TY - JOUR AU1 - Vezzani, Alberto AB - Abstract We construct the dagger realization functor for analytic motives over non-archimedean fields of mixed characteristic, as well as the Monsky–Washnitzer realization functor for algebraic motives over a discrete field of positive characteristic. In particular, the motivic language on the classic étale site provides a new direct definition of the overconvergent de Rham cohomology and rigid cohomology and shows that their finite dimensionality follows formally from one of Betti cohomology for smooth projective complex varieties. 1 Introduction The problem of the definition of a well-behaved cohomology theory “à la de Rham” for analytic varieties over a non-archimedean field lies on the pathological properties of the de Rham complex in this context. Even though it behaves as expected when applied to proper varieties, or analytifications of algebraic varieties, its (hyper)cohomology computed on very basic affinoid smooth rigid varieties (such as the closed disc $$\mathbb{B}^1$$) is oddly infinite dimensional. This is related to the impossibility to integrate general holomorphic rigid forms while preserving their radius of convergence. This classical problem has been studied and positively resolved by different authors (see [22], [25], [42]). The common strategy is to consider the (hyper)cohomology of an alteration of the de Rham complex, namely the overconvergent complex $$\Omega^{\dagger}$$ that is, the subcomplex of those forms which can be extended on a “strict neighborhood” of the variety. In order to give sense to this definition, one needs first to consider an affinoid variety and endow it with an overconvergent structure, which amounts to embedding it in the interior of a bigger one. Secondly, one has to prove than two different choices would induce two canonically equivalent cohomology groups, and that this definition can be extended functorially to arbitrary varieties. The technical tool which is behind these facts is the rigid version of Artin’s approximation theorem proved by Bosch [14] stating that any map of varieties can be approximated with a new one preserving two given overconvergent structures, combined with a homotopic argument. In this article, we follow the approach of Groxße-Klönne [23] and [25] and we give a motivic version of this procedure which can be stated as follows (see Theorem 4.23): Theorem. For any ring of coefficients $$\Lambda$$ the canonical functor $$l\colon \text{RigSm}^\dagger\!/K\rightarrow\text{RigSm}/K$$ from smooth rigid varieties with an overconvergent structure to rigid varieties induces a monoidal, triangulated equivalence of the associated categories of motives: Ll∗:RigDAe´t†eff(K,Λ)≅RigDAe´teff(K,Λ):Rl∗. □ We will actually prove a relative statement, where the base dagger variety is not necessarily the spectrum of the field $$K$$. We remark that, in particular, it is possible to define the spectrum of the overconvergent de Rham cohomology as the motive $$\mathbb{L} l^*\Omega^{\dagger}$$. The main technical ingredient for the proof is Proposition 3.10 which provides a way to approximate maps of rigid analytic varieties having an overconvergent structure, with maps that preserve such structures. It is reminiscent of (a cubical version of) Artin’s approximation lemma, but completely independent of it. We can apply the previous theorem to give a new definition of rigid cohomology, which is a good cohomological theory “à la de Rham” for algebraic varieties $$X$$ over a discrete field $$k$$ of positive characteristic. The idea, due to Monsky and Washnitzer [42] is to find (whenever possible) a smooth formal model $$\mathcal{X}$$ of $$X$$ over the ring of integers $$K^\circ$$ of a mixed-characteristic valued field $$K$$ with residue $$k$$, and then consider the overconvergent de Rham cohomology of the associated generic fiber $$\mathcal{X}_\eta$$. Also in this case, the major task which is solved in the literature consists in proving that this definition does not depend on the various choices made at each step, as well as in generalizing it to arbitrary varieties. Classically, the tools which are used include the convergent site of Ogus [46] and the crystalline site of Berthelot [11] for proper varieties, or the overconvergent site developed by Le Stum [37]. Using the motivic language, these problems are alternatively solved by the following remark: a theorem of Ayoub [2, Corollary 1.4.24] states that the special-fiber functor $$(\cdot)_\sigma\colon \text{FormSm}/ K^\circ\rightarrow\text{Sm}/k$$ from the category of smooth formal schemes over $$K^\circ$$ to the category of smooth varieties over the residue field $$k$$ induces an equivalence of motives. In particular, by letting $$(\cdot)_\eta\colon \text{FormSm}/ K^\circ\rightarrow\text{RigSm}/K$$ be the generic-fiber functor, there is a motivic triangulated monoidal functor: MW∗:DAe´teff(k,Λ)→∼R(⋅)σ∗FormDAe´teff(K∘,Λ)→L(⋅)η∗RigDAe´teff(K,Λ)→∼Rl∗RigDAe´t†eff(K,Λ). As a whole, by considering the functor $$\mathcal{MW}^*$$ and the complex $$\Omega^\dagger$$ we therefore obtain automatically a functorial cohomology theory on algebraic varieties over $$k$$ satisfying étale descent and homotopy invariance, which coincides with the one of Monsky–Washnitzer whenever this one is defined. It is formal to show that $$\mathcal{MW}^*$$ has a right adjoint $$\mathcal{MW}_*$$ and that the motive $$\mathcal{MW}_*\Omega^\dagger$$ represents the “classic” rigid cohomology, providing an alternative to its usual definition and to the rigid spectrum considered by Deglise–Mazzari [20] and Milne–Ramachandran [41] following Besser [13]. Our construction only uses canonical, explicit functors, the classic étale sites on algebraic and analytic varieties and no hypothesis on the valuation of $$K$$. Another crucial fact which is proved in the literature concerns the finite dimensionality of the cohomological theories mentioned above (the most general statements are in [35]), as well as their compatibility with base extensions. The classic proofs rely on several reduction procedures, involving resolutions of singularities, localizations and homotopy. They decompose the general statement into direct, computable checks on varieties of a special kind, such as the ones which are projective and smooth. We remark that these ad hoc constructions are encapsulated in a fundamental theorem of Ayoub [2, Theorem 2.5.35]. When combined with the results of [52], it states that the category of rigid analytic motives with rational coefficients over a base field $${\bf{RigDA}}^{\text{eff}}_{\acute{\mathrm{e}}\text{t}}(K,\mathbb{Q})$$ is generated, in a suitable sense, by the motives $$M(X)$$ associated with smooth projective algebraic varieties $$X$$ over $$K$$. Admittedly, the proof of the theorem consists in an elaborated composition of the standard reduction procedures, enhanced with the triangulated language allowed by the motivic setting. As shown above, the main outcome of this article is proving that the overconvergent de Rham cohomology and rigid cohomology factor over the triangulated category $${\bf{RigDA}}^{\text{eff}}_{\acute{\mathrm{e}}\text{t}}(K,\mathbb{Q})$$. In particular, using the theorem of Ayoub, we deduce (see Corollary 5.24) their finite dimensionality as well as their compatibility with base change by reducing to the motives $$M(X)$$ of the aforementioned form, and hence to well-known facts related to the classic de Rham cohomology of complex smooth projective varieties $$X(\mathbb{C})$$. Our proof makes no distinction between the discrete-valuation case and the general case, and is independent on the classic proofs (see [10], and partial results in [12], [24], and [40]). In the Appendix, we prove that an overconvergent structure of a variety corresponds to a presentation of its (adic) compactification as an inverse limit (in a weak sense defined by Huber) of strict inclusions of rigid varieties. This connects the theory of dagger spaces of Große-Klönne [23] to the theory of adic spaces of Huber [32] and strengthens the parallel between the techniques used in this paper and the ones of [53] where smooth perfectoid spaces arise as inverse limits of finite maps of rigid varieties. 