TY - JOUR AU - Cho, Sungmun AB - Abstract In this article, we explain a simple and uniform construction of a smooth integral model associated to a quadratic, (anti)-hermitian, and (anti)-quaternionic hermitian lattice defined over an arbitrary local field. As one major application, we explain a conjectural recipe for computing local densities case by case, which is an essential factor in the classification of forms as above over the ring of integers of a number field, by introducing one conjecture about the number of rational points of the special fiber of a smooth integral model. 1 Introduction A long standing central problem in the arithmetic theory of hermitian (or quadratic) forms is the classification of hermitian (or quadratic) forms over the ring of integers of a number field. If we let $$\mathfrak{O}$$ be the ring of integers in a number field $$k$$, then for a totally definite hermitian (or quadratic) $$\mathfrak{O}$$-lattice $$(L, H)$$, the genus of $$(L, H)$$, denoted by $$\mathrm{gen}(L, H)$$, is defined as the set of (equivalence classes of) hermitian (or quadratic) lattices that are locally equivalent to $$(L, H)$$. Since the local-global principle does not hold for a hermitian (or quadratic) lattice $$(L, H)$$, the set $$\mathrm{gen}(L, H)$$ is not trivial in general. However, it is well-known that $$\mathrm{gen}(L, H)$$ is a finite set. The computation of the total mass of $$(L, H)$$ ($$=\sum_{(L^{\prime}, H^{\prime}) \in \mathrm{gen}(L, H)} \frac{1}{\#\mathrm{Aut}_{\mathfrak{O}}(L^{\prime}, H^{\prime})}$$) is an essential ingredient for enumerating all elements of the set $$\mathrm{gen}(L, H)$$ explicitly. The total mass of $$(L, H)$$ can be expressed as a product of local factors, the so-called local densities, by the celebrated Smith–Minkowski–Siegel mass formula. The local density is defined as follows. To simplify notation, we consider a quadratic $$A$$-lattice $$(L, h)$$ at this moment. Here, $$A$$ is the ring of integers of a local field $$F$$ with a uniformizer $$\pi$$ and the residue field $$\kappa$$. The local density was originally defined as the limit of a certain sequence [10], which is described below: βL=12⋅limN→∞q−NdimG#G_′(A/πNA). Here $$\underline{G}^{\prime}$$ is a naive integral model of the orthogonal group $$\mathrm{O}(V, h)$$, where $$V=L\otimes_AF$$, such that $$\underline{G}^{\prime}(R)=\mathrm{Aut}_{R}(L\otimes_AR, h\otimes_AR)$$ for any commutative $$A$$-algebra $$R$$, and $$q$$ is the cardinality of $$\kappa$$. This limit stabilizes after finite steps, and the formula is found in [5] and [8] for $$\mathbb{Z}_p$$. However, for a given quadratic lattice defined over a finite (especially ramified) extension of $$\mathbb{Z}_2$$, it is a nontrivial task to compute the above limit. Later, W. T. Gan and J.-K. Yu found another approach for computing local densities in [6]. (We explain their approach here briefly and for a detailed exposition including a general lattice such as a hermitian lattice, see Section 3 of [6].) The local density of a quadratic $$A$$-lattice $$(L, h)$$ can also be defined as an integral of a certain volume form $$\omega^{\mathrm{ld}}$$ associated to $$\underline{G}^{\prime}$$, which is described below: βL=12∫G_′(A)|ωld|. On the other hand, the following integral is known. ∫G_′(A)|ωcan|=q−dimG⋅#G_(κ). Here $$\underline{G}$$ is the unique smooth affine group scheme (called a smooth integral model) of $$\mathrm{O}(V, h)$$ such that $$\underline{G}(R)=\underline{G}^{\prime}(R)$$ for any étale $$A$$-algebra $$R$$ (Proposition 3.7 in [6]) and $$\omega^{\mathrm{can}}$$ is a volume form associated to $$\underline{G}$$. Therefore, in order to obtain an explicit formula for the local density, it suffices to determine $$\underline{G}$$ and the number of rational points of the special fiber of $$\underline{G}$$ (i.e., $$\# \underline{G}(\kappa)$$); compare two volume forms $$\omega^{\mathrm{ld}}$$ and $$\omega^{\mathrm{can}}$$. The difference between the two volume forms can be calculated directly from the construction of the smooth integral model $$\underline{G}$$. Therefore, constructing the smooth integral model $$\underline{G}$$ and investigating the special fiber of $$\underline{G}$$ will lead us to an explicit formula. The local density formula together with a smooth integral model $$\underline{G}$$ and its special fiber has been fully studied in [6] $$(p\neq 2)$$, [2] (quadratic lattice with $$A/\mathbb{Z}_2$$ unramified), and [3] and [4] (ramified hermitian lattice with $$A/\mathbb{Z}_2$$ unramified). Indeed, the constructions of $$\underline{G}$$ in [2–4], when $$A/\mathbb{Z}_2$$ is unramified, are much more complicated than that of [6] when $$p\neq 2$$. More precisely, if $$p\neq 2$$, then the construction of $$\underline{G}$$ is based on the dual lattice $$L^{\#}$$ of $$L$$. In the case that $$A/\mathbb{Z}_2$$ is unramified, one needs to consider a series of sublattices of $$L$$ in order to construct $$\underline{G}$$. Therefore, if $$A/\mathbb{Z}_2$$ is ramified, then it is natural to expect that one needs much more involved sublattices of $$L$$ beyond a series of sublattices of $$L$$ used in the unramified case in order to construct $$\underline{G}$$. One can also expect that a construction of $$\underline{G}$$ in a quadratic lattice is different from that of a hermitian lattice, as in [2–4]. There are two main parts in this article. In the first main part (Section 3), we explain a uniform construction of a smooth integral model $$\underline{G}$$ associated to a quadratic or a hermitian lattice over any non-Archimedean local field without taking account of any sublattice of $$L$$. Our construction of $$\underline{G}$$ is simple and canonical (independent of the choice of a basis of $$L$$). As one major application, we obtain the following theorem about the local density for a