWY21, Section 5]. Given a (\mathbb{Q}-)derivation :pp\der\colon\mathbb{Q}_{p}\to\mathbb{Q}_{p} we let EE be {αp:αp}\{\alpha\in\mathbb{Z}_{p}:\der\alpha\in\mathbb{Z}_{p}\}. It is easy to see that EE is a subring of p\mathbb{Z}_{p}. Fact 6.1 is a summary of the statements of [JWY21, Section 5].

Fact 6.1.

If is not identically zero then:

  1. (1)

    EE is a one-dimensional Noetherian henselian local ring with fraction field p\mathbb{Q}_{p}, and

  2. (2)

    the EE-adic topology on p\mathbb{Q}_{p} strictly refines the pp-adic topology.

Furthermore E^\widehat{E} is isomorphic to p[X]/(X2)\mathbb{Z}_{p}[X]/(X^{2}), hence EE is not excellent.

Let 𝒟\mathscr{D} be the set of derivations pp\mathbb{Q}_{p}\to\mathbb{Q}_{p} which are not constant zero. We say that ,𝒟\der,{}^{*}\in\mathscr{D} are constant multiples of each other if λ=\lambda\der={}^{*} for some λp\lambda\in\mathbb{Q}_{p}. We prove:

Theorem 6.2.

If ,𝒟\der,{}^{*}\in\mathscr{D} are not constant multiples of each other, then the EE-adic topology does not refine the EE_{{}^{*}}-adic topology and vice versa. There is I𝒟I\subseteq\mathscr{D} such that |I|=220|I|=2^{2^{\aleph_{0}}} and if ,I\der,{}^{*}\in I, \der\neq{}^{*} then the EE-adic topology does not refine the EE_{{}^{*}}-adic topology.

Thus there are 2202^{2^{\aleph_{0}}}-distinct EE-adic topologies on p\mathbb{Q}_{p}. We explain how the second claim follows from the first. Let BB be a transcendence basis for p\mathbb{Q}_{p}. By the usual rules for extending derivations to separable field extensions [FJ05, Section 2.8], it is easy to see that any function BpB\to\mathbb{Q}_{p} uniquely extends to a derivation pp\mathbb{Q}_{p}\to\mathbb{Q}_{p}. Since |B|=20\lvert B\rvert=2^{\aleph_{0}}, this shows |𝒟|=220\lvert\mathscr{D}\rvert=2^{2^{\aleph_{0}}}. As every element of 𝒟\mathscr{D} is a constant multiple of precisely |p×|=20\lvert\mathbb{Q}_{p}^{\times}\rvert=2^{\aleph_{0}} other elements of 𝒟\mathscr{D}, this shows that there are 2202^{2^{\aleph_{0}}} classes of elements of 𝒟\mathscr{D} under the equivalence relation of being a constant multiple of one another, and we may take I𝒟I\subseteq\mathscr{D} to be a set of representatives for this equivalence relation.

Lemma 6.3.

Suppose that ,:pp\der,{}^{*}\colon\mathbb{Q}_{p}\to\mathbb{Q}_{p} are derivations and neither is a constant multiple of the other. Then {(α,α,α):αp}\{(\alpha,\der\alpha,{}^{*}\alpha):\alpha\in\mathbb{Q}_{p}\} is pp-adically dense in p3\mathbb{Q}^{3}_{p}.

Proof.

As ,\der,{}^{*} are not constant multiples of each other there are s,tps,t\in\mathbb{Q}_{p} such that (s,t)(\der s,\der t) and (s,t)({}^{*}s,{}^{*}t) are not scalar multiples of each other in p2\mathbb{Q}^{2}_{p}. Any derivation pp\mathbb{Q}_{p}\to\mathbb{Q}_{p} is \mathbb{Q}-linear and vanishes on \mathbb{Q}, hence for α=a+sb+tc\alpha=a+sb+tc with a,b,ca,b,c\in\mathbb{Q} we have

(α,α,α)=(a+sb+tc,bs+ct,bs+ct).\displaystyle(\alpha,\der\alpha,{}^{*}\alpha)=(a+sb+tc,b\der s+c\der t,b{}^{*}s+c{}^{*}t).

We let T:p3p3T\colon\mathbb{Q}^{3}_{p}\to\mathbb{Q}^{3}_{p} be the \mathbb{Q}-linear transformation given as follows:

T(xyz)=(x+sy+tz(s)y+(t)z(s)y+(t)z)T\begin{pmatrix}x\\ y\\ z\end{pmatrix}=\begin{pmatrix}x+sy+tz\\ (\der s)y+(\der t)z\\ ({}^{*}s)y+({}^{*}t)z\end{pmatrix}

Note that T(3){(α,α,α):αp}T(\mathbb{Q}^{3})\subseteq\{(\alpha,\der\alpha,{}^{*}\alpha):\alpha\in\mathbb{Q}_{p}\}, so it is enough to show that T(3)T(\mathbb{Q}^{3}) is dense in p3\mathbb{Q}^{3}_{p}. As 3\mathbb{Q}^{3} is dense in p3\mathbb{Q}^{3}_{p} and TT is linear it is sufficient to note that TT is invertible since

det(T)=det(1st0st0st)=det(stst)0.\det(T)=\det\begin{pmatrix}1&s&t\\ 0&\der s&\der t\\ 0&{}^{*}s&{}^{*}t\end{pmatrix}=\det\begin{pmatrix}\der s&\der t\\ {}^{*}s&{}^{*}t\end{pmatrix}\neq 0.\qed
Proof of Theorem 6.2.

Suppose ,\der,{}^{*} are not constant multiples of each other. We show that the EE_{{}^{*}}-adic topology does not refine the EE-adic topology. By Lemma 3.1 is enough to show that aEEaE_{{}^{*}}\nsubseteq E for any ap×a\in\mathbb{Q}^{\times}_{p}. Let

U={(b,b,b′′)p×p×p:ab+(a)bpp}.U=\{(b,b^{\prime},b^{\prime\prime})\in\mathbb{Z}_{p}\times\mathbb{Q}_{p}\times\mathbb{Z}_{p}:ab^{\prime}+(\der a)b\in\mathbb{Q}_{p}\setminus\mathbb{Z}_{p}\}.

Then UU is open and nonempty as (0,(pa)1,0)U(0,(pa)^{-1},0)\in U. By Lemma 6.3 we have (b,b,b)U(b,\der b,{}^{*}b)\in U for some bpb\in\mathbb{Q}_{p}. Then b,bpb,{}^{*}b\in\mathbb{Z}_{p}, so bEb\in E_{{}^{*}}, and (ab)=a(b)+b(a)p\der(ab)=a(\der b)+b(\der a)\notin\mathbb{Z}_{p}, so abEab\notin E. ∎

7. The étale open topology on pseudo-algebraically closed fields

Recall that a field KK is pseudo-algebraically closed (PAC) if every geometrically integral KK-variety has a KK-point.

Proposition 7.1.

Let KK be a PAC field. Then the étale open topology on varieties over KK is not induced by a field topology on KK.

Proof.

Suppose for a contradiction that there is a field topology τ\uptau on KK inducing K\mathscr{E}_{K}. We consider the morphism α:PGL2,K×K1K1\alpha\colon\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\to\mathbb{P}^{1}_{K} given by the natural group action, as well as the projection morphisms π1:PGL2,K×K1PGL2,K\pi_{1}\colon\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\to\operatorname{PGL}_{2,K} and π2:PGL2,K×K1K1\pi_{2}\colon\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\to\mathbb{P}^{1}_{K}. For later use we observe that the morphism (π1,α):PGL2,K×K1PGL2,K×K1(\pi_{1},\alpha)\colon\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\to\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K} is an isomorphism, since it has an obvious inverse given by acting with the inverse group element. In particular, the morphism α=π2(π1,α)\alpha=\pi_{2}\circ(\pi_{1},\alpha) is smooth since π2\pi_{2} is smooth (as it is a base change of the smooth morphism PGL2,KSpecK\operatorname{PGL}_{2,K}\to\operatorname{Spec}K).