2 Overconvergent Rigid Varieties From now on we fix a complete valued field $$K$$ endowed with a non-archimedean valuation of rank $$1$$ and residue characteristic $$p>0$$. We denote by $$\pi$$ a pseudo-uniformizer of $$K$$, that is, an invertible, topologically nilpotent element. We also denote by $$K^\circ$$ the ring of integers and by $$k$$ the residue field. We consider rigid analytic varieties as adic spaces, using the language of Huber [32]. In particular, when we consider a point $$x\in X$$ for a variety $$X$$ we mean a point in the sense of Huber (or, equivalently, a point of the $$G$$-topos of $$X$$). We only consider rigid analytic varieties over $$K$$ which are separated and taut (i.e., the closure of a quasi-compact subset is quasi-compact, see [33, Definition 5.1.2]). If $$R$$ is a Tate algebra, we sometimes denote by $$\text{Spa}\,R$$ the associated affinoid space $$\text{Spa}\,(R,R^\circ)$$. The starting point to define overconvergent, or dagger varieties are the so-called dagger algebras. For the sake of completeness, we report here their definition and some basic properties, proved in [23] and [42]. We refer to the Appendix for a link between these definitions and the language of adic spaces of Huber. Definition 2.1 ([23], [42]). For $$c\in K$$ and $$m,d\in\mathbb{N}_{>0}$$ we denote by $$K\langle c^{m/d}{\tau_1,\ldots,c^{m/d}\tau_n}\rangle$$ or simply by $$K\langle c^{m/d}\underline{\tau}\rangle$$ the subring of $$K\langle{\tau_1,\ldots,\tau_n}\rangle$$ of those power series $$\sum a_\alpha \underline{\tau}^\alpha$$ such that $$\lim |a_\alpha| \lambda^{|\alpha|}=0$$ for $$\lambda=|c|^{m/d}$$. We denote by $$K\langle\underline{\tau}\rangle^\dagger=K\langle \tau_1,\ldots,\tau_n\rangle^\dagger$$ the following topological subring of $$K\langle \tau_1\ldots,\tau_n\rangle$$ lim→h⁡K⟨π1/hτ_⟩=lim→h⁡K⟨π1/hτ1,…,π1/hτn⟩ that is, the ring of those power series $$\sum a_\alpha \underline{\tau}^\alpha$$ such that $$\lim |a_\alpha| \lambda^{|\alpha|}=0$$ for some $$\lambda\in\mathbb{R}_{>1}$$. A dagger algebra is a topological $$K$$-algebra $$R$$ isomorphic to a quotient $$K\langle\underline{\tau}\rangle^\dagger\!/I$$ of $$K\langle\underline{\tau}\rangle^\dagger$$. Its completion $$\hat{R}$$ is the Tate algebra $$K\langle\underline{\tau}\rangle/I$$. A morphism of dagger algebras $$R \rightarrow S$$ is a $$K$$-linear (hence continuous) morphism. The category $$\text{Aff}^\dagger$$ of affinoid dagger spaces is the opposite category of dagger algebras. We denote by $$\text{Spa}^\dagger R$$ the object on $$\text{Aff}^\dagger$$ associated with $$R$$. We say that its limit is $$\text{Spa}\,\hat{R}$$ where $$\hat{R}$$ is the completion of $$R$$ and vice-versa we say that $$\text{Spa}^\dagger R$$ is a dagger structure of $$\text{Spa}\,\hat{R}$$. We say that $$\text{Spa}^\dagger R$$ [resp. a morphism $$\text{Spa}^\dagger R'\rightarrow\text{Spa}^\dagger R$$] has the property $$\mathbf{P}$$ if $$\text{Spa}\,\hat{R}$$ [resp. the induced morphism of affinoid spaces $$\text{Spa}\,\hat{R}' \rightarrow\text{Spa}\,\hat{R}$$] has the property $$\mathbf{P}$$. The category of smooth morphisms of affinoid dagger spaces $$X \rightarrow S$$ to a fixed affinoid dagger space $$S=\text{Spa}^\dagger R$$ is denoted by $$\text{AffSm}^\dagger\!/S$$. Similarly, a collection of morphisms $$\{\text{Spa}^\dagger R_i \rightarrow\text{Spa}^\dagger R \}$$ in $$\text{AffSm}/S$$ is a cover if the induced collection $$\{\text{Spa}\,\hat{R}_i \rightarrow\text{Spa}\,\hat{R} \}$$ is a topological cover, that is, if $$\bigsqcup\text{Spa}\,\hat{R}_i \rightarrow\text{Spa}\,\hat{R}$$ is surjective. □ Remark 2.2. The functor $$\text{Spa}^\dagger R\mapsto\text{Spa}\,\hat{R}$$ is faithful, and $$\text{Hom}(\text{Spa}^\dagger R',\text{Spa}^\dagger R)$$ corresponds to those maps in $$\text{Hom}(\text{Spa}\,\hat{R}',\text{Spa}\,\hat{R})$$ such that the image of $$R\subset \hat{R}$$ lies in $$R'\subset\hat{R}'$$. □ Example 2.3. We denote by $$\mathbb{B}^{n\dagger}$$ the smooth affinoid dagger space $$\text{Spa}^\dagger K\langle\tau_1,\ldots,\tau_n\rangle^\dagger$$. □ Proposition 2.4 ([23, Paragraph 1.4]). The dagger algebra $$K\langle\underline{\tau}\rangle^\dagger$$ is a Noetherian factorial Jacobson ring. In particular, any dagger algebra is isomorphic to a quotient $$K\langle\tau_1,\ldots,\tau_n\rangle^\dagger\!/(f_1,\ldots,f_k)$$ with $$f_i\in K\langle \pi^{1/N}\underline{\tau}\rangle$$ for a sufficiently big $$N$$. □ Definition 2.5. Choose a presentation of a dagger algebra $$R \cong K\langle\tau_1,\ldots,\tau_n\rangle^\dagger\!/(f_1,\ldots,f_k)$$ with completion $$\hat{R}$$. We denote by $$\hat{R}_h$$ the Tate algebra $$K\langle\pi^{1/(H+h)}\underline{\tau}\rangle/(f_i)$$. It is well defined for all $$h\geq1$$ for a sufficiently big $$H$$. The ring $$R$$ is the union $$\varinjlim_h \hat{R}_h$$. If we denote by $$X$$ the space $$\text{Spa}^\dagger R$$ we also denote by $$X_h$$ the space $$\text{Spa}\,\hat{R}_h$$ and by $$\hat{X}$$ the space $$\text{Spa}\,\hat{R}$$. □ Remark 2.6. The definition of the algebras $$\hat{R}_h$$ above depend on the presentation of the dagger algebra $$R$$. Whenever we use this notation, we consider a possible choice of presentation of $$R$$. If we let $$R^{ +}$$ be $$\varinjlim \hat{R}_h^\circ$$ then the affinoid Huber space $$\text{Spa}\,(R ,R^{ +})$$ is the compactification of $$\hat{X}$$ over $$K$$ and is a (weak) inverse limit of adic spaces $$\varprojlim X_h$$ following Huber’s definition [33, Definition 2.4.2]. We refer to the Appendix for the details (see Proposition A.22). □ Remark 2.7. Let $$U$$ and $$V$$ be two open subvarieties of $$X$$. We write $$U\Subset_XV$$ if the closure of $$U$$ lies in $$V$$ (see [2, Proposition 2.1.13]). By [2, Proposition 2.1.16] we have that $$\hat{X}\Subset_{X_1} X_{h}\Subset_{X_1}X_1$$. Moreover, the sequence $$\{X_h\}$$ is coinitial with respect to $$\Subset$$ among rational subspaces of $$X_1$$ with this property. □ We now recall some basic facts about the category of dagger spaces. In particular, we isolate in the following proposition the fundamental Artin’s approximation lemma. It will not be used under this general form, but rather in a smooth “cubical” fashion (see 3.10). Proposition 2.8 ([22, Corollary 7.5.10]). Suppose $$\text{char} K=0$$ or $$\text{char} K=p>0$$ and $$[K\colon K^p]<\infty$$. Let $$X$$ and $$Y$$ be two affinoid dagger spaces with limit $$\hat{X}$$ and $$\hat{Y,}$$ respectively. We fix a Banach norm $$||\cdot||$$ on $$\mathcal{O}(\hat{X})$$ and $$\mathcal{O}(\hat{Y})$$. For any [iso-]morphism $$\phi\colon \hat{X}\rightarrow \hat{Y}$$ and any $$\varepsilon>0$$ there exists a [iso-]morphism $$\psi\colon X \rightarrow Y$$ such that $$||\mathcal{O}(\phi)(f)-\mathcal{O}(\hat{\psi})(f)||\leq\epsilon$$ for all $$f\in \mathcal{O}(\hat{Y})$$ with $$||f||\leq1$$. □ Remark 2.9. Following the notations of the previous proposition, the property $$||\mathcal{O}(\phi)(f)-\mathcal{O}(\hat{\psi})(f)||\leq\epsilon$$ for all $$f\in \mathcal{O}(\hat{Y})$$ with $$||f||\leq1$$ is typically denoted by $$||{\phi} - \hat{\psi}||\leq\epsilon$$. We will also follow this convention in what follows. □ Proposition 2.10 ([23, [Paragraph 1.16]). The category of dagger algebras has coproducts $$\otimes^\dagger$$. The categories $$\text{Aff}^\dagger$$ and $$\text{AffSm}/S$$ have fibered products. □ We recall how coproducts of dagger algebras are formed. Suppose given three dagger algebras $$T$$, $$R=K\langle\underline{\tau}\rangle^\dagger/I$$ and $$S=K\langle\underline{\sigma}\rangle^\dagger/J$$ and two maps $$T\rightarrow R$$, $$T\rightarrow S$$. The dagger algebra $$R\otimes^\dagger_TS$$ is the image of $$K\langle\underline{\tau},\underline{\sigma}\rangle^\dagger$$ under the canonical map to the Tate algebra $$\hat{R}\widehat{\otimes}_{\hat{T}}\hat{S}$$. Definition 2.11. Let $$R$$ be a dagger algebra. We denote by $$R \langle\tau_1\,\ldots,\tau_n\rangle^\dagger$$ the dagger algebra $$R \otimes^\dagger K\langle\tau_1\,\ldots,\tau_n\rangle^\dagger$$. □ Whenever $$f_1,\ldots,f_n,g$$ are elements of a Tate algebra $$R$$ generating the unit ideal, we denote by $$U(f_1,\ldots,f_n/g)$$ the rational space of the affinoid variety $$\text{Spa}\,R$$ defined by the conditions $$|f_i(x)|\leq|g(x)|$$. Proposition 2.12 ([23, Proposition 2.6, Paragraph 2.11]). Let $$X =\text{Spa}^\dagger R$$ be an affinoid dagger space. Any rational open subset $$U$$ of $$\text{Spa}\,\hat{R}$$ can be written as $$U(f_1,\ldots,f_n/g)$$ with $$f_i,g\in R$$ generating the unit ideal. The dagger space $$U =\text{Spa}^\dagger \mathcal{O}^\dagger(U)$$ with O†(U)=R⟨τ1,…,τn⟩†/(gτi−fi) is an open rational subspace of $$\text{Spa}^\dagger R$$ canonically independent on the choice of $$f_i,g$$. Moreover $$\mathcal{O}^\dagger$$ is a sheaf of topological $$K$$-algebras on $$\text{Spa}\,\hat{R}$$. □ Definition 2.13. Let $$X$$ be an affinoid dagger space with limit $$\hat{X}$$. We denote by $$\mathcal{O}^\dagger_{X }$$ or simply by $$\mathcal{O}^\dagger$$ the sheaf of topological algebras on the rational site of $$X$$ as well as the sheaf of topological algebras on $$\hat{X}$$ introduced in Proposition 2.12. □ In the category of affinoid rigid analytic spaces, the functor $$\text{Spa}\,R\mapsto R^\circ$$ is represented by $$\mathbb{B}^1$$. The next proposition shows the role of the dagger disc $$\mathbb{B}^{1\dagger}$$ introduced