given lattice $$(L, h)$$: Theorem 1.1. (Theorem 3.25) Let $$\tilde{G}$$ (respectively $$G$$) be the special fiber (respectively generic fiber) of $$\underline{G}$$ and let $$q$$ be the cardinality of $$\kappa$$. Then the local density of ($$L,h$$) is βL=1[G:Go]qN⋅q−dim G⋅#G~(κ), where $$G^o$$ is the identity component of $$G$$ and $$N$$ is a suitable integer. □ Therefore, in order to compute the local density after constructing $$\underline{G}$$, we only have to know two ingredients in the above theorem; $$q^N$$ and $$\#\tilde{G}(\kappa)$$. Here $$q^N$$ quantifies the difference between the two volume forms $$\omega^{\mathrm{ld}}$$ and $$\omega^{\mathrm{can}}$$ and this can be computed easily based on the construction of $$\underline{G}$$, as explained in Theorem 3.25. Thus, once we construct $$\underline{G}$$, the remaining challenging ingredient is the computation of $$\#\tilde{G}(\kappa)$$. The second main part of this article (Section 4) is devoted to explaining a conjectural recipe for computing $$\#\tilde{G}(\kappa)$$ case by case. For a given quadratic lattice $$(L, h)$$, let $$L=\bigoplus_{i\geq 0}L_i$$ be a Jordan splitting of $$(L, h)$$. Then there exists a morphism as algebraic groups $$\varphi_{i} : \tilde{G} \rightarrow \mathrm{O}(\bar{V_i}, \bar{h_i})$$ which is naturally induced from the structure of $$(L, h)$$ (Proposition 4.9). Here, $$\bar{V_i}$$ and $$\bar{h_i}$$ are defined in Proposition 4.9. Let φ=∏iφi:G~⟶∏iO(Vi¯,hi¯). The formation of $$\varphi$$ is compatible with finite unramified extension of $$A$$. To compute $$\#\tilde{G}(\kappa),$$ we propose the following conjecture. Conjecture 1.2. (Conjecture 4.10) (1) The morphism $$\varphi$$ is surjective. (2) The smooth (possibly disconnected) algebraic group $$U := (\mathrm{ker~}\varphi)_{\rm{red}}$$ is unipotent with constant component group. □ Applying this general conjecture after any unramified extension of $$A$$ implies the surjectivity in (1) on $$\kappa'$$-points for all finite extensions $$\kappa'$$ of $$\kappa$$. Unipotence implies the finite étale component group $$\pi_0(U) = U/U^0$$ has order $$2^{\beta}$$ for some $$\beta$$, and the identity component $$U^0$$ is isomorphic to an affine space $$\mathbb{A}^{\ell}$$ for some $$\ell$$ (as for any smooth unipotent group over a perfect field of characteristic 2). Here, $$\ell={\mathrm{dim}}$$$$\tilde{G}$$ - $$\sum_i \mathrm{dim}\,\,\, \mathrm{O}(\bar{V_i}, \bar{h_i})$$. The dimension of $$\tilde{G}$$ is the same as the dimension of $$G$$ which is $$n(n-1)/2$$, where $$G$$ is the generic fiber of $$\underline{G}$$ and $$n$$ is the rank of $$L$$, since $$\underline{G}$$ is flat over $$A$$. Once one verifies the conjecture, we obtain that #G~(κ)=∏i#(O(Vi¯,hi¯)(κ))⋅qℓ⋅2β. The conjecture tells us more than $$\#\tilde{G}(\kappa)$$. Since we conjecture that $$\varphi_{\kappa'}$$ is surjective for every finite extension $$\kappa'$$ of $$\kappa$$, it follows that the local density is given by a rational function of $$q$$ as we extend the base ring $$A$$ to finite unramified extensions $$A'$$ (cf. Remark 4.11). This had not been expected before this article. The conjecture appears to be very difficult in the general case. For the special cases when $$A$$ is a finite unramified extension of $$\mathbb{Z}_2$$ in [2–4], we proved this by writing down formal matrix forms of an element of $$\tilde{G}(\kappa)$$ and the group homomorphism $$\varphi_{\kappa}$$ explicitly based on a matrix form of an element of $$\underline{G}(A)$$. In the general case, it seems unlikely that such formal matrix forms can be written because the classification of hermitian (or quadratic) lattices over a ramified extension of $$\mathbb{Z}_2$$ is a highly involved problem [9, 12, 13]. However, the construction of $$\underline{G}$$ studied in this article gives an explicit matrix form of an element of $$\underline{G}(A)$$ which yields explicit formal matrix forms of $$\varphi_{\kappa}$$ as well as an element of $$\tilde{G}(\kappa)$$, case by case. In Remark 4.12, we explain a possible framework to prove the conjecture case by case, based on the proof of the conjecture in a specific case as given in Theorem 4.4 (unimodular quadratic lattices with odd rank). In order to show how the framework in Remark 4.12 works case by case, we provide Examples 4.13–4.16 (non-unimodular quadratic lattices) and prove the conjecture in those cases. Especially, Examples 4.14–16 and Remark 4.17 serve as evidence for constancy of the component group of $$(\mathrm{ker~}\varphi)_{\rm{red}}$$. Example 4.14 serves as evidence that the special fiber (and the local density) depends on lower ramification groups of the Galois group as well as the ramification index, which had not been known before this article. In Remark 4.17, we explain one expected phenomenon about the number of component groups (i.e., the value of $$\beta$$) in terms of “distance” between Jordan components. This article does not render the papers [2–4, 6] obsolete since we do not offer any new method to prove the conjecture. One of the main contributions in those articles is to prove special cases of the conjecture (for an unramified extension of $$\mathbb{Z}_2$$). Consequently, by assuming that the above conjecture is true, this article gives a new formula in terms of geometrically meaningful invariants for computing the local density of a given quadratic (or hermitian) lattice case by case, especially, defined over an arbitrary ramified extension of $$\mathbb{Z}_2$$. One strength of this conjectural recipe is that all computations for local densities of given lattices are reduced to computations on finitely generated $$A$$-modules (conditionally depending on Conjecture 4.10 and Remark 4.12), limiting the language of schemes to as little as possible. This article is organized as follows. After fixing the notations in