The étale open topology on PGL2,K(K)×K1(K)\operatorname{PGL}_{2,K}(K)\times\mathbb{P}^{1}_{K}(K) is the product topology of the étale open topologies on PGL2,K(K)\operatorname{PGL}_{2,K}(K) and K1(K)\mathbb{P}^{1}_{K}(K), since the analogous statement is true for the τ\uptau-topology and the two topologies agree on the KK-points of every variety. Let U1(K)\emptyset\neq U\subseteq\mathbb{P}^{1}(K) be open. We show that UU is necessarily cofinite.

The group scheme action α\alpha induces a map PGL2,K(K)×1(K)1(K)\operatorname{PGL}_{2,K}(K)\times\mathbb{P}^{1}(K)\to\mathbb{P}^{1}(K) on KK-points, which we also denote by α\alpha. It is continuous by the defining properties of the étale open topology, and so there exist non-empty étale open subsets of PGL2,K(K)\operatorname{PGL}_{2,K}(K) and 1(K)\mathbb{P}^{1}(K) whose product is contained in the preimage of UU under α\alpha. In other words, there exist two KK-varieties XX and YY with étale maps XPGL2,KX\to\operatorname{PGL}_{2,K}, YK1Y\to\mathbb{P}^{1}_{K} such that X(K),Y(K)X(K),Y(K)\neq\emptyset and UU contains the image of X(K)×Y(K)X(K)\times Y(K) under the composite

g:X×YPGL2,K×K1𝛼K1.g\colon X\times Y\to\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\overset{\alpha}{\to}\mathbb{P}^{1}_{K}.

The KK-schemes XX and YY are smooth since they are étale over the smooth KK-schemes PGL2,K\operatorname{PGL}_{2,K} and K1\mathbb{P}^{1}_{K}, respectively. Passing to a connected component of XX and YY if necessary, we may additionally assume that both XX and YY are connected (as schemes, i.e. not in relation to the topologies K\mathscr{E}_{K} or τ\uptau). Since they both have a KK-point, XX and YY are then geometrically connected [Poo17, Proposition 2.3.24] and hence (by smoothness) geometrically integral [Poo17, Proposition 3.5.67].

We claim that the generic fibre of gg is geometrically integral (as a variety over the function field K(K1)K(\mathbb{P}^{1}_{K})), i.e. that the function field K(X×Y)K(X\times Y) is a regular extension of the function field K(K1)K(\mathbb{P}^{1}_{K}) via the map gg. Let us defer the proof of this claim for the moment. By [Sta20, Tags 0578 and 0559], all but finitely many fibres of gg are geometrically integral. In particular, for all but finitely many x1(K)x\in\mathbb{P}^{1}(K), the KK-variety g1(x)g^{-1}(x) has a KK-point by the PAC property, and thus xg(X(K)×Y(K))Ux\in g(X(K)\times Y(K))\subseteq U.

This shows that UU is cofinite. Thus the étale open topology on K=𝔸K1(K)K1(K)K=\mathbb{A}^{1}_{K}(K)\subseteq\mathbb{P}^{1}_{K}(K), and therefore the topology τ\uptau, is the cofinite topology. Since the cofinite topology is not a field topology on any infinite field, this yields the desired contradiction.

It remains to prove the claim. This is purely a matter of algebraic geometry, so the topologies K\mathscr{E}_{K} and τ\uptau no longer intervene. As a consequence of Zariski’s Main Theorem, we can embed XX and YY as open subschemes of normal integral schemes X¯\overline{X}, Y¯\overline{Y} with finite morphisms p1:X¯PGL2,Kp_{1}\colon\overline{X}\to\operatorname{PGL}_{2,K}, p2:Y¯K1p_{2}\colon\overline{Y}\to\mathbb{P}^{1}_{K} extending the étale morphisms from XX respectively YY. (See for instance [Poo17, Theorem 3.5.52 (c)] (recalling that XX and YY are separated by our convention on varieties), where X¯\overline{X} and Y¯\overline{Y} are described concretely as normalisations of PGL2,K\operatorname{PGL}_{2,K} (respectively K1\mathbb{P}^{1}_{K}) in the function field of XX (respectively YY).)

Via the dominant morphism X¯×Y¯p1×p2PGL2,K×K1𝛼K1\overline{X}\times\overline{Y}\overset{p_{1}\times p_{2}}{\longrightarrow}\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\overset{\alpha}{\to}\mathbb{P}^{1}_{K}, which restricts to the morphism gg considered earlier on X×YX\times Y, we can consider K(X×Y)=K(X¯×Y¯)K(X\times Y)=K(\overline{X}\times\overline{Y}) as an extension field of K(K1)K(\mathbb{P}^{1}_{K}). Let FK(X×Y)F\subseteq K(X\times Y) be the relative algebraic closure of K(K1)K(\mathbb{P}^{1}_{K}) therein. Then F/KF/K is regular since K(X×Y)/KK(X\times Y)/K is regular, due to the geometric integrality of XX and YY. Let CK1C\to\mathbb{P}^{1}_{K} be the normalisation of K1\mathbb{P}^{1}_{K} in FF. Thus C/KC/K is a geometrically integral normal projective curve and CK1C\to\mathbb{P}^{1}_{K} is a finite morphism. We shall show using a ramification argument that in fact CK1C\to\mathbb{P}^{1}_{K} is an isomorphism.

Let us consider the following diagram:

X¯×Y¯\textstyle{\overline{X}\times\overline{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1×p2\scriptstyle{p_{1}\times p_{2}}PGL2,K×K1\textstyle{\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(π1,α)\scriptstyle{(\pi_{1},\alpha)}PGL2,K×C\textstyle{\operatorname{PGL}_{2,K}\times C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PGL2,K×K1\textstyle{\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}}

All varieties occurring are geometrically integral and normal, the vertical morphism on the right is an isomorphism, the top horizontal morphism is finite and generically étale, and the bottom morphism (given by the identity on PGL2,K\operatorname{PGL}_{2,K} and the previous map CK1C\to\mathbb{P}^{1}_{K}) is finite.

We can complete the diagram by a finite morphism on the left side, shown as a dashed arrow: Observe first that by construction, the function field K(X¯×Y¯)K(\overline{X}\times\overline{Y}) is an extension of the function field of PGL2,K×C\operatorname{PGL}_{2,K}\times C, i.e. we can find a rational function on the left side making the diagram commute. In particular, we then have a normalisation of PGL2,K×C\operatorname{PGL}_{2,K}\times C in the function field K(X¯×Y¯)K(\overline{X}\times\overline{Y}) (see for instance [Liu02, Definition 4.1.24]), which is also a normalization of PGL2,K×K1\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K} in this field by construction. However, the morphism (π1,α)(p1×p2)(\pi_{1},\alpha)\circ(p_{1}\times p_{2}) already describes X¯×Y¯\overline{X}\times\overline{Y} as the normalisation of PGL2,K×K1\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K} within K(X¯×Y¯)K(\overline{X}\times\overline{Y}); therefore, by uniqueness of normalisations, X¯×Y¯\overline{X}\times\overline{Y} must already be the normalisation of PGL2,K×C\operatorname{PGL}_{2,K}\times C in K(X¯×Y¯)K(\overline{X}\times\overline{Y}), and the morphism on the left side of the diagram making it commutative is none other but the normalisation morphism.