above. Proposition 2.14. The affinoid dagger space $$\mathbb{B}^{1\dagger}$$ represents the functor $$\text{Spa}^\dagger R \mapsto \hat{R}^\circ\cap R$$. □ Proof. Let $$X =\text{Spa}^\dagger R$$ be an affinoid dagger space. A continuous map $$K\langle\tau\rangle\rightarrow \hat{R}$$ amounts to the choice of an element $$s\in \hat{R}^\circ$$ and it preserves the dagger structures if and only if for any $$n$$ the induced map $$K\langle \pi^{1/n}\tau\rangle\rightarrow \hat{R}$$ factors over a map $$K\langle\pi^{1/n}\tau\rangle\rightarrow \hat{R}_h$$ for some $$h$$, that is, if and only if $$\pi s^n \in \varinjlim \hat{R}_h^\circ$$ for all $$n$$. Since $$R$$ is a $$K$$-algebra, integrally closed in $$\hat{R}$$ (see [15, Theorem 2]) we deduce that such an $$s$$ lies in $$\hat{R}^\circ\cap R$$. Vice-versa, we claim that any element $$s\in \hat{R}^\circ\cap R^\dagger$$ satisfies the condition. If $$f\in \hat{R}_1\cap R^\circ$$ then we deduce from [2, Proposition 2.1.16] the following inclusions in $$X_1=\text{Spa}\,\hat{R}_1$$: X^⊂U(f/1)⋐X1U(πf/1). Since $$\{X_h\}$$ is coinitial among the rational subvarieties $$W$$ of $$X_1$$ such that $$\hat{X }\Subset_{X_1}W$$ we deduce in particular that $$X_h\subset U(\pi f/1)$$ for some $$h$$ that is, $$\pi f\in\varinjlim \hat{R}_h^\circ$$. This proves the inclusions π(R∩R^∘)⊂lim→⁡R^h∘⊂R∩R^∘ and therefore our claim. ■ The following proposition already appears in [21, Theorem 2.3]. We present here an alternative proof based on the methods developed in the Appendix. Proposition 2.15. Let $$X$$ be an affinoid dagger space with limit $$\hat{X}$$. 1. The functor $$U \mapsto \hat{U}$$ defines an equivalence between the categories of inclusions of rational subspaces in $$X$$ and in $$\hat{X}$$. 2. The functor $$U\mapsto \hat{U}$$ defines an equivalence between the categories of finite étale affinoid spaces over $$X$$ and over $$\hat{X}$$. □ Proof. The first claim follows from Proposition 2.12. For the second claim, by [48, Lemma 7.5] we know that, up to shifting indices, any map $$\hat{U}\rightarrow \hat{V}$$ of finite étale affinoid spaces over $$\hat{X}$$ is induced by a map $$U_1\rightarrow V_1$$ of finite étale spaces over $$X_1$$ with $$\hat{U}=U_1\times_{X_1}\hat{X}$$ and $$V=V_1\times_{X_1}\hat{X}$$. Let $$U_h$$ be $$U_1\times_{X_1}X_h$$. We are left to prove that $$\varinjlim\mathcal{O}(U_h)$$ is a dagger algebra. We now use the equivalence between dagger affinoid spaces and their presentations, proved in the Appendix (see Proposition A.22). In particular, we can alternatively prove that the sequence $$U_h$$ is a presentation of $$\hat{U}$$. It suffices to show that $$\hat{U}$$ lies in $$\text{Int}(U_1)$$ (see the notations of Definition A.7). Since $$f\colon U_1\rightarrow X_1$$ is finite, by [9, Corollary 2.5.13(i)] and Corollary A.8 we deduce that $$\text{Int}(U_1/X_1)=U_1$$ and hence by Corollary A.11 we get $$\text{Int}(U_1)=f^{-1}\text{Int}(X_1)$$. We then need to prove that the image of $$\hat{U}$$ lies in the interior of $$X_1$$ and this is clear as it factors over $$\hat{X}$$ lying in $$\text{Int}( X_1)$$. ■ Remark 2.16. If $$Y\rightarrow X$$ is a finite étale morphism of affinoid dagger spaces, the dagger algebra $$\mathcal{O}^\dagger(Y )$$ associated with $$Y$$ is of the form $$\varinjlim(\hat{S}_1\otimes_{\hat{R}_1}\hat{R}_h)=\hat{S}_1\otimes_{\hat{R}_1}\mathcal{O}^\dagger(X )$$ for some finite étale algebra $$\hat{S}_1$$ over $$\hat{R}_1$$ (up to shifting indices). In particular, it is finite étale over $$\mathcal{O}^\dagger(X )$$. □ Corollary 2.17. Let $$X$$ be an affinoid dagger variety with limit $$\hat{X}$$. If $$\mathcal{V}$$ is an étale cover of $$\hat{X}$$, then it can be refined into another cover $$\hat{\mathcal{U}}$$ induced by an étale cover $$\mathcal{U}$$ of $$X$$. □ Proof. By Proposition 2.15 and the fact that any étale cover can be refined into a new one which is a composition of rational embeddings and finite étale maps, we can find a refinement $$\hat{\mathcal{U}}$$ of the cover which is the limit of a family $$\mathcal{U}$$ of maps of dagger spaces. This is also a cover of $$X$$ by definition. ■ Corollary 2.18. Let $$X$$ be an affinoid dagger variety with limit $$\hat{X}$$. The maps of the small rational and étale sites $$\hat{X}\rightarrow X$$ induces equivalences on the associated topoi. □ Proof. It suffices to use the criteria of [33, Appendix A]. ■ Definition 2.19. Let $$X =\text{Spa}^\dagger R$$ be an affinoid dagger space and $$M$$ be a finite $$R$$-module. We define $$\tilde{M}$$ to be the sheaf $$\tilde{M}=M\otimes_{R }\mathcal{O}^\dagger$$ on $$\hat{X}$$. A coherent $$\mathcal{O}^\dagger$$-module over a dagger space $$X$$ is a $$\mathcal{O}^\dagger$$-module in the category of sheaves of $$K$$-algebras over $$\hat{X}$$ isomorphic to $$\tilde{M}$$ for some finite $$R$$-module $$M$$. □ Remark 2.20. In [23, Theorem 2.16] it is proved that the notion of coherent sheaves over $$\text{Spa}^\dagger R$$ coincides with the notion of coherent modules over the ringed space $$(\text{Spa}\,\hat{R},\mathcal{O}^\dagger)$$ of [26, Chapter 0, Section 5.3]. □ Proposition 2.21. Let $$X$$ be an affinoid dagger space and $$\mathcal{F}$$ be a coherent $$\mathcal{O}^\dagger$$-module over it. 1. ([23, Proposition 3.1]) $$H^i_{\text{an}}(X ,\mathcal{F})=0$$ if $$i>0$$. 2. $$\mathcal{F}$$ extends to an étale sheaf over $$X$$ defined by putting $$f^*\mathcal{F}=f^{-1}\mathcal{F}\otimes_{f^{-1}\mathcal{O}^{\dagger}_X}\mathcal{O}^\dagger_Y$$ for each étale map $$Y \rightarrow X$$. 3. $$H^i_{\acute{\mathrm{e}}\text{t}}(X ,\mathcal{F})=0$$ if $$i>0$$. □ Proof. Suppose $$\mathcal{F}=\tilde{M}$$ and let $$R$$ be the dagger algebra $$\mathcal{O}^\dagger(X )$$. By means of Corollary 2.18, [22, Proposition 8.2.1] and the proof of [22, Proposition 8.2.3(2)] we are left to prove that or each surjective finite étale map $$\hat{Y}\rightarrow \hat{X}$$ the following sequence is exact 0→M→M⊗RS⇉M⊗R(S⊗R†S), where we denote by $$S$$ the dagger algebra associated with $$Y$$ (see Proposition 2.15). By Remark 2.16 the map $$R \rightarrow S$$ is finite étale, and in particular the dagger tensor product coincides with the usual one (see [23, Lemma 1.10]). The claim then follows from [27, Section I.3]. ■ We now recall the definition of Große-Klönne of “global” dagger spaces Definition 2.22. A dagger space is a pair $$X =(\hat{X},\mathcal{O}^\dagger)$$ where $$\hat{X}$$ is a rigid analytic space, and $$\mathcal{O}^\dagger$$ is a sheaf of topological $$K$$-algebras on $$\hat{X}$$ such that for some affinoid open cover $$\{\hat{U}_i\rightarrow \hat{X}\}$$ there are dagger structures $$U_i$$ on $$\hat{U}_i$$ with $$\mathcal{O}^\dagger|_{\hat{U}_i}\cong\mathcal{O}^\dagger_{U_i}$$ (see Proposition 2.12). A morphism of dagger spaces is a morphism of the underlying locally ringed spaces over $$K$$ (see [23, Definition 2.12]) and the category they form is denoted by $$\text{Rig}^\dagger$$. We say that the rigid space $$\hat{X}$$ is the limit of $$X$$ and vice-versa we say that $$X$$ is a dagger structure of $$\hat{X}$$. We say that $$X$$ [resp. a morphism $$X \rightarrow X'$$] has the property $$\mathbf{P}$$ if $$\hat{X}$$ [resp. the induced morphism of rigid spaces $$\hat{X}\rightarrow\hat{ X}'$$] has the property $$\mathbf{P}$$. Whenever $$S$$ is a dagger space, we denote by $$\text{Rig}^\dagger$$ the category of rigid spaces over it. The category of smooth morphisms of dagger spaces $$X \rightarrow S$$ to a fixed dagger space $$S$$ is denoted by $$\text{RigSm}^\dagger\!/S$$ and its full subcategory of affinoid objects by $$\text{AffSm}^\dagger\!