Section 2, we explain a uniform construction of a smooth integral model $$\underline{G}$$ in Section 3. The constructions of $$\underline{G}$$ in [2–4, 6] are special and simple cases (at most $$\alpha=2$$) of the construction studied in this article, which is explained in Remark 3.23. Then we give the local density formula in Theorem 3.25. In the first part of Section 4 (4.1–4.7), we work out completely the mass of unimodular quadratic lattices in odd dimension. In the second part of Section 4 (4.8-4.17), we explain a conjectural recipe for computing $$\#\tilde{G}(\kappa)$$ in the general case by introducing Conjecture 4.10 and Remark 4.12 about the number of rational points of the special fiber of $$\underline{G}$$. We wish to point out that our method also works for symplectic lattices, hermitian lattices, anti-hermitian lattices, quaternionic hermitian lattices, and anti-quaternionic hermitian lattices over any non-Archimedean local field. 2 Notations This section is taken from [6]. 2.0.1 Let $$F$$ be a non-Archimedean local field with $$A$$ its ring of integers and $$\kappa$$ its residue field. Let $$p$$ be the residue characteristic of $$F$$. Choose a uniformizing element $$\pi$$ of $$A$$. 2.0.2 Let $$(K, \sigma)$$ be one of the following $$F$$-algebras with involution: $$K=F$$ with $$\textit{char } F\neq 2$$, $$\sigma$$ identity; $$K=E$$ a separable quadratic extension, $$\sigma$$ the unique nontrivial automorphism of $$E/F$$; $$K=F\oplus F$$, $$\sigma(x,y)=(y,x)$$; $$K=D$$ is the unique nonsplit quaternion algebra over $$F$$, $$\sigma$$ the standard involution. 2.0.3 Let $$B$$ be a maximal $$A$$-order in $$K$$. Then $$B$$ is uniquely determined. If $$K=E$$ is a ramified quadratic extension of $$F$$, or if $$K=D$$, we let $$\pi_K$$ be a uniformizer of $$K$$; in all other cases, we let $$\pi_K=\pi$$. Note that $$\pi$$ in $$K=F\oplus F$$ means $$(\pi, \pi)$$. 2.0.4 Let $$\epsilon$$ be either $$1$$ or $$-1$$. The triple $$(K, \sigma, \epsilon)$$ will be fixed throughout this article, and by a hermitian form we always mean a $$(\sigma, \epsilon)$$-hermitian form. If $$(K, \epsilon, p)\neq (F, 1, 2)$$, we consider a $$B$$-lattice $$L$$ (i.e., a free right$$B$$-module of finite rank) of rank $$n$$ with a hermitian form $$h : L \times L \rightarrow B.$$ Our convention is h(v⋅a,w⋅b)=σ(a)h(v,w)b, h(w,v)=ϵσ(h(v,w)). If $$(K, \epsilon, p)= (F, 1, 2)$$, we consider a $$B (=A)$$-lattice $$L$$ with a quadratic form $$h : L \rightarrow B$$, and we write $$h(v, w)=1/2\cdot (h(v+w)-h(v)-h(w))$$. When $$K=F$$, the characteristic of $$F$$ is not 2 as explained in Section 2.0.2. We assume that $$V=L\otimes_AF$$ is nondegenerate with respect to $$h$$. The right $$B$$-module $$L$$ is also regarded as a left $$B$$-module by the rule $$a\cdot v=v\cdot \sigma(a)$$. Definition 2.1. We define the dual lattice of $$L$$, denoted by $$L^{\#}$$, as L#={x∈L⊗AF:h(x,L)⊂B}. □ The pair ($$L, h$$) is fixed throughout this article. Lemma 2.2. For a given pair ($$L, h$$), we have the following identity: (L#)#=L. □ This lemma can be deduced from the fact that $$L^{\#}$$ is isomorphic to $$\mathrm{Hom}_B(L, B)$$ with its right $$B$$-module structure via $$\sigma$$, from which the double duality follows automatically, see for example, II.2.2 of [11]. 2.2.1 Let $$G$$ be the reductive algebraic group over $$F$$ such that G(F)=AutK⊗FF(V⊗FF,h⊗FF) for any commutative $$F$$-algebra $$\mathfrak{F}$$. Then $$G$$ is a classical group, not necessarily connected. Indeed, it is essential that $$G$$ is smooth in our construction of the smooth integral model $$\underline{G}$$ due to Theorem 3.1 below. Thus we exclude the case in Subsection 2.0.2 that $$G$$ is an orthogonal group over a field of characteristic 2, that is, the case $$K=F$$ with $$\textit{char } F= 2$$ and $$\sigma$$ is the identity. We denote by $$\underline{\mathrm{GL}_K(V)}_{/F}$$ the $$F$$-group scheme whose group of $$\mathfrak{F}$$-valued points is $$\mathrm{GL}_{K\otimes_F\mathfrak{F}}(V\otimes_F\mathfrak{F})$$ for any commutative $$F$$-algebra $$\mathfrak{F}$$. If $$K$$ is commutative, this is just the Weil restriction $$\mathrm{Res}_{K/F}\underline{\mathrm{GL}_K(V)}$$. 3 Smooth integral model We start with the following theorem which is crucially used in this article. Theorem 3.1. (Proposition 3.7 in [6]) Assume that $$A$$ is a discrete valuation ring. Let $$\underline{G}^{\prime}$$ be an affine group scheme over $$A$$ of finite type with smooth generic fiber $$G$$. Then there exists a unique smooth affine group scheme $$\underline{G}$$ over $$A$$ with generic fiber $$G$$ such that G_(R)=G_′(R)for any étale A-algebra R. □ Let $$\underline{G}^{\prime}$$ be the naive integral model of $$G$$ such that for any commutative $$A$$-algebra $$R$$, $$\underline{G}^{\prime}(R)=\mathrm{Aut}_{B\otimes_AR}(L\otimes_AR, h\otimes_AR)$$. Then, the above theorem guarantees the existence of a unique smooth integral model $$\underline{G}$$ of $$G$$ such that $$\underline{G}(R)=\underline{G}^{\prime}(R)$$ for any étale $$A$$-algebra $$R$$. In this section, we give an explicit construction of the smooth integral model $$\underline{G}$$. We briefly explain the strategy of our construction. The naive integral model $$\underline{G}^{\prime}$$ is defined as the closed subgroup scheme of the group of invertible elements in $$\mathrm{End}_{B}(L)$$ stabilizing the hermitian form $$h$$. We will first construct a certain congruence subgroup of $$\mathrm{End}_{B}(L)$$ (denoted by $$\widetilde{T}$$) from a sequence of functors $$T^m$$ from the category of flat $$A$$-algebras to the category of abelian groups. Then we will show that the closed subgroup scheme of the group of invertible elements in $$1+\widetilde{T}$$ (denoted by $$\underline{M}$$) stabilizing the hermitian form $$h$$ is the desired smooth integral model $$\underline{G}$$. We will show that the functors $$T^m$$ are represented by a polynomial ring in some number of variables over $$A$$, in the following sense. Firstly, $$T^m$$ is represented as a sheaf on the small fppf site over $$A$$. Secondly, a polynomial ring in some number of variables over $$A$$ means a ring of the form $$A[t_1, \cdots, t_l]$$, where $$t_i$$’s are algebraically independent over $$A$$. Finally, by saying that a functor $$T^m$$ is represented by a polynomial ring $$A[t_1, \cdots, t_l]$$, we mean that $$T^m$$ is the functor of points of an affine space $$\mathbb{A}^l$$. 