Let us show that the morphism of curves CK1C\to\mathbb{P}^{1}_{K} is unramified. First observe that the only prime divisors of PGL2,K×K1\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K} which ramify under the map p1×p2p_{1}\times p_{2} are of the form D×K1D\times\mathbb{P}^{1}_{K} or PGL2,K×D\operatorname{PGL}_{2,K}\times D^{\prime}, where DD ramifies under p1p_{1} or DD^{\prime} ramifies under p2p_{2}. Since α\alpha is a transitive group action, the image of such a prime divisor under the automorphism (π1,α)(\pi_{1},\alpha) of PGL2,K×K1\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K} is never of the form PGL2,K×{x}\operatorname{PGL}_{2,K}\times\{x\} for a closed point (i.e., prime divisor) xx of K1\mathbb{P}^{1}_{K}. In other words, for every closed point xx of K1\mathbb{P}^{1}_{K}, the prime divisor PGL2,K×{x}\operatorname{PGL}_{2,K}\times\{x\} does not ramify along the map (π1,α)(p1×p2):X¯Y¯PGL2,K×K1(\pi_{1},\alpha)\circ(p_{1}\times p_{2})\colon\overline{X}\to\overline{Y}\to\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}. Due to the commutative diagram above, it follows that the prime divisor in question cannot ramify along PGL2,K×CPGL2,K×K1\operatorname{PGL}_{2,K}\times C\to\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K} either, and so xx is not a branch point of C×K1C\times\mathbb{P}^{1}_{K}. Since xx was arbitrary, this shows that CK1C\to\mathbb{P}^{1}_{K} is unramified. Since CC is a geometrically integral projective curve and K1\mathbb{P}^{1}_{K} is geometrically simply connected (see [Liu02, Corollary 7.4.20]), it follows that the map CK1C\to\mathbb{P}^{1}_{K} is an isomorphism, and so F=K(C)=K(K1)F=K(C)=K(\mathbb{P}^{1}_{K}). In other words, the field K(K1)K(\mathbb{P}^{1}_{K}) is relatively algebraically closed in K(X×Y)K(X\times Y).

Finally, the morphism gg is smooth, since it factors as the composition of the étale morphism X×YPGL2,K×K1X\times Y\to\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}, and the smooth morphism α\alpha. Smoothness of gg at the generic point means that K(X×Y)/K(K1)K(X\times Y)/K(\mathbb{P}^{1}_{K}) is a separable field extension, so (together with relative algebraic closedness) we have shown that it is a regular field extension. This finishes the proof of the claim that the generic fibre of gg is geometrically integral. ∎

Remark 7.2.

The precise choice of the morphism PGL2,K×K1K1\operatorname{PGL}_{2,K}\times\mathbb{P}^{1}_{K}\to\mathbb{P}^{1}_{K} in the proof above is not very important. We only used that it is a transitive group action on a geometrically simply connected variety. In characteristic zero, one can instead use the simpler addition action 𝔸K1×𝔸K1𝔸K1\mathbb{A}^{1}_{K}\times\mathbb{A}^{1}_{K}\to\mathbb{A}^{1}_{K}, but in positive characteristic 𝔸K1\mathbb{A}^{1}_{K} is not geometrically simply connected.

8. gt-henselian field topologies

8.1. Background on topological fields

We develop the basics of a theory of gt-henselian field topologies extending the Prestel-Ziegler theory of t-henselian field topologies. Recall our convention that all field topologies are Hausdorff and non-discrete. Throughout, we fix such a field topology τ\uptau on the field KK.

Definition 8.1.

We say that τ\uptau is generalized (topologically) henselian, for short gt-henselian, if for every nn and every neighbourhood PKP\subseteq K of 1-1 there is a neighbourhood OKO\subseteq K of zero such that the polynomial Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in PP for any a0,,an1Oa_{0},\ldots,a_{n-1}\in O.

As the terminology suggests, gt-henselianity generalizes t-henselianity. For more on t-henselianity, see [PZ78, Section 7].

Remark 8.2.

In fact, the field topology τ\uptau is gt-henselian if and only if KK is τ\uptau-henselian in the sense considered in [Pop14, Examples 1.7], as follows from the characterization we give in Proposition 8.6(4) below. However, Pop’s notion of τ\uptau-henselian rings does not seem to have been studied in any depth in the literature. We prefer the name gt-henselianity to stress the link with [PZ78].

Another notion of henselianity for rings in the literature is given by the henselian semi-normed rings of [FP11] (on which Pop’s definition of weak τ\uptau-henselianity is modelled), but there do not appear to be interesting examples of field topologies obtained in this way, except in the well-known case of a field with an absolute value.

Recall from [PZ78, Theorem 7.2 a)] (which we may as well take as a definition) that the field topology τ\uptau on KK is t-henselian if and only if it is a VV-topology (see [PZ78, Section 3]) and for every n1n\geq 1 there is a τ\uptau-neighbourhood UU of 0 such that any polynomial f=Xn+Xn1+an2Xn2++a1X+a0K[X]f=X^{n}+X^{n-1}+a_{n-2}X^{n-2}+\dotsb+a_{1}X+a_{0}\in K[X] with an2,,a0Ua_{n-2},\dotsc,a_{0}\in U has a zero in KK. We show that gt-henselianity generalises t-henselianity.

Proposition 8.3.

The topology τ\uptau is tt-henselian if and only if it is gtgt-henselian and a V-topology.

For the proof we need the following fact, a special case of the polynomial implicit function theorem for t-henselian fields [PZ78, Theorem 7.4]. We can also prove a polynomial implicit function theorem for locally bounded gt-henselian field topologies, but we will not do so here.

Fact 8.4.

Suppose that τ\uptau is t-henselian, fK[Y1,,Yn,X]f\in K[Y_{1},\ldots,Y_{n},X], and (α,β)Kn×K(\alpha,\beta)\in K^{n}\times K is such that f(α,β)=0f/Xf(α,β)f(\alpha,\beta)=0\neq\der f/\der Xf(\alpha,\beta). Then there are τ\uptau-neighbourhoods U1KnU_{1}\subseteq K^{n}, U2KU_{2}\subseteq K of α,β\alpha,\beta, respectively, and a τ\uptau-continuous function g:U1U2g\colon U_{1}\to U_{2} such that

{(a,g(a)):aU1}={(a,b)U1×U2:f(a,b)=0}\{(a,g(a)):a\in U_{1}\}=\{(a,b)\in U_{1}\times U_{2}:f(a,b)=0\}
Proof of Proposition 8.3.

It follows directly from the definitions that a gt-henselian V-topology is t-henselian. Suppose that τ\uptau is t-henselian. Then τ\uptau is necessarily a V-topology. Fix n1n\geq 1 and a neighbourhood PKP\subseteq K of 1-1. We let fK[Y1,,Yn1,X]f\in K[Y_{1},\ldots,Y_{n-1},X] be the polynomial Xn+1+Xn+Yn1Xn++Y1X+Y0X^{n+1}+X^{n}+Y_{n-1}X^{n}+\ldots+Y_{1}X+Y_{0}. Then we have f(0,,0,1)=0f/X(0,,0,1)f(0,\ldots,0,-1)=0\neq\der f/\der X(0,\ldots,0,-1). Let U1U_{1}, U2U_{2}, and gg be as in Fact 8.4. Let O=g1(PU2)O=g^{-1}(P\cap U_{2}). Then OO is a neighbourhood of zero. By construction, if a0,,an1Oa_{0},\ldots,a_{n-1}\in O then Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in PP. ∎

A significant set of examples for gt-henselian field topologies is furnished by RR-adic topologies for RR henselian.