/S$$. A collection of morphisms $$\{X_i \rightarrow X \}$$ in $$\text{RigSm}/S$$ is a cover if the induced collection $$\{\hat{X}_i\rightarrow \hat{X}\}$$ is a topological cover, that is, if $$\bigsqcup \hat{X}_i\rightarrow \hat{X}$$ is surjective. □ Example 2.23. Any rigid variety without boundary has a dagger structure by [23, Theorem 2.27]. We denote by $$\mathbb{P}^{1\dagger}$$ a dagger variety having as limit the projective line $$\mathbb{P}^1$$. □ The following proposition is straightforward. Proposition 2.24 Let $$S$$ be a dagger space. The functor Aff†/S →Rig†/SSpa†R ↦(Spa R^,O†) is fully faithful and induces an equivalence of the associated open analytic and the étale topoi. □ From now on, we will use the term affinoid dagger space also to indicate the objects in the essential image of the functor above. We easily obtain also the following version. Corollary 2.25. Let $$S$$ be a dagger space. The functor AffSm†/S →RigSm†/SSpa†R ↦(Spa R^,O†) is fully faithful and induces an equivalence of the associated open analytic and the étale topoi. □ 3 Approximation Results From now on, we fix a dagger space $$S$$ with limit $$\hat{S}$$. In this section, we recall the analytic version of the inverse function theorem and we use it as an alternative to Artin’s approximation theorem for smooth dagger algebras [14]. As a matter of fact, it induces a weaker form of this theorem (see Corollary 3.4) but also a cubical version of it that we will need in what follows (see Propositions 3.9 and 3.10). The reader who believes in Proposition 3.10 can safely skip this technical section. Proposition 3.1. Let $$R$$ be a dagger algebra with completion $$\hat{R}$$. If an element $$\xi$$ of $$\hat{R}$$ is algebraic over $$\text{Frac}\,R$$ then it lies in $$R$$. □ Proof. We denote by $$X$$ the space $$\text{Spa}^\dagger R$$. Since $$R$$ is algebraically closed in $${\hat{R}}$$ (see [15, Theorem 2]) we conclude that $$\text{Frac}\,R$$ is algebraically closed in $$( R \setminus\{0\})^{-1}\hat{R}$$ and therefore $$\xi\in\text{Frac}\,R$$. We can also assume $$\xi\in \text{Frac}\,\hat{R}_1$$ up to shifting indices. Let $$I_{\xi}=(d_1,\ldots,d_n)$$ be the ideal of denominators of $$\hat{R}_1$$ associated with the meromorphic function $$\xi$$ (see [22, Lemma 4.6.5]) and let $$V(I_{\xi})$$ be the induced (Zariski) closed subvariety of $$X_1=\text{Spa}\,\hat{R}_1$$. From now on, we denote by $$T^c$$ the closure of a subset $$T$$ in $$X_1$$. We recall that $$\xi$$ is analytic around a point $$x$$ if and only if $$(I_{\xi})_x=\mathcal{O}_{x}$$. Since $$\mathcal{O}_{x}$$ is local with maximal ideal equal to the support of the valuation at $$x$$ (see [32, Lemma 1.6(i)]) we deduce that $$\xi$$ is regular around $$x$$ if and only if $$x\notin V(I_\xi)$$ that is, if $$|d_i(x)|\neq0$$ for some $$i$$. By [33, Lemma 1.1.10] if $$x\in{\{y\}}^c$$ for some point $$y$$ then $$|d_i(x)|=0$$ if and only if $$|d_i(y)|=0$$. Using [31, Remark 2.1(iii)] and the regularity of $$\xi$$ on $$\hat{X}$$ we then deduce that $${\hat{X}}^c\cap V(I_\xi)=\emptyset$$. On the other hand, by [33, Lemma 1.5.10(1a)] this set $${\hat{X}}^c\cap V(I_\xi)$$ coincides with the intersection of the nested closed subsets $$\{{X}^c_h\cap V(I_{\xi})\}_h$$ inside the quasi-compact space $$X_1$$. From Cantor’s intersection theorem (see e.g., [44, Theorem 3-5.9]). we conclude $$X_h\cap V(I_\xi)=\emptyset$$ for $$h$$ large enough and hence $$\xi$$ is regular on $$X_h$$, as wanted. ■ We obtain in particular the following fact. Corollary 3.2. Let $$R$$ be a dagger algebra with completion $$\hat{R}$$. If $$f\in R$$ is invertible in $$\hat{R}$$ then it is invertible in $$R$$. □ We recall the following version of the inverse mapping theorem in the analytic context. Proposition 3.3 ([53, Corollary A.2]). Let $$\hat{R}$$ be a non-archimedean Banach $$K$$-algebra, let $$ {\sigma}=(\sigma_1,\ldots,\sigma_n)$$ and $$ {\tau}=(\tau_1,\ldots,\tau_m)$$ be two systems of coordinates, let $$\bar{\sigma}=(\bar{\sigma}_1,\ldots,\bar{\sigma}_n)$$ and $$ \bar{\tau}=(\bar{\tau}_1,\ldots,\bar{\tau}_m)$$ two sequences of elements of $$\hat{R}$$ and let $$ {P}=(P_1,\ldots,P_m)$$ be a collection of polynomials in $$\hat{R}[ {\sigma}, {\tau}]$$ such that $$ {P}( {\sigma}=\bar{\sigma}, {\tau}=\bar{\tau})=0$$ and $$\det(\frac{\partial P_i}{\partial \tau_j})( {\sigma}=\bar{\sigma}, {\tau}=\bar{\tau})\in \hat{R}^\times$$. There exists a unique collection $$ {F}=(F_1,\ldots,F_m)$$ of $$m$$ formal power series in $$\hat{R}[[ {\sigma-\bar{\sigma}}]]$$ such that $$ {F}( {\sigma}=\bar{\sigma})=\bar{\tau}$$ and $$ {P}( {\sigma}, {F}( {\sigma}))=0$$ in $$\hat{R}[[ {\sigma-\bar{\sigma}}]]$$ and they have a positive radius of convergence around $$\bar{\sigma}$$. □ Corollary 3.4. Suppose that $$S$$ is affinoid. Let $$X$$ and $$Y$$ be two affinoid dagger algebras smooth over $$S$$ such that $$\hat{Y}$$ is étale over a poly-disc $$\mathbb{B}^m\times \hat{S}$$. For any [iso-]morphism $$\phi\colon \hat{X}\rightarrow \hat{Y}$$ over $$\hat{S}$$ and any $$\varepsilon>0$$ there exists a [iso-]morphism $$\psi \colon X \rightarrow Y$$ over $$S$$ such that $$||\phi-\hat{\psi}||\leq\epsilon$$. □ Proof. We suppose $$S=\text{Spa}^\dagger A$$, $$X=\text{Spa}^\dagger R$$ and we denote their limits by $$\hat{S}=\text{Spa}\,\hat{A}$$, $$\hat{X}=\text{Spa}\,\hat{R}$$. Note that $$\hat{Y}$$ is isomorphic to $$\text{Spa}\,\hat{A}\langle\sigma_1,\ldots,\sigma_m,\tau_1,\ldots,\tau_n\rangle/(P_1,\ldots, P_n)$$ such that each $$P_i$$ is a polynomial in $$A[\sigma,\tau]$$ and $$\det\left(\frac{\partial P}{\partial \tau}\right)$$ is invertible in $$\mathcal{O}(\hat{Y})$$ by means of [2, Lemma 1.1.51]. We first assume that $$Y=\text{Spa}^\dagger A\langle \underline{\sigma},\underline{\tau}\rangle^\dagger\!/(\underline{P}(\underline{\sigma},\underline{\tau}))$$. We also remark that $$\det\left(\frac{\partial P}{\partial \tau}\right)$$ is invertible in $$\mathcal{O}^\dagger (Y)$$ by Corollary 3.2. The map $$\phi$$ is uniquely determined by the association $$({\sigma},{\tau})\mapsto(s,t)$$ from $$\hat{A}\langle \underline{\sigma},\underline{\tau}\rangle/(\underline{P})$$ to $$\hat{R}$$ for an $$m$$-tuple $$s$$ and an $$n$$-tuple $$t$$ in $$\hat{R}$$. By Corollary 3.3 there exists power series $$F=(F_{1},\ldots,F_{m})$$ in $$\hat{R}[[\sigma-\bar{\sigma}]]$$ such that (σ,τ)↦(s~,F(s~))∈R^ efines a new map $$\psi$$ from $$\hat{X}$$ to $$\hat{Y}$$ for any choice of $$\tilde{s}\in \hat{R}^\circ\cap R$$ such that $$\tilde{s}$$ is in the convergence radius of $$F$$ and $$F(\tilde{s})$$ is in $$\hat{R}^\circ$$. The field $$\text{Frac}\,(R )(F(\tilde{s}))$$ is finite étale over $$\text{Frac}\,(R )$$. By Proposition 3.1, we deduce that $$F(\tilde{s})$$ is in $$R$$ and therefore $$\psi$$ is a map from $$X$$ to $$Y$$. By the density of $$R$$ in $$\hat{R}$$ and the continuity of $$F$$ we can also assume that the $$m$$-tuple $$\tilde{s}$$ induces a map $$\psi$$ such that $$||\phi-\psi||\leq\epsilon$$. If we take $$\hat{X}=\hat{Y}$$ we can find an endomorphism $$\psi$$ of $$\hat{X}$$ inducing a map $$X \rightarrow Y$$ such that $$||\text{id}-\psi||\leq\epsilon$$. If we take $$\epsilon$$ sufficiently small, we deduce that $$\psi$$ is an automorphism of $$\hat{X}$$ and hence any two dagger structures on it are isomorphic. In particular, we obtain the general case of the proposition. ■ In the previous proof, we also showed the following structure theorem. Corollary 3.5. Suppose $$S$$ affinoid. Let $$\hat{Y}$$ be an affinoid rigid space which is étale over the poly-disc $$\mathbb{B}^{m}\times \hat{S}$$. Then it admits a dagger structure $$Y$$ isomorphic to $$\text{Spa}\,R$$ with R=O†(S^)⟨σ1,…,σm,τ1,…,τn⟩†/(P1,…,Pn), where each $$P_i$$ lies in $$\mathcal{O}^\dagger(\hat{S})[\underline{\sigma},\underline{\tau}]$$ and $$\det(\frac{\partial P_i}{\partial\tau_j})$$ is invertible in $$R$$. □ Proposition 3.6. Any rigid dagger space smooth over $$S$$ is locally étale over a dagger poly-disc $$\mathbb{B}^{n\dagger}\times{V }$$ for some affinoid rational open subset $$V\subset S$$. □ Proof. The claim is local on $$S$$ so we can assume it is affinoid. Let $$X$$ be a smooth dagger variety over it. By Proposition 2.15 and [2], we can find a rational covering $$\{U _i\rightarrow X \}$$ such that each limit $$\hat{U}_i$$ is étale over some poly-disc $$\mathbb{B}^n_{\hat{V}_i}$$ where $$\hat{V}_i$$ is rational inside $$\hat{S}$$. The claim then follows from Corollary 3.5. ■ We recall (see [16, Definition 1.1.9/1]) that a morphism of normed groups $$\phi\colon G\rightarrow H$$ is strict if the homomorphism $$G/\ker\phi\rightarrow\phi(G)$$ is a homeomorphism, where the former group is endowed with the quotient topology and the latter with the topology inherited from $$H$$. In particular, we say that a sequence of normed $$K$$-vector spaces R→fR′→gR″ is strict and exact at $$R'$$ if it exact at $$R'$$ and if $$f$$ is strict that is, the quotient norm and the norm induced by $$R'$$ on $$R/\ker(f)\cong\ker(g)$$ are equivalent. Lemma 3.7. For any map $$\sigma\colon T_\sigma\rightarrow\{0,1\}$$ defined on a subset $$T_\sigma$$ of $$\{1,\ldots,n\}$$ we denote by $$I_\sigma$$ the ideal generated by $$\theta_i-\sigma(i)$$ as $$i$$ varies in $$T_\sigma$$. For any finite set $$\Sigma$$ of such maps and any dagger algebra $$R$$ with limit $$\hat{R}$$ the following diagram of topological $$K$$-algebras has vertical inclusions and strict and exact lines Moreover, the ideal $$\bigcap_{\sigma\in\Sigma} I_\sigma$$ is generated by a finite set of polynomials with coefficients in $$\mathbb{Z}$$. □ Proof. The fact that the first line is strict and exact as well as the description of the generators of $$\bigcap I_\sigma$$ is proved in [53]. The same statement applied to the rings $$\hat{R}_h\langle\pi^{1/h}\underline{\theta}\rangle$$ and a direct limit argument show that also the second line is exact. We now prove that the vertical maps are inclusions. This can be proved only for the last two columns, where the statement is clear. As the second line is isometrically contained in the first, we also deduce that it is strict as well. ■ Let $$\sigma$$ and $$\sigma'$$ be maps defined from two subsets $$T_\sigma$$ resp. $$T_{\sigma'}$$ of $$\{1,\ldots,n\}$$ to $$\{0,1\}$$. We say that they are compatible if $$\sigma(i)=\sigma'(i)$$ for all $$i\in T_\sigma\cap T_{\sigma'}$$ and in this case we denote by $$(\sigma,\sigma')$$ the map from $$T_\sigma\cup T_{\sigma'}$$ extending them. Lemma 3.8. Let $$R$$ be a dagger algebra with completion $$\hat{R}$$ and $$\Sigma$$ a set as in Lemma 3.7. For any $$\sigma\in\Sigma$$ let $$\bar{f}_\sigma$$ be an element of $$\hat{R}\langle\underline{\theta}\rangle/I_\sigma$$ such that $$\bar{f}_\sigma|_{(\sigma,\sigma')}=\bar{f}_{\sigma'}|_{(\sigma,\sigma')}$$ for any couple $$\sigma,\sigma'\in\Sigma$$ of compatible maps. 1. There exists an element $$f\in \hat{R}\langle\underline{\theta}\rangle$$ such that $$f|_\sigma=\bar{f}_\sigma$$. 