3.2 Constructions of $$T^0$$ and $$H^0$$ based on [6] We first define a functor $$T^0$$ from the category of commutative flat $$A$$-algebras to the category of rings. For any commutative flat $$A$$-algebra $$R$$, set T0(R)={X∈EndB⊗AR(L⊗AR):X(L#⊗AR)⊂L#⊗AR}. The functor $$T^0$$ is represented by a unique flat $$A$$-algebra which is a polynomial ring over $$A$$ in $$n^2\cdot[K:F]$$ variables (cf. Section 5.2 of [6]). Moreover, it is easy to see that $$T^0$$ has the structure of a ring scheme since $$T^0(R)$$ is closed under addition and multiplication. For future use, we state the following proposition. Proposition 3.3. For a flat $$A$$-algebra $$R$$, choose $$X \in T^0(R)$$. We define the adjoint $$X^{ad}$$ of $$X$$ characterized as $$h(X(v), w)=h(v, X^{ad}(w))$$, where $$v, w \in L \otimes_A R$$. Then we have that Xad∈T0(R) and (Xad)ad=X. □ Proof. By Lemma 2.2, it is easy to see that $$X^{ad}$$ stabilizes $$L^{\#}$$ as well as $$L$$. This completes the proof. ■ The proposition yields that $$X^{ad}\cdot Y\in T^0(R)$$, where $$X, Y \in T^0(R)$$. This will be used later on. We define another functor $$H^0$$ from the category of commutative flat $$A$$-algebras to the category of abelian groups. For any commutative flat $$A$$-algebra $$R$$, set H0(R)={f : f is an hermitian form on L⊗AR such that f(L⊗AR,L#⊗AR)⊂B⊗AR}. The functor $$H^0$$ is also represented by a flat $$A$$-algebra which is a polynomial ring over $$A$$ in $$n^2\cdot[K:F]-\mathrm{dim~}G$$ variables (cf. Section 5.4 of [6]). Moreover, it is easy to see that $$H^0$$ has the structure of an abelian group scheme since $$H^0(R)$$ is closed under addition. The fixed hermitian form $$h$$ is an element of $$H^0(A)$$. 3.4 Constructions of three functors $$\boldsymbol{\varphi_{0}}$$, $$\boldsymbol{\psi_{0}}$$, $$\boldsymbol{\bar{\varphi}_{0}}$$ 3.4.1 We define a morphism of functors $$\varphi_0 : T^0 \rightarrow H^0$$ defined on a flat $$A$$-algebra $$R$$ by φ0,R:T0(R)⟶H0(R), X↦h∘X. Here, $$h\circ X$$ is a hermitian form on $$L\otimes_AR$$ such that $$h\circ X(v, w)=h(X(v), X(w))$$ for $$v, w \in L\otimes_AR$$. Then $$\varphi_0$$ is represented by a morphism of $$A$$-schemes. 3.4.2 Let $$R$$ be a commutative $$A$$-algebra. Since $$T^0$$ is represented by a polynomial ring in some number of variables, the fiber of the identity in the morphism $$T^0(R[\varepsilon]/\varepsilon^2) \rightarrow T^0(R)$$ is naturally identified with $$T^0(R)$$. Similarly, the fiber of the fixed hermitian form $$h$$ in the morphism $$H^0(R[\varepsilon]/\varepsilon^2) \rightarrow H^0(R)$$ is identified with $$H^0(R)$$. Then the morphism $$\varphi_0$$ induces a morphism of functors $$\psi_0 : T^0 \rightarrow H^0$$ defined on a flat $$A$$-algebra $$R$$ by ψ0,R:T0(R)⟶H0(R), ψ0,R(X)(v,w)=h(v,X(w))+h(X(v),w)). Both $$T^0(R)$$ and $$H^0(R)$$ have $$R$$-module structures and that $$\psi_{0, R}$$ is an $$R$$-module homomorphism. Like $$\varphi_0$$, the map $$\psi_{0}$$ is represented by a morphism of $$A$$-schemes. Indeed, $$\psi_0$$ is the differential of $$\varphi_0$$ and is $$A$$-linear. 3.4.3 We define a morphism of functors $$\bar{\varphi}_0 : T^0 \rightarrow \mathrm{coker}(\psi_0)$$ defined on a flat $$A$$-algebra $$R$$ by φ¯0,R:T0(R)⟶H0(R)⟶coker(ψ0)(R). Here, the first map is $$\varphi_{0, R}$$ and the second map is the quotient map. Note that $$\mathrm{coker}(\psi_0)$$ is a representable sheaf on the small fppf site over $$A$$ such that $$\mathrm{coker}(\psi_0)(R)=H^0(R)/\mathrm{Im~}\psi_{0, R}$$ since $$\psi_{0, R}$$ is a morphism of $$R$$-modules (cf. Lemma 6.3.3 in [6]). Then $$\mathrm{coker}(\psi_0)(R)$$ is an abelian group since it is a quotient as $$R$$-modules. The map $$\bar{\varphi}_{0, R}$$ is then a group homomorphism since $$\varphi_{0, R}(X+Y)=\varphi_{0, R}(X)+\varphi_{0, R}(Y)+\psi_{0, R}(X^{ad}\cdot Y)$$ for $$X, Y \in T^0(R)$$. Here, $$X^{ad}\cdot Y \in T^0(R)$$ by Proposition 3.3 and thus $$\psi_{0, R}(X^{ad}\cdot Y) \in \mathrm{Im~}\psi_{0, R}$$. We point out that $$\bar{\varphi}_{0, R}$$ is not a morphism of $$R$$-modules even though it is a morphism of abelian groups. 3.5 Construction of $$T^1$$ We define the functor $$T^1$$ from the category of commutative flat $$A$$-algebras to the category of abelian groups as follows: T1(R)={ker(φ¯0,A)∩ker(φ¯0,A)adif R=A;R⊗AT1(A)if R is a flat A-algebra.  Note that $$\mathrm{ker}(\bar{\varphi}_{0, A})\cap \mathrm{ker}(\bar{\varphi}_{0, A})^{ad}$$ can also be characterized as the set $$\{X\in \mathrm{ker~}\bar{\varphi}_{0, A} : X^{ad} \in \mathrm{ker~}\bar{\varphi}_{0, A}\}$$ by using Proposition 3.3. To show that the functor $$T^1$$ is well-defined, it is enough to show that the set $$\mathrm{ker}(\bar{\varphi}_{0, A})\cap \mathrm{ker}(\bar{\varphi}_{0, A})^{ad}$$ is an $$A$$-module, which is obvious. We can also see that $$T^1(R)$$ is an $$R$$-submodule of $$T^0(R)$$ for a flat $$A$$-algebra $$R$$ since $$T^0(R)=R\otimes_AT^0(A)$$. The abelian group $$T^1(R)$$ has the following description for an étale $$A$$-algebra $$R$$. Theorem 3.6. Let $$R$$ be a flat $$A$$-algebra. Then $$R\otimes_AT^1(A) \subseteq \mathrm{ker}(\bar{\varphi}_{0, R})\cap \mathrm{ker}(\bar{\varphi}_{0, R})^{ad}$$ as $$R$$-submodules of $$T^0(R)$$. The equality holds assuming that $$R$$ is étale over $$A$$. □ Proof. The first claim