Proposition 8.5.

Let RKR\subsetneq K be a henselian local domain with fraction field KK. Then the RR-adic topology is gt-henselian.

Proof.

Let PKP\subseteq K be an RR-adic neighbourhood of 1-1. Then PP contains 1+αR-1+\alpha R for some αK×\alpha\in K^{\times}. By multiplying α\alpha with a suitable element of RR, we may assume that αR\alpha\in R and α\alpha is not a unit. It now suffices to show that for every nn and all a0,,an1αRa_{0},\dotsc,a_{n-1}\in\alpha R, the polynomial Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\dotsb+a_{1}X+a_{0} has a root in 1+αR-1+\alpha R. This precisely means that (R,αR)(R,\alpha R) is a henselian pair (see the characterization in [Sta20, Tag 09XI (5)]), which follows from [Sta20, Tag 0DYD] since (R,𝔪)(R,\mathfrak{m}) is a henselian pair (where 𝔪\mathfrak{m} is the maximal ideal of RR). ∎

We let Poln\mathrm{Pol}_{n} be the KK-variety parameterizing degree nn monic polynomials, so Poln\mathrm{Pol}_{n} is just a copy of 𝔸n\mathbb{A}^{n}. Recall that αK\alpha\in K is a simple root of fK[X]f\in K[X] if f(α)=0f(\alpha)=0 and f(α)0f^{\prime}(\alpha)\neq 0.

Proposition 8.6.

The following are equivalent:

  1. (1)

    τ\uptau is gt-henselian.

  2. (2)

    For any nn and neighbourhood PP of 11 there is a neighbourhood OO of 0 such that if a0,,an1Oa_{0},\ldots,a_{n-1}\in O then Xn+1Xn+an1Xn1++a1X+a0X^{n+1}-X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in PP.

  3. (3)

    For any nn and neighbourhood PP of 1-1 there is a neighbourhood OO of 0 such that if c2,,cnOc_{2},\ldots,c_{n}\in O then 1+X+c2X2++cnXn1+X+c_{2}X^{2}+\ldots+c_{n}X^{n} has a root in PP.

  4. (4)

    If αK\alpha\in K is a simple root of a monic polynomial fK[X]f\in K[X], degf=n\deg f=n, and PKP\subseteq K is a neighbourhood of α\alpha then there is a neighbourhood OPoln(K)O\subseteq\mathrm{Pol}_{n}(K) of ff such that every fOf^{*}\in O has a simple root in PP.

  5. (5)

    V(K)W(K)V(K)\to W(K) is τ\uptau-open for any étale morphism VWV\to W.

  6. (6)

    V(K)W(K)V(K)\to W(K) is τ\uptau-open for any smooth morphism VWV\to W.

Definition 8.7.

A basic standard étale morphism is a morphism π:VW\uppi\colon V\to W where WW is an affine KK-variety, VV is the subvariety of W×𝔸1W\times\mathbb{A}^{1} given by f=0gf=0\neq g for f,g(K[W])[X]f,g\in(K[W])[X] such that ff is monic, f/X0\der f/\der X\neq 0 on VV, and π\uppi is the restriction of the projection W×𝔸1WW\times\mathbb{A}^{1}\to W to VV. A standard étale morphism is a morphism π:VW\uppi\colon V\to W of KK-varieties such that there is a KK-variety isomorphism ρ:VV\uprho\colon V^{*}\to V with πρ:VW\uppi\circ\uprho\colon V^{*}\to W basic standard étale.

Fact 8.8 is [Sta20, Tag 02GT].

Fact 8.8.

Any étale morphism of KK-varieties is locally standard étale. That is, if VWV\to W is an étale morphism of KK-varieties and pVp\in V then there is a Zariski open neighbourhood VVV^{*}\subseteq V of pp and an affine Zariski open neighbourhood WWW^{*}\subseteq W of f(p)f(p) such that f(W)Vf(W^{*})\subseteq V^{*} and VWV^{*}\to W^{*} is standard étale.

In the following proof, we work with respect to τ\uptau throughout.

Proof of Proposition 8.6.

The equivalence of (1) and (3) is clear by considering the substitution Y=1/XY=1/X. The equivalence of (1) and (2) is likewise clear by considering the substitution Y=XY=-X. The implication from (6) to (5) is clear since étale morphisms are smooth, and the converse holds since a smooth morphism is locally the composition of an étale morphism and a product projection [Sta20, Tag 054L], see also [WY23, Proposition 3.1].

We show that (4) implies (5). Suppose (4) and let π:VW\uppi\colon V\to W be étale. We show that V(K)W(K)V(K)\to W(K) is τ\uptau-open. By Fact 8.8 we may suppose that π\uppi is basic standard étale. Let π\uppi, ff, and gg be as in Definition 8.7. Given αW(K)\alpha\in W(K) let fαK[X]f_{\alpha}\in K[X] be given by evaluating ff at α\alpha and let ι:W(K)Poln(K)\upiota\colon W(K)\to\mathrm{Pol}_{n}(K) be ι(α)=fα\upiota(\alpha)=f_{\alpha}. Note that ι\upiota is continuous with respect to τ\uptau. It is enough to fix (α,β)V(K)(\alpha,\beta)\in V(K) and a neighbourhood PW(K)×KP\subseteq W(K)\times K of (α,β)(\alpha,\beta) and show that π(V(K)P)\uppi(V(K)\cap P) is a neighbourhood of α\alpha. We may suppose that PP is contained in the open subvariety of W×𝔸1W\times\mathbb{A}^{1} given by g0g\neq 0. As the τ\uptau-topology on W(K)×KW(K)\times K is the product topology we suppose that P=O×UP=O^{*}\times U for a neighbourhood OW(K)O^{*}\subseteq W(K) of α\alpha and a neighbourhood UKU\subseteq K of β\beta. Note that β\beta is a simple root of fαf_{\alpha} as f/X\der f/\der X does not vanish at (α,β)(\alpha,\beta). Hence there is a neighbourhood OPoln(K)O\subseteq\mathrm{Pol}_{n}(K) such that every fOf^{*}\in O has a simple root in UU. We show that Oι1(O)O^{*}\cap\upiota^{-1}(O) is contained in π(V(K)P)\uppi(V(K)\cap P), note that Oι1(O)O^{*}\cap\upiota^{-1}(O) is a neighbourhood of α\alpha. Fix γOι1(O)\gamma\in O^{*}\cap\upiota^{-1}(O). Then fγOf_{\gamma}\in O, hence fγf_{\gamma} has a simple root η\eta in UU. We show that (γ,η)V(K)P(\gamma,\eta)\in V(K)\cap P. Note f(γ,η)=fγ(η)=0f(\gamma,\eta)=f_{\gamma}(\eta)=0. As γO\gamma\in O^{*} and ηU\eta\in U we have (γ,η)P(\gamma,\eta)\in P, so g(γ,η)0g(\gamma,\eta)\neq 0, hence (γ,η)V(K)(\gamma,\eta)\in V(K).