2. There exists a constant $$C=C(\Sigma)$$ such that if for some $$g\in \hat{R}\langle\underline{\theta}\rangle$$ one has $$|\bar{f}_\sigma-{g}|_\sigma|<\varepsilon$$ for all $$\sigma$$ then the element $$f$$ can be chosen so that $$|f-g|0$$ there exist elements $$\tilde{s}_1,\ldots,\tilde{s}_N$$ of $$\hat{R}\langle\underline{\theta}\rangle^\circ\cap R\langle\underline{\theta}\rangle^\dagger$$ satisfying the following conditions. 1. $$|s_\alpha-\tilde{s}_\alpha|<\varepsilon$$ for each $$\alpha$$. 2. For any $$\alpha,\beta\in\{1,\ldots,N\}$$ and any $$k\in\{1,\ldots,n\}$$ such that $$s_\alpha|_{\theta_k=0}=s_\beta|_{\theta_k=0}$$ we also have $$\tilde{s}_\alpha|_{\theta_k=0}=\tilde{s}_\beta|_{\theta_k=0}$$. 3. For any $$\alpha,\beta\in\{1,\ldots,N\}$$ and any $$k\in\{1,\ldots,n\}$$ such that $$s_\alpha|_{\theta_k=1}=s_\beta|_{\theta_k=1}$$ we also have $$\tilde{s}_\alpha|_{\theta_k=1}=\tilde{s}_\beta|_{\theta_k=1}$$. 4. For any $$\alpha\in\{1,\ldots,N\}$$ if $$s_\alpha|_{\theta_1=1}\in R\langle\underline{\theta}\rangle^\dagger$$ then $$\tilde{s}_\alpha|_{\theta_1=1}={s}_\alpha|_{\theta_1=1}$$. □ Proof. We will actually prove a stronger statement, namely that we can reinforce the previous conditions with the following: 5. For any $$\alpha,\beta\in\{1,\ldots,N\}$$ any subset $$T$$ of $$\{1,\ldots,n\}$$ and any map $$\sigma\colon T\rightarrow\{0,1\}$$ such that $$s_\alpha|_{\sigma}=s_\beta|_{\sigma}$$ then $$\tilde{s}_\alpha|_{\sigma}=\tilde{s}_\beta|_{\sigma}$$. 6. For any $$\alpha\in\{1,\ldots,N\}$$ any subset $$T$$ of $$\{1,\ldots,n\}$$ containing $$1$$ and any map $$\sigma\colon T\rightarrow\{0,1\}$$ such that $$s_\alpha|_{\sigma}\in R\langle\underline{\theta}\rangle^\dagger$$ then $$\tilde{s}_\alpha|_{\sigma}={s}_\alpha|_{\sigma}$$. Above we denote by $$s|_\sigma$$ the image of $$s$$ via the substitution $$(\theta_{t}=\sigma(t))_{t\in T}$$. We proceed by induction on $$N$$, the case $$N=0$$ being trivial. We remark that if $$\epsilon$$ is sufficiently small, any element $$a$$ such that $$|a-s_k|<\varepsilon$$ lie in $$\hat{R}\langle\underline{\theta}\rangle^\circ$$ as this ring is open. We are left to prove that we can pick elements $$\tilde{s}_k$$ in $$R\langle\underline{\theta}\rangle^\dagger$$. Consider the conditions we want to preserve that involve the index $$N$$. They are of the form si|σ=sN|σ and are indexed by some pairs $$(\sigma, i)$$ where $$i$$ is an index and $$\sigma$$ varies in a set of maps $$\Sigma$$. Our procedure consists in determining by induction the elements $$\tilde{s}_1,\ldots,\tilde{s}_{N-1}$$ first, and then deduce the existence of $$\tilde{s}_N$$ by means of Lemma 3.8 by lifting the elements $$\{\tilde{s}_i|_{\sigma}\}_{(\sigma,i)}$$. Therefore, we first define $$\varepsilon'\colon = \frac{1}{C}\varepsilon$$ where $$C=C(\Sigma)$$ is the constant introduced in Lemma 3.8 and then apply the induction hypothesis to the first $$N-1$$ elements with respect to~$$\varepsilon'$$. By the induction hypothesis, the elements $$\tilde{s}_i|_\sigma$$ satisfy the compatibility condition of Lemma 3.8 and lie in $$R\langle\underline{\theta}\rangle^\dagger$$. By Lemma 3.8 we can find an element $$\tilde{s}_N$$ of $$R\langle\underline{\theta}\rangle^\dagger$$ lifting them such that $$|\tilde{s}_N-s_N|0}\times\delta^\mathbb{Z}$$ induced by putting $$1<\delta<\mathbb{R}_{>1}$$ and let $$||\cdot||$$ be the Gauss norm on $$K[\tau]$$. Consider the valuation ring $$K(\tau)^+$$ in $$K(\tau)$$ induced by the following valuation, taking values in $$\mathbb{R}_{>0}\times\delta^\mathbb{Z}\cup\{0\}$$: |⋅|∞:f=∑aiτi↦||f||⋅δmax{i:|ai|=||f||}. The associated valuation of rank $$1$$ is the Gauss norm on $$K(\tau)$$ which defines a point in $$\mathbb{B}^1$$. Nonetheless, there is no lift of $$\text{Spa}\,(K(\tau),K(\tau)^+)\rightarrow\text{Spa}\,K$$ to $$\mathbb{B}^1$$ since $$|\tau|_{\infty}=\delta>1$$. □ Definition A.3. ([33, Theorem 5.1.5]). The universal compactification of a map of rigid analytic varieties $$f\colon X\rightarrow S$$ is a factorization $$X\stackrel{j}{\rightarrow} X^{\text{cp}}_f\stackrel{f'}{\rightarrow} S$$ of adic spaces such that $$j$$ is locally closed, $$f'$$ is partially proper and such that for any other factorization $$X\stackrel{h}{\rightarrow} Y\stackrel{g}{\rightarrow} S$$ with $$g$$ partially proper, there exists a unique map $$i\colon X^{\text{cp}}_f\rightarrow Y$$ making the following diagram commute: The universal compactification of $$X\rightarrow\text{Spa}\,K$$ will be simply called the compactification of $$X$$ and denoted by $$X^{\text{cp}}$$. □ Example A.4. The universal compactification of a map of affinoid rigid analytic varieties $$\text{Spa}\,S\rightarrow\text{Spa}\,R$$ induced by a map $$\phi\colon R\rightarrow S$$ is given by the affinoid (yet not rigid analytic!) space $$\text{Spa}\,(S,S^+)$$ where $$S^+$$ is the integral closure in $$S$$ of the ring $$\phi(R^\circ)+S^{\circ\circ}$$. In particular, the compactification of $$\text{Spa}\,R$$ is the space $$\text{Spa}\,(R,R^+)$$ where $$R^+$$ is the minimal choice among rings of integral elements in $$R$$ over $$K$$, namely the integral closure of $$K^\circ+R^{\circ\circ}$$ in $$R$$. It contains $$\text{Spa}\,R$$ as an open dense subset. □ Definition A.5. Let $$X\subset Y$$ be an open immersion of rigid analytic varieties over a variety $$S$$. We write $$X\Subset_S Y$$ if the inclusion factors over the adic compactification of $$X$$ over $$S$$ (see [33, Theorem 5.1.5]). In case $$S=\text{Spa}\,K$$ we simply write $$X\Subset Y$$. □ Remark A.6. Let $$Y\rightarrow X$$ be an open immersion of rigid varieties over $$S$$. Then $$Y\Subset_S X$$ if and only if the compactification of $$Y$$over $$S$$ coincides with the compactification of $$Y$$ over $$X$$. □ Definition A.7. Let $$f\colon X\rightarrow S$$ be a morphism of rigid analytic varieties over a field $$K$$. The interior$$\text{Int}(X/S)$$ [resp. the border$$\partial(X/S)$$] of $$f$$ is [the complementary of] the union of its open supsbaces $$U$$ such that $$U\Subset_S X$$. If $$S=\text{Spa}\,K$$ we simply write $$\text{Int}(X)$$ [resp. $$\partial(X)$$]. □ We recall that the Berkovich space $$X^{\mathbb{B}erk}$$ associated with a rigid analytic (taut) variety $$X$$ is the universal Hausdorff quotient of the topological space underlying $$X$$ (see [48, Theorem 2.24]). In particular, there is a continuous quotient morphism $$\mathbb{B}erk\colon X\rightarrow X^{\mathbb{B}erk}$$. Our notations coincide with the one of [2] by means of the following interpretation in terms of Tate and Berkovich spaces. Proposition A.8. Let $$X\subset Y$$ be an open immersion of rigid analytic varieties over a variety $$S$$. Then $$X\Subset_SY$$ if and only if $$X^{\mathbb{B}erk}$$ lies in the Berkovich interior of $$Y^{\mathbb{B}erk}$$ over $$S^{\mathbb{B}erk}$$. □ Proof. By [9, Proposition 2.5.17], we can assume that all varieties are affinoid $$X=\text{Spa}\,(B,B^\circ)$$, $$Y=\text{Spa}\,(A,A^\circ)$$, $$S=\text{Spa}\,(C,C^\circ)$$. By [9, Proposition 2.5.2(d) and Proposition 2.5.9], $$X$$ lies in $$\text{Int}(Y/S)$$ if and only if the image of $$A^\circ$$ in $$B^\circ/B^{\circ\circ}$$ is integral over the image of $$C^\circ$$. This amounts to say that $$A^\circ$$ is mapped in the integral closure $$B^+$$ of $$C^\circ+B^{\circ\circ}$$ in $$B$$ which is precisely the ring of integers of the compactification $$\text{Spa}\,(B,B^+)$$ of $$X$$ over $$S$$. ■ We then immediately obtain the following result. Corollary A.9. Let $$f\colon X\rightarrow S$$ be a morphism of rigid analytic varieties over a field $$K$$. The space $$\text{Int}(X/S)$$ is the inverse image via $$X\rightarrow X^{\mathbb{B}erk}$$ of the Berkovich interior of $$X^{\mathbb{B}erk}$$ over $$S^{\mathbb{B}erk}$$. □ We also recall the following fundamental formula of Berkovich. Definition A.10. A rigid analytic variety $$X$$ is good