is obvious. For the second claim, let $$R$$ be an étale local ring over $$A$$. Such $$R$$ is finite over $$A$$ since any étale local ring $$R$$ over a henselian local ring is finite by Proposition 4 of Section 2.3 in [1] and $$A$$ is henselian by the assumption made in Subsection 2.0.1. Then the morphism $$\textit{Spec R}\rightarrow \textit{Spec A}$$ is a Galois covering (cf. Section 6.2, Example B in [1]). We remark that $$T^0(R)=R\otimes_AT^0(A)$$ and $$H^0(R)=R\otimes_AH^0(A)$$. It is, then, easy to show that the set $$\mathrm{ker}(\bar{\varphi}_{0, A})\cap \mathrm{ker}(\bar{\varphi}_{0, A})^{ad}$$ is stabilized by any element of $$\mathrm{Aut}_A(R)$$. Therefore, the above set descends to an $$A$$-submodule $$\mathcal{M}$$ of $$T^0(A)$$ by Galois Descent. Choose $$X\in \mathcal{M}$$. Since $$\varphi_{0, R}(X), \varphi_{0, R}(X^{ad})\in \mathrm{Im~}\psi_{0, R} \cap H^0(A)$$, we can see that $$\varphi_{0, A}(X), \varphi_{0, A}(X^{ad})\in \mathrm{Im~}(\psi_{0, A})$$. We conclude that $$X\in T^1(A)$$. Thus $$\mathcal{M} \subseteq T^1(A)$$, which concludes the proof. ■ In the above theorem, the equality does not hold for a general flat $$A$$-algebra $$R$$. For example, assume that $$(L, h)$$ is a quadratic $$\mathbb{Z}_2$$-lattice of rank 1 and let $$R$$ be a ramified quadratic extension of $$\mathbb{Z}_2$$. Then $$R\otimes_AT^1(A)$$ is the ideal of $$R$$ generated by 2, whereas $$\mathrm{ker}(\bar{\varphi}_{0, R})\cap \mathrm{ker}(\bar{\varphi}_{0, R})^{ad}$$ is the maximal ideal of $$R$$ generated by a uniformizer. Thus if we define $$T^1(R)$$ to be $$\mathrm{ker}(\bar{\varphi}_{0, R})\cap \mathrm{ker}(\bar{\varphi}_{0, R})^{ad}$$ for any flat $$A$$-algebra $$R$$, then the functor $$T^1$$ may not be represented by a polynomial ring. We now claim that the functor $$T^1$$ is represented by a unique polynomial ring in some number of variables. More precisely, Lemma 3.7. The functor $$T^1$$ is represented by a flat $$A$$-algebra which is a polynomial ring over $$A$$ in $$n^2\cdot[K:F]$$ variables. □ Proof. Let $$R$$ be a flat $$A$$-algebra. Since $$T^1(R)=R\otimes_AT^1(A)$$ is a finitely generated free $$R$$-module, $$T^1$$ is represented by a polynomial ring in some number of variables over $$A$$. Thus the relative dimension of $$T^1$$, considered as an affine space, over $$\textit{Spec A}$$ is the same as the dimension of the generic fiber of $$T^1$$ over $$\textit{Spec F}$$, which is the same as the dimension of $$T^1(F)$$ as an $$F$$-vector space. We claim that $$T^1(F)~~(=F\otimes_AT^1(A))~~=T^0(F)$$. It is easy to show that $$F\otimes_AT^1(A)=\{X\in \mathrm{ker~}\bar{\varphi}_{0, F} : X^{ad} \in \mathrm{ker~}\bar{\varphi}_{0, F}\}$$. Since $$\psi_{0, F}$$ is surjective, The latter is the same as $$T^0(F)$$, whose dimension as an $$F$$-vector space is $$n^2\cdot [K:F]$$. The surjectivity of $$\psi_{0, F}$$ follows from smooothness of the generic fiber $$G$$. thus $$\psi_{0, F}$$ is not surjective if $$(L, q)$$ is a quadratic $$A$$-lattice and the characteristic of $$F$$ is $$2$$. However, this case is excluded along the article as mentioned in Sections 2.0.2, 2.0.4, and 2.2.1. ■ There is a natural morphism from $$T^1$$ to $$T^0$$ mapping $$X$$ to $$X$$ itself, where $$X\in T^1(R)$$ for a flat $$A$$-algebra $$R$$. This morphism is represented by a morphism of schemes and we denote it by $$\iota_0$$. We note that $$\iota_0$$ is an injection on flat $$A$$-algebras, but not an immersion of schemes. For example, if $$R$$ is a torsion $$A$$-algebra, then $$\iota_{0, R}$$ is no longer injective. Let $$\varphi_1$$ (resp. $$\psi_1$$) be the morphism from $$T^1$$ to $$H^0$$ induced by $$\varphi_0$$ (respecticvely $$\psi_0$$) composed with $$\iota_0$$, illustrated in the following commutative diagram of morphisms of schemes: $$ $$ 3.8 Construction of the functor $$\boldsymbol{T^{m+1}}$$ on the category of étale $$A$$-algebras In this section and the next, our aim is to recursively define functors $$T^{m+1}$$ from the category of flat $$A$$-algebras to the category of abelian groups and recursively define morphisms of functors $$\iota_m : T^{m+1} \rightarrow T^m$$, $$\varphi_m, \psi_m : T^m \rightarrow H^0$$, and $$\bar{\varphi}_m : T^m \rightarrow \mathrm{coker}(\psi_m)$$ for all $$m\geq 0$$. First we define them on the category of étale $$A$$-algebras, and then in Section 3.11, we extend them to the category of all flat $$A$$-algebras. We define the functor $$T^{m+1}$$, for all $$m \geq 0$$, from the category of étale $$A$$-algebras to the category of abelian groups as follows: Tm+1(R)=ker(φ¯m,R)∩ker(φ¯m,R)ad with the following morphisms: {ιm−1,R:Tm(R)⟶Tm−1(R), X↦X;φm,R,ψm,R:Tm(R)⟶H0(R), φm,R=φm−1,R∘ιm−1,R, ψm,R=ψm−1,R∘ιm−1,R;φ¯m,R:Tm(R)⟶H0(R)⟶coker(ψm).  The set $$\mathrm{ker}(\bar{\varphi}_{m, R})\cap \mathrm{ker}(\bar{\varphi}_{m, R})^{ad}$$ can be characterized as the set $$ \{X\in \mathrm{ker~}\overline{\varphi}_{m, R} : X^{ad} \in \mathrm{ker~}\overline{\varphi}_{m, R}\}$$ by using Proposition 3.3. For convenience, let $$T^{-m}=T^0$$ and let $$\varphi_{-m, R}=\varphi_{0, R}$$ and $$\psi_{-m, R}=\psi_{0, R}$$ for all $$m\geq 0$$. We will show that the above functors are well-defined in the following theorem. Theorem 3.9. Let $$R$$ be an étale $$A$$-algebra. Then for any integer $$m\geq 0$$, we have the followings: (1) The morphisms $$\iota_{m-1, R}, \varphi_{m, R},$$ and $$\psi_{m, R}$$ are well-defined and $$\psi_{m, R}$$ is a group homomorphism under addition; (2) $$X^{ad}\cdot Y\in T^m(R)$$ with $$X, Y\in T^m(R)$$; (3) $$\bar{\varphi}_{m, R}$$ is a group homomorphism under addition; (4) $$T^{m+1}(R)$$ is a group under addition. □ Proof. The statement (4) is a direct consequence of the statements (1)–(3). We prove this by induction. When $$m=0$$, the theorem follows from Proposition 3.3 and Section 3.4.3. Suppose that the theorem is true for