We show that (5) implies (4). Suppose (5) and fix n2n\geq 2 ((4) is trivial for n=1n=1). Let VV be the subvariety of SpecK[Y1,,Yn,X]=𝔸n×𝔸1\operatorname{Spec}K[Y_{1},\ldots,Y_{n},X]=\mathbb{A}^{n}\times\mathbb{A}^{1} given by Xn+Yn1Xn1++Y1X+Y0=0X^{n}+Y_{n-1}X^{n-1}+\ldots+Y_{1}X+Y_{0}=0 and (/X)[Xn+Yn1Xn1++Y1X+Y0]0(\der/\der X)[X^{n}+Y_{n-1}X^{n-1}+\ldots+Y_{1}X+Y_{0}]\neq 0. Let π:V𝔸n\uppi\colon V\to\mathbb{A}^{n} be the projection. Then π\uppi is standard étale, hence the projection V(K)KnV(K)\to K^{n} is open. Suppose a=(a0,,an1)Kna=(a_{0},\ldots,a_{n-1})\in K^{n}, f(X)=Xn+an1Xn1++a1X+a0f(X)=X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0}, bKb\in K is a simple root of ff, and PKP\subseteq K is a neighbourhood of bb. Note that (a,b)V(K)(a,b)\in V(K). Let O=π([Kn×P]V(K))O=\uppi([K^{n}\times P]\cap V(K)), so OO is a neighbourhood of aa. It is easy to see that f(X)=Xn+an1Xn1++a1X+a0f^{*}(X)=X^{n}+a^{*}_{n-1}X^{n-1}+\ldots+a^{*}_{1}X+a^{*}_{0} has a simple root in PP for any (a0,,an1)O(a^{*}_{0},\ldots,a^{*}_{n-1})\in O.

We show that (4) implies (1). Let PP be a neighbourhood of 1-1. Note that 1-1 is a simple root of Xn+1+XnX^{n+1}+X^{n}. Hence there is a neighbourhood OKnO\subseteq K^{n} of (1,0,,0)(1,0,\ldots,0) such that if a=(an,,a0)Oa=(a_{n},\ldots,a_{0})\in O then Xn+1+anXn++a1X+a0X^{n+1}+a_{n}X^{n}+\ldots+a_{1}X+a_{0} has a root in PP. Fix a neighbourhood QKQ\subseteq K of 0 such that {1}×QnO\{1\}\times Q^{n}\subseteq O. Then Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in PP for all a0,,an1Qa_{0},\ldots,a_{n-1}\in Q. Hence τ\uptau is gt-henselian.

We finish by showing that (3) implies (4). Suppose (3). Let fK[X]f\in K[X] be monic of degree nn, and let αK\alpha\in K be a simple root of ff. The change of variables Y=XαY=X-\alpha induces an automorphism of Poln(K)\mathrm{Pol}_{n}(K), so we may assume without loss of generality that α=0\alpha=0. Thus f=a1X++an1Xn1+Xnf=a_{1}X+\dotsb+a_{n-1}X^{n-1}+X^{n} with coefficients aiKa_{i}\in K, a1=f(0)0a_{1}=f^{\prime}(0)\neq 0.

Let PKP\subseteq K be a neighbourhood of 0. Let PPP^{\prime}\subseteq P be a smaller neighbourhood of 0 such that 1P-1\not\in P^{\prime} and a11P(1+P)1(1+P)Pa_{1}^{-1}\cdot P^{\prime}\cdot(1+P^{\prime})^{-1}(-1+P^{\prime})\subseteq P. By (3) there exists a neighbourhood OO of 0 such that every polynomial 1+X+c2X2++cnXn1+X+c_{2}X^{2}+\dotsb+c_{n}X^{n} with ciOc_{i}\in O has a root in 1+P-1+P^{\prime}. By shrinking OO and PP^{\prime}, we may assume that any root in 1+P-1+P^{\prime} of any such polynomial is simple.

Let OO^{\ast} be the set of polynomials b0+b1X++bn1Xn1+Xnb_{0}+b_{1}X+\dotsb+b_{n-1}X^{n-1}+X^{n} in Poln(K)\mathrm{Pol}_{n}(K) with b0Pb_{0}\in P^{\prime}, b1a1(1+P)b_{1}\in a_{1}(1+P^{\prime}), and bib0i1b1iOb_{i}b_{0}^{i-1}b_{1}^{-i}\in O for all i=2,,ni=2,\dotsc,n. This is a neighbourhood of ff in Poln(K)\mathrm{Pol}_{n}(K).

Let us show that every g=b0+b1X++bn1Xn1+XnOg=b_{0}+b_{1}X+\dotsb+b_{n-1}X^{n-1}+X^{n}\in O^{\ast} has a simple root in PP. If b0=0b_{0}=0, then 0 is a simple root of gg. Otherwise, consider the polynomial h=b01g(b0b11X)K[X]h=b_{0}^{-1}g(b_{0}b_{1}^{-1}X)\in K[X]. By construction, hh has the form 1+X+c2X2++cnXn1+X+c_{2}X^{2}+\dotsb+c_{n}X^{n} with ciOc_{i}\in O, and thus has a simple zero in 1+P-1+P^{\prime}. Therefore gg has a simple zero in b0b11(1+P)Pb_{0}b_{1}^{-1}(-1+P^{\prime})\subseteq P, as desired. ∎

Remark 8.9.

We have seen in Section 6 that the field p\mathbb{Q}_{p} carries 2202^{2^{\aleph_{0}}} many pairwise incomparable locally bounded gt-henselian topologies. This is in marked contrast to t-henselian topologies, where a field which is not separably closed can admit at most one such ([PZ78, Theorem 7.9], essentially F. K. Schmidt’s theorem on independent henselian valuations). Therefore, while it is sensible to speak of t-henselian fields and the t-henselian topology on one such (forbidding separably closed fields), we avoid the analogous terminology in the gt-henselian case.

The analysis of the topological field (K,τ)(K,\uptau) simplifies when τ\uptau is ω\omega-complete, i.e. it the collection of neighbourhoods of 0 is closed under countable intersections. Using an ultrapower argument, Prestel-Ziegler in [PZ78, Theorem 1.1] show that every (K,τ)(K,\uptau) may be replaced by some (K,τ)(K^{\ast},\uptau^{\ast}) which is “locally equivalent” to (K,τ)(K,\uptau) and such that τ\uptau^{\ast} is ω\omega-complete. Here local equivalence means that (K,τ)(K,\uptau) and (K,τ)(K^{\ast},\uptau^{\ast}) satisfy the same sentences in a certain logic extending first-order logic in the language of rings, allowing restricted second-order quantification over neighbourhoods of 0. See [PZ78, Section 1] for details on this formalism.

Lemma 8.10.

Let (K,τ)(K^{\ast},\uptau^{\ast}) be locally equivalent to (K,τ)(K,\uptau). Then τ\uptau^{\ast} is gt-henselian (t-henselian) if and only if τ\uptau is gt-henselian (t-henselian).

Proof.

It is immediate from the definition that gt-henselianity is expressed by a collection of local sentences. The same holds for t-henselianity (as is already expressed in [PZ78, Corollary 7.3]). ∎

For ω\omega-complete field topologies, we have the following.

Fact 8.11.

Suppose that τ\uptau is ω\omega-complete. Then τ\uptau is locally bounded if and only if τ\uptau is the SS-adic topology for a local subring SS of KK with K=Frac(S)K=\operatorname{Frac}(S). Furthermore τ\uptau is a V-topology if and only if τ\uptau is the SS-adic topology for a valuation subring SS of KK and τ\uptau is t-henselian if and only if τ\uptau is the SS-adic topology for a henselian valuation subring SS of KK.

We note that Fact 8.11 can fail without ω\omega-completeness. For instance, it fails for the usual topology on \mathbb{R} or \mathbb{C}.

Proof of Fact 8.11.

The first claim is in the proof of [PZ78, Theorem 2.2 (b)], the second is [PZ78, Lemma 3.3], and the third follows from [PZ78, Theorem 7.2]. ∎

A subset of KK is a henselian ideal if it is the maximal ideal of a henselian local subring of KK with fraction field KK. We say that τ\uptau is induced by a henselian local ring if τ\uptau is the RR-adic topology for a henselian local subring RR of KK with Frac(R)=K\operatorname{Frac}(R)=K.