if for any point $$x\in X$$ there exists an open subaffinoid of $$X$$ containing the closure of $$\{x\}$$ in $$X$$ (see [33, Proposition 8.3.7]). □ Corollary A.11 ([9, Proposition 2.5.8(iii)]). Let $$f\colon X\rightarrow Y$$ and $$g\colon Y\rightarrow S$$ be two morphisms of good rigid analytic varieties over a field $$K$$. It holds Int(X/S)=Int(X/Y)∩f−1Int(Y/S). □ Remark A.12. We recall that a morphism $$X\rightarrow Y$$ of rigid analytic spaces is partially proper if its border is empty, and it is proper if it is partially proper and quasi-compact. Thanks to the properties above, these definitions coincide with Berkovich’s and with Huber’s. The results of [50] and [51] also show that the notion of properness coincides with Kiehl’s (see also [33, Remark 1.3.19] for the discrete-valuation case). □ Proposition A.13 ([2, Proposition 2.1.16]). Let $$X=U(f_i/g)$$ be a rational subvariety of the affinoid space $$X_1$$. The sequence $$X_h=U(\pi f_i^h/g^h)$$ of rational subspaces of $$X_1$$ satisfies $$X\Subset_{X_1}X_{h+1}\Subset_{X_1}X_h$$ and is coinitial with respect to $$\Subset_{X_1}$$ among rational subspaces greater than $$X$$. □ We remark that the category of adic spaces doesn’t have all inverse limits. Nonetheless, there is a notion of being similar to the inverse limit for an object $$X$$ having maps towards a directed system $$\{X_{i+1}{\rightarrow} X_{i}\}$$ given by Huber (see [33, Section 2.4]) and denoted by $$X\sim\varprojlim X_h$$. In this case, the étale topos of $$X$$ is the filtered limit topos of those associated with $$X_i$$ ([33, Proposition 2.4.4]). Proposition A.14. Let $$X=\text{Spa}\,\hat{R}$$ be a rational subspace of $$X_1=\text{Spa}\,\hat{R}_1$$ and let $$X_h=\text{Spa}\,\hat{R}_h$$ be a sequence of rational affinoids in $$X_1$$ totally ordered with respect to $$\Subset_{X_1}$$ and coinitial among the opens $$W$$ with $$X\Subset_{X_1}W$$. Let $$R$$ [resp. $$R^{+}$$] be $$\varinjlim \hat{R}_h$$ [resp. $$\varinjlim \hat{R}_h^\circ$$]. If we endow $$R$$ with the topology induced by $$\hat{R}$$ then $$\text{Spa}\,(R,R^{+})$$ is an affinoid adic space with global sections equal to $$\hat{R}$$ and Spa(R,R+)∼lim←h⁡Spa R^h. Moreover if $$\text{Spa}\,(S,S^+)$$ is an affinoid adic space over $$K$$ with $$S^+$$ bounded is $$S$$ then Hom(Spa(S,S+),Spa(R,R+))≅lim←h⁡Hom(Spa(S,S+),Spa R^h). □ Proof. By Proposition A.13 we can assume that $$X=U(f_i/g)$$ and $$X_h=U(\pi(f_i^h/g^h))$$ as subspaces of $$X_1$$. We first observe that $$R$$ is dense in $$\hat{R}$$. Indeed, elements in $$\hat{R}_1\langle \upsilon\rangle/(g\upsilon-f)$$ having a polynomial as representative are dense, so that in particular the image of $$\hat{R}_1\langle \pi\upsilon\rangle$$ which is included in $$R$$ is also dense. We now claim that the ring $$R^{+}$$ is open and bounded in $$R$$. Since $$\hat{R}$$ is reduced, then $$\hat{R}^\circ$$ is bounded and the topology on $$R$$ is induced by the ring of definition $$R\cap \hat{R}^\circ$$. We already proved in Proposition 2.14 the chain of inclusions π(R∩R^∘)⊂R+⊂R∩R^∘ and therefore our claim. Since $$R^{+}$$ is open and integrally closed in $$R$$ the pair $$(R,R^{+})$$ is an affinoid pair. By what we proved above and [22, Lemma 7.5.3] its completion coincides with the pair $$(\hat{R},\hat{R}^+)$$ where $$\hat{R}^+$$ is the $$\pi$$-adic completion of $$R^{+}$$. As $$\text{Spa}\,\hat{R}$$ is adic (i.e., the structure presheaf is a sheaf) then also $$\text{Spa}\,(\hat{R},\hat{R}^+)=\text{Spa}\,(R,R^{+})$$ is adic (see [32, Theorem 2.2]). Maps of affinoid spaces $$\text{Spa}\,(T,T^+)\rightarrow\text{Spa}\,(S,S^+)$$ over $$K$$ for which $$S^+$$ and $$T^+$$ are bounded are uniquely determined by the maps of abstract $$K^\circ$$-algebras $$S^+\rightarrow T^+$$. The last isomorphism follows then from our definitions and [32, Proposition 2.1(i)]. If we apply it to spectra of valued fields $$\text{Spa}\,(L,L^+)$$ we deduce in particular $$|\text{Spa}\,(R,R^{+})|\cong\varprojlim_h|\text{Spa}\,\hat{R}_h|$$ and therefore $$\text{Spa}\,(R,R^{+})\sim\varprojlim_h\text{Spa}\,\hat{R}_h$$. ■ We warn the reader that the completion of $$R^{+}$$ may vary among the rings of integral elements of $$R$$, as the next examples show. Example A.15. If we take $$X=X_h=X_1$$ we obtain $$\text{Spa}\,(R,R^{+})=\text{Spa}\,(\hat{R},\hat{R}^\circ)$$. □ Example A.16. Suppose that $$X_2\Subset X_1$$. By Proposition A.11 we deduce $$X_{h+1}\Subset X_h$$ and therefore $$\mathcal{O}^\circ(X_h)\subset\overline{K^\circ+\mathcal{O}^{\circ\circ}(X_{h+1})}$$ so that $$R^+$$ is contained in the algebraic closure of $$K^\circ+\hat{R}^{\circ\circ}$$ in $$\hat{R}$$. We conclude that $$\text{Spa}\,(R,R^+)$$ is the compactification of $$X$$ over $$K$$. □ We now make specific examples of this last situation. Example A.17. Consider the rational inclusion $$X=\mathbb{B}^n=\text{Spa}\,K\langle\underline{\tau}\rangle\rightarrow\text{Spa}\,K\langle\pi\underline{\tau}\rangle=X_1\cong\mathbb{B}^n$$ and let $$X_h$$ be the rational space $$U(\pi^{1/h}\underline{\tau})\colon= U\left((\pi\underline{\tau})^{h}/\pi^{h-1}\right)$$ of $$X_1$$. The sequence $$X_{h+1}\Subset_{X_1}X_h$$ is coinitial among opens $$W$$ such that $$X\Subset_{X_1}W$$ and $$X_{h+1}\Subset X_h$$. Moreover, $$R$$ coincides with $$K\langle\underline{\tau}\rangle^\dagger$$. □ We remark that Proposition A.14 applied to the example A.17 generalizes the claim at the end of [54, Example 7.58]. This last example can be extended to the following situation. Example A.18. Consider a rational inclusion $$X=X_1(f_1,\ldots,f_m/g)\Subset X_1$$ of affinoid rigid spaces. By the proofs of [9, Proposition 2.5.2, Proposition 2.5.9] we can suppose that there are presentations O(X1)=K⟨ρ1−1τ1,…,ρn−nτn⟩/IO(X)=K⟨π−1ρ1−1τ1,…,π−1ρn−1τn,υ1,…,υm⟩/((υig−fi)+I) with $$\rho_i\in\sqrt{K^\times}$$. We then define $$X_h$$ to be the rational subset of $$X_1$$ with O(Xh)=K⟨π1h−1ρ1−1τ1,…,π1h−1ρn−1τn,π1hυ1,…,π1hυm⟩/((υig−fi)+I). The sequence $$X_{h+1}\Subset X_h$$ is coinitial among opens $$W$$ such that $$X\Subset_{X_1}W$$. We also obtain R=K⟨(πρ1)−1τ1,…,(πρn)−1τn,υ1,…,υm⟩†/((υig−fi)+I) which is a dagger algebra. □ We are then inclined to make the following definition. Definition A.19. Fix an affinoid rigid space $$X$$. A presentation of a dagger structure on $$X$$ is a pro-affinoid variety $$\varprojlim X_h$$ where $$X$$ and all $$X_h$$ are rational subspaces of $$X_1$$, such that $$X\Subset X_{h+1}\Subset X_h$$ and the system is coinitial among rational subsets of $$X_1$$ containing $$X$$ in their interior. A morphism of presentations between $$\varprojlim X_h$$ and $$\varprojlim Y_k$$ is a morphism of pro-objects, that is, an element of $$\varprojlim_k\varinjlim_h\text{Hom}(X_h,Y_k)$$. □ Example A.20. The dagger poly-disc $$\mathbb{B}^{n\dagger}$$ has the presentation $$\varprojlim X_h$$ described in Example A.17. □ Remark A.21. The system $$X_h$$ of A.15 is not a presentation of a dagger structure on $$X$$ since $$X_{h+1}$$ is not contained in the interior of$$X_h$$. □ We summarize the previous discussion in the following proposition, drawing the link between the definition of dagger algebras and the language of (weak) inverse limits of adic spaces due to Huber. Proposition A.22. Let $$\hat{X}=\text{Spa}\,(\hat{R},\hat{R}^\circ)$$, be an affinoid space and let $$\varprojlim {X}_h$$ be a presentation of a dagger structure on $$\hat{X}$$. 1. $$\varinjlim \mathcal{O}( {X}_h)$$ is a dagger algebra $$R$$ dense in $$\hat{R}$$. 2. The functor $$\varprojlim {X}_h\mapsto \text{Spa}^\dagger R$$ induces an equivalence of categories between dagger affinoid spaces $$ \text{Aff}^\dagger$$ as introduced in Definition 2.1 and their presentations. 3. $$\hat{X}^{\text{cp}}=\text{Spa}\,(\hat{R},\hat{R}^+)\cong\text{Spa}\,(R,R^{+})\sim\varprojlim \hat{X}_h$$ where $$\hat{R}^+$$ is the integral closure of $${K^\circ+\hat{R}^{\circ\circ}}$$ in $$\hat{R}$$ and $$R^+$$ is $$\varinjlim\hat{R}_h^\circ$$. 