all $$n\leq m-1$$, where $$m \geq 1$$, so that Tm(R)=ker(φ¯m−1,R)∩ker(φ¯m−1,R)ad is well-defined and a group under addition. Then the statement (1) is obvious. To prove the statement (2), let $$X, Y\in T^{m}(R)$$ and choose $$Z\in T^{m-1}(R)$$ such that $$\varphi_{m-1, R}(X^{ad})=\psi_{m-1, R}(Z)$$. To show that $$X^{ad}\cdot Y\in T^m(R)$$, we consider the following identity: φm−1,R(Xad⋅Y)=ψm−1,R(Yad⋅Z⋅Y). Since $$Z\cdot Y$$ and $$Y^{ad}\cdot Z\cdot Y \in T^{m-1}(R)$$ by hypothesis of induction (statement (2)), we conclude that $$\overline{\varphi}_{m-1, R}(X^{ad}\cdot Y)=0$$. Similarly, $$\overline{\varphi}_{m-1, R}(Y^{ad}\cdot X)=0$$ so $$X^{ad}\cdot Y\in T^m(R)$$. To show that $$\overline{\varphi}_{m, R}$$ is a group homomorphism, choose $$X, Y \in T^m(R)$$. Let us consider the following identity: φm,R(X+Y)=φm,R(X)+φm,R(Y)+ψm,R(Xad⋅Y). Since $$X^{ad}\cdot Y\in T^m(R)$$, $$\overline{\varphi}_{m, R}$$ is a group homomorphism. This completes the proof. ■ Corollary 3.10. Let $$R$$ be an étale $$A$$-algebra. Let $$X\in T^{m+1}(R)$$ and let $$Y\in T^{m}(R)$$. Then $$\iota_{m, R}(X)\cdot Y\in \mathrm{ker~}\overline{\varphi}_{m, R}$$. □ Proof. Let $$\varphi_{m, R}\left( \iota_{m, R}(X)\right)=\psi_{m, R}(Z)$$ for some $$Z\in T^m(R)$$. Then $$\varphi_{m, R}(\iota_{m, R}(X)\cdot Y)=\psi_{m, R}(Y^{ad}\cdot Z\cdot Y)$$ and $$Y^{ad}\cdot Z\cdot Y\in T^m(R)$$ by the above theorem. This completes the proof. ■ 3.11 Constructions of the functors $$\boldsymbol{T^{m+1}}$$ and $$\boldsymbol{\widetilde{T}}$$ on the category of flat $$\boldsymbol{A}$$-algebras We extend the functor $$T^{m+1}$$ to a functor from the category of commutative flat $$A$$-algebras to the category of abelian groups as follows: Tm+1(R)={ker(φ¯m,A)∩ker(φ¯m,A)adif R=A;R⊗ATm+1(A)if R is a flat A-algebra.  By similar arguments used in Subsection 3.5 and Theorem 3.6 and Lemma 3.7, combined with induction on $$m$$, we can show that (1) the set $$\mathrm{ker}(\bar{\varphi}_{m, A})\cap \mathrm{ker}(\bar{\varphi}_{m, A})^{ad}$$ is an $$A$$-module so that $$T^{m+1}$$ is well-defined; (2) $$\iota_{m, R} : T^{m+1}(R) \rightarrow T^{m}(R)$$ is an injective $$R$$-module homomorphism for a flat $$A$$-algebra $$R$$; (3) for an étale $$A$$-algebra $$R$$, R⊗ATm+1(A)(=Tm+1(R))=ker(φ¯m,R)∩ker(φ¯m,R)ad, as $$R$$-submodules of $$T^{m}(R)$$; (4) $$T^{m+1}$$ is represented by a flat $$A$$-algebra which is a polynomial ring over $$A$$ in $$n^2\cdot[K:F]$$ variables; (5) the morphisms $$\iota_{m-1, R}, \psi_{m, R}, \bar{\varphi}_{m, R}$$, and $$\varphi_{m, R}$$, for a flat $$A$$-algebra $$R$$, are represented by morphisms of $$A$$-schemes. First three are morphisms of group schemes under addition, whereas the last is not. Note that $$T^m(F)=T^0(F)$$ for all $$m$$. Let us consider the following sequence induced by $$\iota_m$$: T0(R)⊇T1(R)⊇⋯⊇Tm(R)⊇⋯ for a flat $$A$$-algebra $$R$$. We emphasize that all ranks of $$T^i(R)$$, as finitely generated free $$R$$-modules, are the same, which is $$n^2\cdot[K:F]$$. Indeed, there is a suitable integer $$\eta$$ which makes the above sequence stabilize and this is proved in the following theorem. Theorem 3.12. There exists an integer $$\eta ~(\geq 0)$$ such that $$T^m=T^\eta$$ for all $$m \geq \eta$$. □ Proof. Let $$M=T^0(A)/T^1(A)$$. It is then clear that $$M$$ is a torsion $$A$$-module since the rank of $$T^0(A)\otimes_AF$$ is the same as that of $$T^1(A)\otimes_AF$$. Let $$l$$ be the smallest non-negative integer such that $$\pi^l\cdot M=0$$. Let T′(R)=π2lT0(R) for a flat $$A$$-algebra $$R$$. Then $$T'$$ is represented by a flat $$A$$-algebra which is a polynomial ring in some number of variables over $$A$$. Let $$\varphi'$$ and $$\psi'$$ be morphisms from $$T'$$ to $$H^0$$ induced from $$\varphi_0$$ and $$\psi_0$$, respectively. We choose an element $$\pi^{2l}\cdot X\in T'(R)$$ for an étale $$A$$-algebra $$R$$. Since $$\pi^l\cdot X\in T^1(R)$$, we have that $$\varphi_{0, R}(\pi^l\cdot X)=\psi_{0, R}(Y)$$ for a certain $$Y\in T^0(R)$$. Then, φR′(π2l⋅X)=π2l⋅φR′(πl⋅X)=π2l⋅ψ0,R(Y)=ψR′(π2l⋅Y) with $$\pi^{2l}\cdot Y\in T'(R)$$. Here, the first equality follows from the fact that $$\varphi'_R(aX)=a^2\varphi'_R(X)$$ for $$a\in A$$ and the last equality follows from the fact that $$\psi'_R$$ is $$A$$-linear. Therefore, $$T'(R)\subset T^m(R)$$ for every integer $$m$$, where $$R$$ is an étale $$A$$-algebra. On the other hand, if $$T^l=T^{l+1}$$ for certain non-negative integer $$l$$, then it is clear that $$T^l=T^{l'}$$, for all $$l'\geq l$$. This fact yields that a sequence of the form $$T^l(R)=T^{l+1}(R)\nsupseteq T^{l+2}(R)$$, for a flat $$A$$-algebra $$R$$, cannot happen. Thus we conclude the existence of an integer $$\eta ~(\geq 0)$$ stabilizing the sequence explained just before this theorem. ■ Let $$\alpha$$ be the smallest non-negative integer satisfying that $$T^m=T^{\alpha}$$ for all $$m \geq \alpha$$. We finally define the functor T~:=Tα, equipped with two morphisms $$\varphi_{\alpha}$$ and $$\psi_{\alpha}$$ mapping to $$H^0$$. We note that Theorem 3.12 and its proof only assert stability of a sequence of functors $$T^m$$. In order to find $$\alpha$$, one can construct $$T^{m}(A)$$ explicitly starting from $$m=0$$. Then $$\alpha$$ is the first integer $$m$$ such that $$T^{m}(A)=T^{m+1}(A)$$ (equivalently, the map $$\overline{\varphi}_{m, A} : T^m(A) \rightarrow H^0(A)/\mathrm{Im~}\psi_{m, A}$$ is zero). Indeed, we expect that $$\alpha$$ is at most $$e'+1$$, where the ramification index $$e=2e'$$ or $$e=2e'-1$$, as this is true in the unramified case (cf. Remark 3.23) and for the examples covered in Section 4.1 and Examples 4.13–4.16. 