Proposition 8.12.

Suppose τ\uptau is ω\omega-complete. The following are equivalent:

  1. (1)

    τ\uptau is gt-henselian.

  2. (2)

    τ\uptau admits a neighbourhood basis at zero consisting of henselian ideals.

If τ\uptau is also locally bounded then τ\uptau is gt-henselian if and only if τ\uptau is induced by a henselian local ring.

Hence an ω\omega-complete gt-henselian field topology is a union of henselian field topologies. For the proof of the proposition, we partly follow the proof of [PZ78, Theorem 2.2], see also [PZ78, Theorem 7.2].

Proof.

Let us first assume that (2) holds and show that this implies (1). Any τ\uptau-neighbourhood PKP\subseteq K of 1-1 contains a set 1+I-1+I, where II is a henselian ideal which is a τ\tau-neighbourhood of 0. Applying condition (3) from Fact 2.8, we see that O=IO=I satisfies the condition from Definition 8.1. Thus τ\uptau is gt-henselian.

For the converse direction, let us suppose that τ\uptau is gt-henselian. We wish to show that (2) holds. We fix a neighbourhood QQ of zero and construct an open henselian ideal PKP\subseteq K which is contained in QQ. We use Fact 2.8 to show that PP is a henselian ideal. Let KprK_{\mathrm{pr}} be the prime subfield of KK.

Claim.

Suppose that OO is a neighbourhood of 1-1. Then there is a neighbourhood PQP\subseteq Q of zero such that:

  1. (1)

    Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in OO when a0,,an1Pa_{0},\ldots,a_{n-1}\in P,

  2. (2)

    Kpr+PK_{\mathrm{pr}}+P is a local subring of KK with fraction field KK and maximal ideal PP.

Proof.

By gt-henselianity, for every n2n\geq 2 we may fix a neighbourhood UnU_{n} of 0 such that if a0,,an1Una_{0},\ldots,a_{n-1}\in U_{n} then Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in KK. By ω\omega-completeness there is a neighbourhood UU of zero such that UUnU\subseteq U_{n} for all nn. We may suppose that UQU\subseteq Q and that UU does not contain 11. Let r1,r2,r_{1},r_{2},\ldots be an enumeration of KprK_{\mathrm{pr}}. Construct a descending sequence (Pi:i)(P_{i}:i\in\mathbb{N}) of open neighbourhoods of zero such that P0=UP_{0}=U, and for all i1i\geq 1 the sets Pi+PiP_{i}+P_{i}, PiPiP_{i}-P_{i}, PiPiP_{i}\cdot P_{i}, and r1Pi,,riPir_{1}P_{i},\ldots,r_{i}P_{i} are all contained in Pi1P_{i-1} and (1+Pi)11+Pi1(1+P_{i})^{-1}\subseteq 1+P_{i-1}. By ω\omega-completeness P:=iPiP:=\bigcap_{i\in\mathbb{N}}P_{i} is a neighbourhood of zero. The proof of [PZ78, Theorem 2.2] shows that Kpr+PK_{\mathrm{pr}}+P is a local subring of KK with maximal ideal PP. Finally, Kpr+PK_{\mathrm{pr}}+P is open so Fact 2.6 shows that K=Frac(Kpr+P)K=\operatorname{Frac}(K_{\mathrm{pr}}+P). ∎

Inductively construct sequences (Pi:i)(P_{i}:i\in\mathbb{N}), (Oi:i)(O_{i}:i\in\mathbb{N}) of open neighbourhoods of 0, 1-1, respectively such that P0QP_{0}\subseteq Q, O0OO_{0}\subseteq O and for each ii\in\mathbb{N} and n1n\geq 1:

  1. (1)

    Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in OiO_{i} for any a0,,an1Pia_{0},\ldots,a_{n-1}\in P_{i},

  2. (2)

    Kpr+PiK_{\mathrm{pr}}+P_{i} is a local ring with maximal ideal PiP_{i} and fraction field KK,

  3. (3)

    PiOi1P_{i}\subseteq O_{i}-1 and Oi+1Pi+1O_{i+1}\subseteq P_{i}+1.

Let P:=iPiP:=\bigcap_{i\in\mathbb{N}}P_{i} and O:=iOiO:=\bigcap_{i\in\mathbb{N}}O_{i}. Note that O=P+1O=P+1. By ω\omega-completeness PP is a neighbourhood of 0. Let R=Kpr+PR=K_{\mathrm{pr}}+P. Since Kpr+PiK_{\mathrm{pr}}+P_{i} is a local ring with maximal ideal PiP_{i} for each ii, we easily check that RR is a ring and PP is an ideal with residue field KprK_{\mathrm{pr}}. Furthermore, every element of 1+P1+P is invertible in RR, since we have (1+Pi)11+PiR(1+P_{i})^{-1}\subseteq 1+P_{i}\subseteq R for every ii. It follows that RR is a local ring with maximal ideal PP. As PP is a neighbourhood of zero, RR is open. By Fact 2.6 K=Frac(Kpr+P)K=\operatorname{Frac}(K_{\mathrm{pr}}+P). Note that Xn+1+Xn+an1Xn1++a1X+a0X^{n+1}+X^{n}+a_{n-1}X^{n-1}+\ldots+a_{1}X+a_{0} has a root in P+1P+1 for every a0,,an1Pa_{0},\ldots,a_{n-1}\in P. Hence RR is henselian by Fact 2.8.

We now suppose that τ\uptau is locally bounded. We take QQ in the construction above to be bounded, hence PP is bounded. Then α+P\alpha+P is bounded for all αKpr\alpha\in K_{\mathrm{pr}}, so Kpr+PK_{\mathrm{pr}}+P is bounded as a countable union of bounded sets, since (K,τ)(K,\uptau) is ω\omega-complete.

An application of Fact 2.7 shows that τ\uptau is the RR-adic topology. ∎

Corollary 8.13.

Suppose that τ\uptau is locally bounded. Let (K,τ)(K^{*},\uptau^{*}) be locally equivalent to (K,τ)(K,\uptau) and ω\omega-complete. Then τ\uptau is gt-henselian if and only if τ\uptau^{*} is induced by a henselian local ring.

Proof.

Suppose τ\uptau is gt-henselian. By Lemma 8.10 τ\uptau^{*} is gt-henselian and by Proposition 8.12 τ\uptau^{*} is induced by a henselian local ring. Conversely, if τ\uptau^{*} is induced by a henselian local ring, then τ\uptau^{*} is gt-henselian, hence τ\uptau is gt-henselian by Lemma 8.10. ∎

8.2. When is the étale open topology induced by a locally bounded field topology?

Lemma 8.14.

If τ\uptau is gt-henselian then τ\uptau refines the étale open topology. If τ\uptau induces the étale open topology then τ\uptau is gt-henselian.

Proof.

Suppose τ\uptau is gt-henselian and VV is a KK-variety. By Proposition 8.6, any étale image in V(K)V(K) is τ\uptau-open, and so τ\uptau refines the K\mathscr{E}_{K}-topology on V(K)V(K). If τ\uptau induces the étale open topology then Fact 2.1(2) shows that if VWV\to W is an étale morphism of KK-varieties then V(K)W(K)V(K)\to W(K) is τ\uptau-open. Now again apply Proposition 8.6. ∎

Corollary 8.15 follows from Lemma 8.14 and the fact that KK is large if and only if the étale open topology on KK is not discrete [JTWY22, Theorem C].

Corollary 8.15.

If KK admits a gt-henselian field topology then KK is large.