4. If $$\text{Spa}\,(T,T^+)$$ is an affinoid adic space with $$T^+$$ bounded then Hom(Spa(T,T+),Spa(R,R+))≅lim←h⁡Hom(Spa(T,T+),Spa R^h). □ Proof. The first claim follows from Example A.18. The fully faithfulness of the functor in the second claim follows from Remark 2.14. Its essential surjectivity is immediate: fix a dagger algebra $$R$$ and an integer $$H$$ such that $$R=K\langle\underline{\tau}\rangle^\dagger/(a_i)$$ with $$a_i\in K\langle \pi^{1/H}\underline{\tau}\rangle$$. Then $$R$$ is the image of the sequence $$X_{h+1}\Subset X_h$$ with $$X_h=\text{Spa}\,K\langle \pi^{1/(H+h)}\underline{\tau}\rangle /(a_i)$$. The last two claims follow from Proposition A.14. ■ Remark A.23. Even though $$\hat{X}^{\text{cp}}$$ is a (weak) inverse limit of the spaces $${X}_h$$ as adic space, morphisms between two presentations of dagger spaces $$X$$ and $$Y$$ do not coincide in general with morphisms between $$\hat{X}^{\text{cp}}$$ and $$\hat{Y}^{\text{cp}}$$ as the latter coincide with morphisms from $$\hat{X}$$ to $$\hat{Y}$$. □ We can also promote the equivalence of categories between dagger spaces and their presentations to an equivalence of topoi, using the following definitions. Definition A.24. Let $$\mathbf{P}$$ be a property of morphisms of rigid spaces. We say that a morphism of pro-rigid spaces $$\phi \colon X \rightarrow Y$$ has the property $$\mathbf{P}$$ if $$X \cong\varprojlim X_h$$, $$Y^\dagger\cong\varprojlim Y_h$$ and $$\phi=\varprojlim \phi_h$$ with $$\phi_h\colon X_h\rightarrow Y_h$$ having the property $$\mathbf{P}$$. We say that a collection of open morphisms of pro-rigid spaces $$\{\phi_i\colon \varprojlim_h U_{ih}\rightarrow X \}_{i\in I}$$ is a cover if $$X\Subset \bigcup_i\text{Im}(U_{ih})$$ for all $$h$$. □ Remark A.25. In particular, we have defined open immersions, smooth and étale morphisms of presentations of affinoid dagger spaces. In that case, as the morphisms $$\hat{X}\subset X_h$$ are open immersions (hence étale) we deduce that if a morphism $$X \rightarrow Y$$ is an open immersion [resp. smooth resp. étale] then the associated morphism $$\hat{X}\rightarrow\hat{Y}$$ also is. □ Remark A.26. The topology induced by covers of open immersions is generated by covers of rational embeddings. □ Proposition A.27 Any family of étale maps of presentations of affinoid dagger spaces $$\{\varprojlim_h U_{ih}\rightarrow\varprojlim_h X_h\}_{i\in I}$$ inducing a covering of $$\hat{X}$$ is a covering. □ Proof. It suffices to show that if $$U\Subset U_1$$ and $$f\colon U_1\rightarrow X_1$$ is an étale map, then $$f(U)\Subset f(U_1)$$. By [9, Proposition 2.5.17] we can consider only the case in which $$f$$ is an open immersion, which is clear, and the case in which $$f$$ is finite étale, which we now examine. Since $$U_1\rightarrow f(U_1)$$ is finite, we deduce from [9, Proposition 2.5.8(iii) and Corollary 2.5.13(i)] that $$\text{Int}(U_1)=f^{-1}\text{Int}(X_1)$$. Therefore since $$U\Subset U_1$$ we deduce $$f(U)\Subset f(U_1)$$ as wanted. ■ Corollary A.28. Let $$X$$ be an affinoid dagger variety with a presentation $$\varprojlim X_h$$ and limit $$\hat{X}$$. The maps of the small rational and étale sites $$\hat{X}\rightarrow X \rightarrow\varprojlim X_h$$ induces equivalences on the associated topoi. □ Proof. It suffices to use the criteria of [33, Appendix A]. ■ Remark A.29. The content of the previous proposition may seem to clash with the result of Huber [33, Proposition 2.4.4] giving an equivalence between the étale topos of $$\hat{X}^{\text{cp}}$$ and the direct limit topos $$\varprojlim {\bf{Sh}}_{\acute{\mathrm{e}}\text{t}}(X_h)$$. The point is that the étale site $$\varinjlim X_{h,\acute{\mathrm{e}}\text{t}}$$ giving rise to the direct limit topos $$\varprojlim {\bf{Sh}}_{\acute{\mathrm{e}}\text{t}}(X_h)$$ is not equivalent to the étale site of the presentation. As a hint of this fact, consider for example the constant system $$U_h=\hat{X}$$ which is not a presentation of a dagger structure and hence does not define an open of $$\hat{X}$$. □ In order to be consistent with the “pro-objects” approach, we also introduce dagger spaces with a functorial perspective. We recall that the rational topology on dagger affinoid spaces is sub-canonical, that is, the presheaf $$\mathcal{F}_X$$ over $$\text{Aff}^\dagger$$ represented by an affinoid dagger variety $$X$$ is a sheaf (this follows from Proposition 2.12). By abuse of notation, we sometimes denote it by $$X$$. Definition A.30. A morphism $$\mathcal{F}\rightarrow\mathcal{G}$$ of sheaves of sets over $$\text{Aff}^\dagger$$ with the rational topology has the property $$\mathbf{P}$$ if for any morphism $$X \rightarrow \mathcal{G}$$ from a representable one, the pull-back is representable and the morphism $$\mathcal{F}\times_{\mathcal{G}}X \rightarrow X$$ has the property $$\mathbf{P}$$. A collection of morphisms $$\{\mathcal{F}_i\rightarrow\mathcal{G}\}$$ is a cover if $$\bigsqcup\mathcal{F}_i\rightarrow\mathcal{G}$$ is an epimorphism of sheaves. A [smooth] functorial dagger space is a sheaf $$\mathcal{F}$$ over $$\text{Aff}^\dagger$$ with a cover of open immersions $$\{U_i\rightarrow\mathcal{F}\}$$ where each $$U_i$$ is represented by a [smooth] affine dagger space. □ Remark A.31. The category of functorial dagger spaces has fiber products, generalizing fiber products of affinoid dagger spaces. □ Remark A.32. étale and open covers define a topology on functorial dagger spaces. □ The functor $$l\colon \text{Aff}^\dagger\rightarrow\text{Aff}$$ induced by $$X \mapsto \hat{X}$$ is continuous. If we embed the category of affinoid varieties in the category of locally ringed spaces $${\bf{LRS}}$$, we therefore obtain a left adjoint Kan extension functor: |⋅|:Sh(Aff†)⇄LRS such that $$|X |=(|\hat{X}|,\mathcal{O}_{\hat{X}})$$ for any $$X\in \text{Aff}^\dagger$$ (we denote by $$X$$ also the sheaf represented by $$X$$). Proposition A.33 Let $$\mathcal{F}$$ be a [smooth] functorial dagger space with an open cover by dagger affinoid spaces $$\{U_i \rightarrow\mathcal{F}\}$$. The ringed space $$|\mathcal{F}|$$ is a [smooth] rigid analytic space covered by $$\hat{U}_i$$ and endowed with an extra sheaf $$\mathcal{O}^\dagger$$ such that $$\mathcal{O}^\dagger|_{\hat{U}_i}$$ is the sheaf $$\mathcal{O}^\dagger_{U_i }$$ introduced in Definition 2.13. □ Proof. By [38, IV.7.3 and A.1.1] we can write a coequalizing diagram of sheaves ⨆(Ui×FUj)⇉⨆Ui→F→0 whose arrows on the left are open immersions of representable sheaves. By applying the left adjoint functor $$|\cdot|$$ we deduce that $$|\mathcal{F}|$$ is obtained by gluing the analytic spaces $$\hat{U}_i$$ over open immersions and therefore is an analytic space. The sheaf $$\mathcal{O}^\dagger$$ can be defined using the formula in the statement. ■ Remark A.34. In particular, we proved that a morphism of affine dagger spaces $$U \rightarrow X$$ is open or rational if an only if the associated map of representable sheaves is, and that a collection $$\{U_i \rightarrow X \}$$ of morphisms of affine dagger spaces is an open cover if and only if the induced collection of morphisms of sheaves is. □ We conclude that the functor $$|\cdot|$$ factors over the category of dagger spaces defined by Große-Klönne [23]. Corollary A.35. The category of functorial dagger spaces is equivalent to the category of dagger spaces of Definition 2.22, and this equivalence preserves rational/open immersions and covers. □ Proof. If $$X$$ is a dagger space with a dagger affinoid covering $$\{U_i\rightarrow X\}$$ we can consider the sheaf $$\mathcal{F}_X$$ it represents on the category of affinoid dagger spaces. Since the functor $$X\mapsto\mathcal{F}_X$$ is a right adjoint, it preserves intersections. We then conclude by Yoneda Lemma and Remark A.34 that $$\{\bigsqcup U_i\rightarrow \mathcal{F}_X\}$$ is an epimorphism of open immersions so that $$\mathcal{F}_X$$ is a functorial dagger space. It is immediate to see that $$X\mapsto\mathcal{F}_X$$ and $$\mathcal{F}\mapsto|\mathcal{F}|$$ are quasi-inverse functors. The second part of the statement follows from Remark A.34. ■ Funding This work was supported by the Lebesgue Center of Mathematics. Acknowledgements This paper was been written during my stay at the IRMAR of Rennes, funded by the Lebesgue Center of Mathematics. I am grateful to the hosting organizations for their support. I thank Jean-Yves Etesse for pointing out some relevant results in the literature. I also express my gratitude to Joseph Ayoub, Frédéric Déglise, Elmar Große-Klönne, Bernard Le Stum and an anonymous referee for their precious remarks on earlier versions of this work. References [1] Ayoub J. “A guide to (étale) motivic sheaves.” Proceedings of ICM II, 2014. Available at http://www.icm2014.org/en/vod/proceedings.html (accessed January 4, 2017) . [2] Ayoub J. “Motifs des variétés analytiques rigides.” Mémoire Société mathématique de France (New series) 2015 , no. 140–41 ( 2015 ): vi+386 . [3] Ayoub J. “Les six opérations de Grothendieck et le formalisme des cycles évanescents dans le monde motivique, I.” Astérisque 2007 , no. 314 ( 2008 ): x+466 . [4] Ayoub J. “Les six opérations de Grothendieck et le formalisme des cycles évanescents dans le monde motivique, II.” Astérisque 2007 , no. 315 ( 2008 ): vi+364 . [5] Ayoub J. “La réalisation étale et les opérations de Grothendieck.” Annales scientifiques de l’Ecole normale supérieure 47 , no. 4 ( 2014 ): 1 – 141 . [6] Ayoub J. “L’algèbre de Hopf et le groupe de Galois motiviques d’un corps de caractéristique nulle, I.” Journal für die Reine und Angewandte Mathematik. [Crelle’s Journal] 693 ( 2014 ): 1 – 149 . [7] Ayoub J. “L’algèbre de Hopf et le groupe de Galois motiviques d’un