3.13 Construction of $$\boldsymbol{\widetilde{H}}$$ Recall that $$\psi_{\alpha, R} : \widetilde{T}(R) \longrightarrow H^0(R)$$ is $$R$$-linear for a flat $$A$$-algebra $$R$$. Define the functor $$\widetilde{H}$$ from the category of commutative flat $$A$$-algebras to the category of abelian groups as follows: H~(R)=Im ψα,R. Theorem 3.14. The functor $$\widetilde{H}$$ is represented by a flat $$A$$-algebra which is a polynomial ring in some number of variables over $$A$$ of $$n^2\cdot[K:F]-\mathrm{dim~}G$$ variables and the map $$\psi_{\alpha, R} : \widetilde{T}(R) \rightarrow \widetilde{H}(R)$$ is represented by a morphism of schemes, denoted by $$\widetilde{\psi}$$. In addition, the map $$\widetilde{\psi_{R_{\kappa}}} : \widetilde{T}(R_{\kappa}) \rightarrow \widetilde{H}(R_{\kappa})$$ is surjective for any $$\kappa$$-algebra $$R_{\kappa}$$. □ Proof. Since $$\widetilde{T}(R)=R\otimes_A\widetilde{T}(A)$$ and $$\psi_{\alpha, R}$$ is $$R$$-linear for a flat $$A$$-algebra $$R$$, we have that $$\widetilde{H}(R)=R\otimes_A\widetilde{H}(A)$$. Therefore, $$\widetilde{H}$$ is represented by a polynomial ring in some number of variables over $$A$$. Since $$\widetilde{H}(F)=H^0(F)$$, the dimension of $$\widetilde{H}(F)$$ as an $$F$$-vector space is $$n^2\cdot [K:F]-\mathrm{dim~}G$$, which is the same as the relative dimension of $$\widetilde{H}$$ over $$\textit{Spec A}$$. The representability of $$\psi_{\alpha, R}$$ is obvious. To show that $$\widetilde{\psi_{R_{\kappa}}} : \widetilde{T}(R_{\kappa}) \rightarrow \widetilde{H}(R_{\kappa})$$ is surjective for a $$\kappa$$-algebra $$R_{\kappa}$$, we choose a flat $$A$$-algebra $$R$$ such that $$R\otimes_A\kappa=R_{\kappa}$$. Then we have the following commutative diagram: $$ $$ Since $$\widetilde{T}$$ and $$\widetilde{H}$$ are represented by affine spaces, two vertical maps are surjective. One can also show surjectivity of two vertical maps by using Hensel’s lemma since $$\widetilde{T}$$ and $$\widetilde{H}$$ are smooth. In addition, $$\widetilde{\psi_{R}}$$ is surjective by the definition of $$\widetilde{H}$$. Therefore, $$\widetilde{\psi_{R_{\kappa}}}$$ is surjective as well. ■ Theorem 3.15. Let $$R$$ be a flat $$A$$-algebra. Then the image of $$\widetilde{T}(R)$$, under the map $$\varphi_{\alpha, R}$$, is contained in $$\widetilde{H}(R)$$. Therefore, the morphism $$\varphi_{\alpha} : \widetilde{T} \rightarrow H^0$$ factors through $$\widetilde{H}$$. □ Proof. Let $$R$$ be an étale $$A$$-algebra. Suppose that there is $$X\in T^{\alpha}(R)=\widetilde{T}(R)$$ such that $$\varphi_{\alpha, R}(X) \notin \widetilde{H}(R)$$. Then $$\overline{\varphi}_{\alpha, R}(X)\neq 0$$ and so $$X$$ is not contained in $$T^{\alpha+1}(R)$$. This contradicts the definition of $$\alpha$$. Thus $$\varphi_{\alpha, R}(X) \in \widetilde{H}(R)$$ for any $$X\in \widetilde{T}(R)$$. Let $$R$$ be a flat $$A$$-algebra. Choose $$X\in T^{\alpha}(R)$$. We can write $$X=\sum_ir_i\cdot X_i$$ with $$r_i\in R$$ and $$X_i\in T^{\alpha}(A)=\widetilde{T}(A)$$. Then $$\varphi_{\alpha, R}(X)=\sum_ir_i^2\cdot \varphi_{\alpha, A}(X_i)+\sum_{i1$$ (we assume this to exclude known cases) be the ramification index. Let $$(L, h)$$ be a quadratic lattice (i.e., $$K=F$$ and $$\epsilon=1$$). We use the symbol $$A(a, b)$$ to denote the $$A$$-lattice $$A\cdot e_1+A\cdot e_2$$ with the symmetric bilinear form having Gram matrix $$\begin{pmatrix} a&1\\ 1&b \end{pmatrix}$$ . We denote by $$(t)$$ the $$A$$-lattice of rank 1 equipped with the symmetric bilinear form having Gram matrix $$(t)$$. Suppose that L=⨁iAi(0,0)⊕A(πs,rπ2e−s)⊕(t) of rank $$2m+3$$ with $$r\in A$$ and $$t\equiv 1 \ \ \mathrm{mod}\ \ \pi$$. Here, $$s\leq e$$, $$s$$ is odd if $$s1$$ by the following facts extracted from Examples 93:9 and 93:18 of [13]: (1) (cf. 93:18.(v) of [13]) If the rank of $$L$$ is at least $$5$$, then L≅A(0,0)⊕⋯. (2) (cf. 93:18.(ii), (iv) of [13]) If the rank of $$L$$ is $$3$$, then L≅A(0,0)⊕(t) orL≅A(b,4rb−1)⊕(t). Here, $$b, r\in A$$ such that the exponential order of $$b$$ is odd and at most $$e$$, and $$t\equiv 1 \ \ \mathrm{mod}\ \ \pi$$. (3) (cf. 93:9 of [13]) A(2α,0)≅A(2α+2β,0) for any $$\alpha, \beta\in A$$. If we choose $$\beta=-\alpha$$, then this implies that $$A(2\alpha,0)\cong A(0,0)$$. 4.1 Constructions of $$\widetilde{T}$$ and $$\widetilde{H}$$ We first construct the smooth integral model $$\underline{G}$$ associated to $$L$$. As mentioned in Section 3.24, a key ingredient of the construction of $$\underline{G}$$ is the explicit description of $$\widetilde{T}(A)$$. Basically, the procedure to find $$\widetilde{T}(A)$$ gives certain congruence conditions on the entries of an element of $$T^0(A)$$. A matrix form of an element of $$\widetilde{T}(A)$$ is described below. For $$X\in \widetilde{T}(A)$$, X={(x0y0π[(e−s)/2]z0πe′w0π[(e−s)/2]x1π[(e−s)/2]y1πe−sz1π[e−(s−1)/2]w1x2y2π[(e−s)/2]z2πe′w2πe′x3πe′y3π[e−(s−1)/2]z32w3)if s0.  □ Remark 4.17. Let $$J$$ and $$J'$$ be Jordan components of a quadratic lattice $$L$$. Assume that $$J$$ (respectively $$J'$$) is $$\pi^n$$-modular (respectively $$\pi^{n+i}$$-modular). Then we say that the distance between $$J$$ and $$J'$$ is $$|i|$$. We keep using the notation $$L^a$$ from the above example. For the lattice $$L^0$$ (respectively $$L^a$$ with $$a>0$$), the component group is determined by two equations $$x_0+x_0^2+z_0^2=0, w_0+w_0^2+y_0^2=0$$ (respectively $$x_0+x_0^2=0, w_0+w_0^2=0$$) which are $$(1\times 1)$$ and $$(2\times 2)$$ blocks of a formal matrix equation. The difference between these two lattices $$L^0$$ and $$L^a$$ with $$a>0$$ is the distance between two Jordan components. If the distance between Jordan components is strictly greater than $$2e$$, then two terms $$z_0^2$$ and $$y_0^2$$ are ignored in the above equations. In other words, two Jordan components with distance greater than $$2e$$ do not interrupt each other in a sense of identifying the connected components. Indeed, the $$(1\times 1)$$ (respectively $$(2\times 2)$$) block of $$\underline{G}$$ associated to $$L^a$$ with $$a>0$$ is the same as $$\underline{G}$$ associated to the quadratic