See [Pop14] for a definition and an account of largeness. Corollary 8.15 generalizes the theorem that the fraction field of a henselian local domain is large [Pop10]. Corollary 8.15 was also observed in slightly more general form in [Pop14, Theorem 1.8].

Proposition 8.16.

Suppose that τ\uptau is locally bounded. Then the following are equivalent:

  1. (1)

    τ\uptau induces the étale open topology over KK.

  2. (2)

    τ\uptau is gt-henselian and some nonempty étale image in KK is bounded.

In particular the following are equivalent when RR is a local domain with fraction field KK:

  1. (3)

    The RR-adic topology agrees with the étale open topology.

  2. (4)

    The RR-adic topology is gt-henselian and RR contains a nonempty étale image.

Proof.

The second equivalence follows easily from the first and the definitions. We prove the first equivalence. Suppose (1). Then τ\uptau is gt-henselian by Lemma 8.14. As τ\uptau is locally bounded we may fix a bounded open UKU\subseteq K. Then UU contains a nonempty étale image, which is also bounded. Thus (2) holds. Now suppose (2). By Lemma 8.14 and Fact 2.3 it suffices to show that the K\mathscr{E}_{K}-topology on KK refines τ\uptau. Let f:V𝔸1f\colon V\to\mathbb{A}^{1} be an étale morphism of KK-varieties such that U=f(V(K))U=f(V(K)) is bounded. By Proposition 8.6 UU is τ\uptau-open, hence (αU+β:αK×,βK)(\alpha U+\beta:\alpha\in K^{\times},\beta\in K) is a basis for τ\uptau. Fact 2.1(3) shows that each αU+β\alpha U+\beta is K\mathscr{E}_{K}-open. ∎

Corollary 8.17.

Suppose that KK is perfect and τ\uptau is a locally bounded field topology on KK. Then the following are equivalent:

  1. (1)

    τ\uptau induces the étale open topology over KK.

  2. (2)

    τ\uptau is gt-henselian and f(V(K))f(V(K)) is bounded for some KK-variety morphism f:V𝔸1f\colon V\to\mathbb{A}^{1} with f(V(K))f(V(K)) infinite.

In particular the following are equivalent when RR is a local domain with fraction field KK:

  1. (1)

    The RR-adic topology agrees with the étale open topology.

  2. (2)

    The RR-adic topology is gt-henselian and f(V(K))Rf(V(K))\subseteq R for some KK-variety morphism f:V𝔸1f\colon V\to\mathbb{A}^{1} with f(V(K))f(V(K)) infinite.

Proof.

This follows from Proposition 8.16 and the fact that if KK is perfect and f:V𝔸1f\colon V\to\mathbb{A}^{1} is a KK-variety morphism then f(V(K))f(V(K)) is the union of a definable K\mathscr{E}_{K}-open set and a finite set, see [WY23, Theorem B]. ∎

Lemma 8.18.

Suppose KKK\equiv K^{\prime} and K\mathscr{E}_{K} is induced by some locally bounded field topology τ\uptau. Then K\mathscr{E}_{K^{\prime}} is induced by a locally bounded field topology τ\uptau^{\prime}. Furthermore, (K,τ)(K,\uptau) is locally equivalent to (K,τ)(K^{\prime},\uptau^{\prime}).

Proof.

We first argue that K\mathscr{E}_{K^{\prime}} on K=𝔸K1(K)K^{\prime}=\mathbb{A}^{1}_{K^{\prime}}(K^{\prime}) is a locally bounded field topology. For every d>0d>0, let

𝒰d:={\displaystyle\mathcal{U}_{d}:=\{ UK:U={xK:yK,f(x,y)=0,g(x,y)0,fY(x,y)0}\displaystyle U\subseteq K\colon U=\{x\in K\colon\exists y\in K,f(x,y)=0,g(x,y)\neq 0,\frac{\partial f}{\partial Y}(x,y)\neq 0\}
for some f,gK[X,Y] of total degree d}.\displaystyle\text{ for some $f,g\in K[X,Y]$ of total degree $\leq d$}\}.

Every U𝒰dU\in\mathcal{U}_{d} is an étale open subset of K=𝔸K1(K)K=\mathbb{A}^{1}_{K}(K), and by Fact 8.8 every étale open subset of KK is a union of elements of d𝒰d\bigcup_{d}\mathcal{U}_{d}, so that the étale open topology on K=𝔸K1(K)K=\mathbb{A}^{1}_{K}(K) has d𝒰d\bigcup_{d}\mathcal{U}_{d} as a basis. Let 𝒱d\mathcal{V}_{d} be the collection of V𝒰dV\in\mathcal{U}_{d} such that VV\neq\emptyset and every U𝒰dU\in\mathcal{U}_{d} is a union of scaled translates of VV, i.e. sets of the form aV+baV+b with aK×a\in K^{\times}, bKb\in K.

There exists some τ\uptau-bounded étale image 0UK0\in U\subseteq K, and by possibly shrinking UU we may assume that U𝒰dU\in\mathcal{U}_{d} for some dd which we now fix. The scaled translates of UU form a basis for the étale open topology on K=𝔸K1(K)K=\mathbb{A}^{1}_{K}(K), i.e. a basis of τ\uptau. In particular U𝒱dU\in\mathcal{V}_{d}, so 𝒱d\mathcal{V}_{d} is not empty. Note that by definition for any other element V𝒱dV\in\mathcal{V}_{d} we may write UU as a union of scaled translates of VV.

Let 𝒰d,𝒱d\mathcal{U}^{\prime}_{d},\mathcal{V}^{\prime}_{d} be the collection of subsets of KK^{\prime} defined analogously as 𝒰d\mathcal{U}_{d} and 𝒱d\mathcal{V}_{d} in KK. Since these are definable families and KKK^{\prime}\equiv K, there exists U𝒱dU^{\prime}\in\mathcal{V}^{\prime}_{d}. Moreover, we may assume the scaled translates of UU^{\prime} form a basis of a field topology τ\uptau^{\prime} on KK^{\prime}.

Since the families 𝒱d\mathcal{V}_{d} and 𝒱d\mathcal{V}^{\prime}_{d} of subsets of KK resp. KK^{\prime} are defined by the same (parameter-free) formula, and are bases of τ\uptau resp. τ\uptau^{\prime}, we have local equivalence of (K,τ)(K,\uptau) and (K,τ)(K^{\prime},\uptau^{\prime}). In particular, (K,τ)(K^{\prime},\uptau^{\prime}) is gt-henselian. By Proposition 8.16, τ\uptau^{\prime} induces the étale open topology on KK^{\prime}. ∎

We now want to analyse the situation of a locally bounded ω\omega-complete topology.

Theorem 8.19.

Suppose that τ\uptau is locally bounded and induces the K\mathscr{E}_{K}-topology. Let (K,τ)(K^{*},\uptau^{*}) be locally equivalent to (K,τ)(K,\uptau) and ω\omega-complete. Then τ\uptau^{*} is induced by a henselian local ring and τ\uptau^{*} induces the étale open topology over KK^{*}.

Proof.

By Lemma 8.18, there is a locally bounded field topology τ\uptau^{\prime} on KK^{\ast} inducing K\mathscr{E}_{K^{\ast}}. Since (K,τ)(K^{\ast},\uptau^{\ast}) is gt-henselian by local equivalence, τ\uptau^{\ast} refines τ\uptau^{\prime}. On the other hand, there exists an étale image in KK^{\prime} which is τ\uptau^{\ast}-bounded, since the same holds in (K,τ)(K,\uptau). Thus τ=τ\uptau^{\ast}=\uptau^{\prime} by [PZ78, Lemma 2.1(f)]. The statement now follows from Corollary 8.13. ∎

Theorem 8.19 reduces the question “When is the étale open topology induced by a locally bounded field topology?” to the question “When does the étale open topology agree with the RR-adic topology for a henselian local ring RR?”.