corps de caractéristique nulle, II.” Journal für die Reine und Angewandte Mathematik. [Crelle’s Journal] 693 ( 2014 ): 151 – 226 . [8] Beilinson A. and Vologodsky. V. “A DG guide to Voevodsky’s motives.” Geometric and Functional Analysis 17 , no. 6 ( 2008 ): 1709 – 87 . Google Scholar CrossRef Search ADS [9] Berkovich V. G. Spectral Theory and Analytic Geometry over Non-Archimedean Fields. Mathematical Surveys and Monographs 33 . Providence, RI : American Mathematical Society , 1990 . [10] Berkovich V. G. Integration of One-Forms on P-Adic Analytic Spaces . Annals of Mathematics Studies 162 . Princeton, NJ : Princeton University Press , 2007 . Google Scholar CrossRef Search ADS [11] Berthelot P. “Cohomologie Cristalline des Schémas de Caractéristique $$p>0$$.” Lecture Notes in Mathematics 407 . Berlin, New York : Springer , 1974 . [12] Berthelot P. “Finitude et pureté cohomologique en cohomologie rigide.” Inventiones Mathematicae 128 , no. 2 ( 1997 ): 329 – 77 . With an appendix in English by Aise Johan de Jong . Google Scholar CrossRef Search ADS [13] Besser A. “Syntomic regulators and $$p$$-adic integration. I. Rigid syntomic regulators.” Proceedings of the Conference on p-adic Aspects of the Theory of Automorphic Representations (Jerusalem, 1998), Israel Journal of Mathematics 120 , no. 2 ( 2000 ): 291 . [14] Bosch S. “A rigid analytic version of M. Artin’s theorem on analytic equations.” Mathematische Annalen 255 , no. 3 ( 1981 ): 395 – 404 . Google Scholar CrossRef Search ADS [15] Bosch S. Dwork B. and Robba. P. “Un théorème de prolongement pour des fonctions analytiques.” Mathematische Annalen 252 , no. 2 ( 1979/80 ): 165 – 73 . Google Scholar CrossRef Search ADS [16] Bosch S. Güntzer U. and Remmert. R. Non-Archimedean Analysis , Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] 261 , Berlin : Springer , 1984 . A systematic approach to rigid analytic geometry . Google Scholar CrossRef Search ADS [17] Chiarellotto B. and Tsuzuki N. “Cohomological descent of rigid cohomology for étale coverings.” Rendiconti del Seminario Matematico della Università di Padova 109 ( 2003 ): 63 – 215 . [18] Cisinski D.-C. and Déglise. F. “Mixed Weil cohomologies.” Advances in Mathematics 230 , no. 1 ( 2012 ): 55 – 130 . Google Scholar CrossRef Search ADS [19] Cisinski D.-C. and Déglise. F. Triangulated categories of mixed motives , arXiv: 0912.2110v3 [math.AG] , 2009 . [20] Déglise F. and Mazzari. N. “The rigid syntomic ring spectrum.” Journal of the Institute of Mathematics of Jussieu 14 , no. 4 ( 2015 ): 753 – 99 . Google Scholar CrossRef Search ADS [21] Etesse J.-Y. “Relèvements de schémas et algèbres de Monsky-Washnitzer: théorèmes d’équivalences et de pleine fidélité. II.” Rendiconti del Seminario Matematico della Università di Padova 122 ( 2009 ): 205 – 34 . Google Scholar CrossRef Search ADS [22] Fresnel J. and van der Put. M. Rigid Analytic Geometry and its Applications . Progress in Mathematics 218 . Boston, MA : Birkhäuser Boston , 2004 . Google Scholar CrossRef Search ADS [23] Große-Klönne E. “Rigid analytic spaces with overconvergent structure sheaf.” Journal für die Reine und Angewandte Mathematik. [Crelle’s Journal] 519 ( 2000 ): 73 – 95 . [24] Große-Klönne E. “Finiteness of de Rham cohomology in rigid analysis.” Duke Mathematical Journal 113 , no. 1 ( 2002 ): 57 – 91 . Google Scholar CrossRef Search ADS [25] Große-Klönne E. “De Rham cohomology of rigid spaces.” Mathematische Zeitschrift 247 , no. 2 ( 2004 ): 223 – 40 . Google Scholar CrossRef Search ADS [26] Grothendieck A. “Éléments de géométrie algébrique. I. Le langage des schémas.” Institut des Hautes études Scientifiques. Publications Mathématiques 228 , no. 4 ( 1960 ): 5 – 228 . Google Scholar CrossRef Search ADS [27] Grothendieck A. Fondements de la géométrie algébrique. [Extraits du Séminaire Bourbaki, 1957–1962.] . Paris : Secrétariat mathématique , 1962 . [28] Hirschhorn P. S. Model Categories and Their Localizations . Mathematical Surveys and Monographs 99 . Providence, RI : American Mathematical Society , 2003 . [29] Hovey Mark Model Categories . Mathematical Surveys and Monographs 63 . Providence, RI : American Mathematical Society , 1999 . [30] Hovey M. “Spectra and symmetric spectra in general model categories.” Journal of Pure and Applied Algebra 165 , no. 1 ( 2001 ): 63 – 127 . Google Scholar CrossRef Search ADS [31] Huber R. “Continuous valuations.” Mathematische Zeitschrift 212 , no. 3 ( 1993 ): 455 – 77 . [32] Huber R. “A generalization of formal schemes and rigid analytic varieties.” Mathematische Zeitschrift 217 , no. 4 ( 1994 ): 513 – 51 . Google Scholar CrossRef Search ADS [33] Huber R. Étale Cohomology of Rigid Analytic Varieties and Adic Spaces . Aspects of Mathematics E30 . Braunschweig : Friedr. Vieweg & Sohn , 1996 . Google Scholar CrossRef Search ADS [34] Jardine J. F. “Simplicial presheaves.” Journal of Pure and Applied Algebra 47 , no. 1 ( 1987 ): 35 – 87 . Google Scholar CrossRef Search ADS [35] Kedlaya K. S. “Finiteness of rigid cohomology with coefficients.” Duke Mathematical Journal 134 , no. 1 ( 2006 ): 15 – 97 . Google Scholar CrossRef Search ADS [36] Künneth H. “Über die Bettischen Zahlen einer Produktmannigfaltigkeit.” Mathematische Annalen 90 , no. 1–2 ( 1923 ): 65 – 85 . Google Scholar CrossRef Search ADS [37] Le Stum B. “The overconvergent site.” Mémoires de la Société Mathématique de France. Nouvelle Série ( 2011 ), 127 ( 2012 ): vi+108 pp. [38] Mac Lane S. and Moerdijk. I. Sheaves in Geometry and Logic . New York : Universitext, Springer , 1994 . A first introduction to topos theory, Corrected reprint of the 1992 edition . [39] Mazza C. , Voevodsky V. and Weibel. C. Lecture Notes on Motivic Cohomology . Clay Mathematics Monographs 2 . Providence, RI : American Mathematical Society , 2006 . [40] Mebkhout Z. “Sur le théorème de finitude de la cohomologie $$p$$-adique d’une variété affine non singulière.” American Journal of Mathematics 119 , no. 5 ( 1997 ): 1027 – 81 . Google Scholar CrossRef Search ADS [41] Milne J. S. and Ramachandran. N. Motivic complexes and special values of zeta functions , arXiv:1311.3166 [math.AG] , 2013 . [42] Monsky P. and Washnitzer. G. “Formal cohomology. I.” Annals of Mathematics. Second Series 88 ( 1968 ): 181 – 217 . Google Scholar CrossRef Search ADS [43] Morel F. and Voevodsky. V. “$${\bf A^1}$$-homotopy theory of schemes.” Institut des Hautes études Scientifiques. Publications Mathématiques ( 1999 ), 90 ( 2001 ): 45 – 143 . Google Scholar CrossRef Search ADS [44] Munkres J. R. Topology: A First Course . Englewood Cliffs, NJ : Prentice-Hall , 1975 . [45] Neeman A. “The connection between the $$K$$-theory localization theorem of Thomason, Trobaugh and Yao and the smashing subcategories of Bousfield and Ravenel.” Annales Scientifiques de l’école Normale Supérieure. Quatrième Série 25 , no. 5 ( 1992 ): 547 – 66 . Google Scholar CrossRef Search ADS [46] Ogus A. The Convergent Topos in Characteristic $$p$$. The Grothendieck Festschrift , edited by Cartier P. Katz N. M. Manin Y. I. Illusie L. Laumon G. and Ribet K. A. vol. III , 133 – 62 . Progress in Mathematics 88 , Boston, MA : Birkhäuser Boston , 1990 . [47] Riou J. “Catégorie homotopique stable d’un site suspendu avec intervalle.” Bulletin de la Société Mathématique de France 135 , no. 4 ( 2007 ): 495 – 547 . Google Scholar CrossRef Search ADS [48] Scholze P. “Perfectoid spaces.” Publications Mathématiques. Institut de Hautes études Scientifiques 116 ( 2012 ): 245 – 313 . Google Scholar CrossRef Search ADS [49] The Stacks Project Authors , Stacks Project , 2014 , http://stacks.math.columbia.edu (accessed on January 4, 2017) . [50] Temkin M. “On local properties of non-Archimedean analytic spaces.” Mathematische Annalen 318 , no. 3 ( 2000 ): 585 – 607 . Google Scholar CrossRef Search ADS [51] Temkin M. “On local properties of non-Archimedean analytic spaces. II.” Israel Journal of Mathematics 140 ( 2004 ): 1 – 27 . Google Scholar CrossRef Search ADS [52] Vezzani A. “Effective motives with and without transfers in characteristic $$p$$.” Advances in Mathematics 306 ( 2017 ): 852 – 79 . Google Scholar CrossRef Search ADS [53] Vezzani A. A motivic version of the theorem of Fontaine and Wintenberger , arXiv:1405.4548 [math.AG] , 2014 . [54] Wedhorn T. Adic spaces , Preprint , 2012 . © The Author 2017. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permission@oup.com. This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model) TI - The Monsky–Washnitzer and the overconvergent realizations JF - International Mathematics Research Notices DO - 10.1093/imrn/rnw335 DA - 2017-02-06 UR - https://www.deepdyve.com/lp/oxford-university-press/the-monsky-washnitzer-and-the-overconvergent-realizations-C4axf4X3DA SP - 1 EP - 3489 VL - Advance Article IS - 11 DP - DeepDyve ER -