Proposition 8.20.

Suppose that KK is 1\aleph_{1}-saturated and suppose that the étale open topology over KK is induced by a locally bounded field topology on KK. Then there is a henselian local subring RR of KK such that the étale open topology over KK agrees with the RR-adic topology.

A field is 1\aleph_{1}-saturated if any descending sequence of nonempty definable sets has nonempty intersection. Such fields can for instance be produced using the ultrapower construction, see [CK90, Theorem 6.1.1].

Proof.

Let τ\uptau be the locally bounded field topology inducing the étale open topology over KK. By Theorem 8.19 it is enough to show that τ\uptau is ω\omega-complete. Fix a τ\uptau-bounded étale image UU in KK which contains 0. Then =(αU:αK×)\mathcal{B}=(\alpha U:\alpha\in K^{\times}) forms a neighbourhood basis for τ\uptau at zero consisting of definable sets, see [PZ78, Lemma 2.1 (e)]. By 1\aleph_{1}-saturation any intersection of countably many elements of \mathcal{B} contains an element of \mathcal{B}. Hence τ\uptau is ω\omega-complete. ∎

Theorem 1.6 from the introduction follows from the preceding proposition together with Lemma 8.18.

References

  • [BCR98] J. Bochnak, M. Coste, and M-F. Roy, Real algebraic geometry, Springer, 1998.
  • [Bou06] N. Bourbaki, Éléments de mathématique. Algèbre commutative. Chapitres 5 à 7, Springer, 2006, reprint of the 1975 original.
  • [CK90] C. C. Chang and H. J. Keisler, Model theory, third ed., Studies in Logic and the Foundations of Mathematics, vol. 73, North-Holland Publishing Co., Amsterdam, 1990. MR 1059055
  • [Cut04] Steven Dale Cutkosky, Resolution of singularities, Graduate Studies in Mathematics, vol. 63, American Mathematical Society, Providence, RI, 2004. MR 2058431
  • [DF21] Philip Dittmann and Arno Fehm, Nondefinability of rings of integers in most algebraic fields, Notre Dame Journal of Formal Logic 62 (2021), no. 3, 589–592.
  • [Duc09] Antoine Ducros, Les espaces de Berkovich sont excellents, Ann. Inst. Fourier (Grenoble) 59 (2009), no. 4, 1443–1552. MR 2566967
  • [EP05] Antonio J. Engler and Alexander Prestel, Valued fields, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2005. MR 2183496
  • [FJ05] Michael D. Fried and Moshe Jarden, Field arithmetic, Springer Berlin Heidelberg, 2005.
  • [FP11] Arno Fehm and Elad Paran, Galois theory over rings of arithmetic power series, Adv. Math. 226 (2011), no. 5, 4183–4197. MR 2770445
  • [Gab92] Ofer Gabber, KK-theory of Henselian local rings and Henselian pairs, Algebraic KK-theory, commutative algebra, and algebraic geometry (Santa Margherita Ligure, 1989), Contemp. Math., vol. 126, Amer. Math. Soc., Providence, RI, 1992, pp. 59–70. MR 1156502
  • [Gro65] A. Grothendieck, Éléments de géométrie algébrique. IV. Étude locale des schémas et des morphismes de schémas. II, Inst. Hautes Études Sci. Publ. Math. (1965), no. 24, 231. MR 199181
  • [Gro67] by same author, Éléments de géométrie algébrique. IV. Étude locale des schémas et des morphismes de schémas. IV, Inst. Hautes Études Sci. Publ. Math. (1967), no. 32, 361. MR 238860
  • [Hir64] Heisuke Hironaka, Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II, Ann. of Math. (2) 79 (1964), 109–203; ibid. (2) 79 (1964), 205–326. MR 0199184
  • [HRW04] William Heinzer, Christel Rotthaus, and Sylvia Wiegand, Catenary local rings with geometrically normal formal fibers, Algebra, arithmetic and geometry with applications (West Lafayette, IN, 2000), Springer, Berlin, 2004, pp. 497–510. MR 2037106
  • [ILO14] Luc Illusie, Yves Laszlo, and Fabrice Orgogozo (eds.), Travaux de Gabber sur l’uniformisation locale et la cohomologie étale des schémas quasi-excellents, Société Mathématique de France, Paris, 2014, Séminaire à l’École Polytechnique 2006–2008. [Seminar of the Polytechnic School 2006–2008], With the collaboration of Frédéric Déglise, Alban Moreau, Vincent Pilloni, Michel Raynaud, Joël Riou, Benoît Stroh, Michael Temkin and Weizhe Zheng, Astérisque No. 363-364 (2014) (2014). MR 3309086
  • [JL89] Christian U. Jensen and Helmut Lenzing, Model-theoretic algebra with particular emphasis on fields, rings, modules, Algebra, Logic and Applications, vol. 2, Gordon and Breach Science Publishers, New York, 1989. MR 1057608
  • [Joh20] Will Johnson, Dp-finite fields v: topological fields of finite weight, arXiv preprint arXiv:2004.14732 (2020).
  • [JTWY22] Will Johnson, Minh Chieu Tran, Erik Walsberg, and Jinhe Ye, Étale open topology and the stable field conjecture, arXiv:2009.02319, accepted in J. Eur. Math. Soc. (2022+).
  • [JWY21] Will Johnson, Erik Walsberg, and Jinhe Ye, The étale open topology over the fraction field of a henselian local domain, arXiv preprint arXiv:2108.01868, accepted in Math. Nachr. (2021).
  • [Kuh11] Franz-Viktor Kuhlmann, The defect, Commutative algebra—Noetherian and non-Noetherian perspectives, Springer, New York, 2011, pp. 277–318. MR 2762515
  • [Liu02] Qing Liu, Algebraic geometry and arithmetic curves, Oxford University Press, 2002.
  • [Mat80] Hideyuki Matsumura, Commutative algebra, 2 ed., Benjamin/Cummings, 1980.
  • [Nag75] Masayoshi Nagata, Local rings, Robert E. Krieger Publishing Co., Huntington, N.Y., 1975, Corrected reprint. MR 0460307
  • [Poo17] Bjorn Poonen, Rational points on varieties, Graduate Studies in Mathematics, vol. 186, American Mathematical Society, Providence, RI, 2017. MR 3729254
  • [Pop10] Florian Pop, Henselian implies large, Annals of Mathematics 172 (2010), no. 3, 2183–2195.
  • [Pop14] by same author, Little survey on large fields - old & new, Valuation Theory in Interaction, European Mathematical Society Publishing House, 2014, pp. 432–463.
  • [PZ78] Alexander Prestel and Martin Ziegler, Model theoretic methods in the theory of topological fields., Journal für die reine und angewandte Mathematik 0299_0300 (1978), 318–341.
  • [Rot97] Christel Rotthaus, Excellent rings, henselian rings, and the approximation property, Rocky Mountain Journal of Mathematics 27 (1997), no. 1.
  • [Sri19] Padmavathi Srinivasan, A virtually ample field that is not ample, Israel Journal of Mathematics 234 (2019), no. 2, 769–776.
  • [Sta20] The Stacks Project Authors, Stacks Project, https://stacks.math.columbia.edu, 2020.
  • [Tem13] Michael Temkin, Inseparable local uniformization, J. Algebra 373 (2013), 65–119. MR 2995017
  • [WY23] Erik Walsberg and Jinhe Ye, Éz fields, J. Algebra 614 (2023), 611–649. MR 4499357