lattice $$(1)$$ (respectively $$(4\pi^a)$$) of rank $$1$$. The congruence conditions of the remained blocks of $$\underline{G}$$ associated to $$L^a$$ are the same as those of $$T^0$$. It seems that this phenomenon occurs in the general case with any ramification index. For example, consider the lattice $$M=\left((1)\oplus (\pi)\right)\oplus \left((8)\oplus (16)\right)\oplus \left((64\pi)\oplus (128)\right)$$ with $$e=2$$. Here the distance between $$(\pi)$$ and $$(8)$$, and between $$(16)$$ and $$(64\pi)$$ is $$5>2e$$. Then the block, corresponding to $$(1)\oplus (\pi)$$ (respectively $$(8)\oplus (16)$$, resp. $$(64\pi)\oplus (128)$$), of $$\underline{G}$$ associated to $$M$$ is the same as $$\underline{G}$$ associated to $$(1)\oplus (\pi)$$ (respectively $$(8)\oplus (16)$$, respectively $$(64\pi)\oplus (128)$$), and the congruence conditions of the remained blocks are the same as those of $$T^0$$ associated to $$M$$. We can then enumerate all equations defining $$\tilde{G}$$ (the morphism $$\varphi$$ is trivial) and show that G~≅A15×(Z/2Z)β as varieties over $$\kappa$$. The value of $$\beta$$ is described as follows: β={2+2+2=6ifA/Z2satisfiesCase1;2+1+2=5ifA/Z2satisfiesCase2.  Here, $$\mathbb{A}^{3}\times (\mathbb{Z}/2\mathbb{Z})^{\beta}$$ is determined by three diagonal blocks corresponding to $$(1)\oplus (\pi)$$, $$(8)\oplus (16)$$, and $$(64\pi)\oplus (128)$$ (which can be read off from Examples 4.14–4.15), and $$\mathbb{A}^{12}$$ is determined by the remained non-diagonal blocks (which is easy to identify). Thus the value of $$\beta$$ associated to $$M$$ is the same as the sum of $$\beta$$’s associated to $$(1)\oplus (\pi)$$, $$(8)\oplus (16)$$, and $$(64\pi)\oplus (128)$$. The case when $$A/\mathbb{Z}_2$$ is unramified in [2] (and the case of hermitian lattices in [3] and [4]) also provides evidence of the expectation explained in this remark. □ For other lattices such as hermitian lattices, one can still construct a morphism of algebraic groups over $$\kappa$$ as in Proposition 4.9 φ=∏iφi:G~⟶∏i(classical group). One method to find classical groups in the above is the following: Each Jordan component of a Jordan splitting of a given lattice $$(L, h)$$ involves a hyperbolic part (possibly empty as Examples 4.14–4.16). Let $$g$$ be an element of $$\tilde{M}$$ (which is the special fiber of $$\underline{M}^{\ast}$$) and $$\delta$$ be a matrix associated to a given form $$h$$. Then we consider the formal matrix equation $$\sigma({}^tg)\delta g=\delta$$ as in Theorem 4.4 and Example 4.13. Now we observe the block of the above formal matrix equation corresponding to the hyperbolic part of each Jordan component. This block should not be difficult to figure out and give information about the form of classical groups. The articles [3] and [4] give an explicit construction of such morphism for ramified hermitian lattices defined over an unramified extension of $$\mathbb{Z}_2$$. In this case, such classical groups are orthogonal groups or symplectic groups. We conjecture that, under these assumptions, Conjecture 4.10 still holds. As in Proposition 4.9, we expect that the morphism $$\varphi$$ is constructed by choosing certain sublattices of $$L$$ and using this, one can show that the formation of $$\varphi$$ is compatible with finite unramified extension on $$A$$. The conjecture can also be proved by using the similar argument explained in Theorem 4.4 and Remark 4.12. In addition, one can see that the local density is given by a rational function with respect to the cardinality of the residue field, as one varies a base ring $$A$$ to a finite unramified extension $$A'$$ (cf. Remark 4.11). Funding This work was supported by JSPS KAKENHI Grant No. 16F16316. Acknowledgments The author had a motivation and an initial idea by hearing the discussion of Andrew Fiori and Gabriele Nebe, and by discussing with Brian Conrad during the AIM workshop “Algorithms for lattices and algebraic automorphic forms”, May 2013. The author would like to show deep appreciation to Wai Kiu Chan, Brian Conrad, Andrew Fiori, Abhinav Kumar, Gabriele Nebe, Rudolf Scharlau, Rainer Schulze-Pillot, and John Voight for valuable comments and discussions during the workshop. The author also thanks the organizers of the workshop and the AIM for inviting him and for their hospitality. Section 3, which is a main section of this article, was inspired by a discussion with Radhika Ganapathy. This article was improved by discussing with Wee Teck Gan during the author’s visit to 2013 Pan Asian Number Theory (PANT) conference, Hanoi, July 2013, Vietnam. He thanks the organizers of 2013 PANT conference and VIASM for inviting him and for their hospitality as well. The author would like to express his deep gratitude to Juan Marcos Cerviño, Wai Kiu Chan, Brian Conrad, Wee Teck Gan, Radhika Ganapathy, Benedict Gross, Manish Mishra, Gopal Prasad, Rudolf Scharlau, Rainer Schulze-Pillot, Olivier Taïbi, Sandeep Varma, Jiu-Kang Yu, and his wife Bogume Jang for their encouragement and support on various aspects of the article and the author’s mathematical life, and to the referee for many valuable comments and corrections. Present address: Department of Mathematics, Kyoto University, Kitashirakawa Oiwake-cho, Sakyo-ku, Kyoto 606-8502, Japan Communicated by Prof. Wee Teck Gan References [1] Bosch S. , Lütkebohmert W. and Raynaud. M. “Néron Models.” Ergebnisse der Mathematik und ihrer Grenzgebiete (3) 21 . Berlin : Springer , 1990 . Google Scholar CrossRef Search ADS [2] Cho S. “Group Schemes and Local Densities of Quadratic Lattices in Residue Characteristic 2.” Compositio Mathematica 151 , no. 5 ( 2015 ): 793 – 827 . Google Scholar CrossRef Search ADS [3] Cho S. “Group Schemes and Local Densities of Ramified Hermitian Lattices in Residue Characteristic 2 Part I.” Algebra & Number Theory 10 , no. 3 ( 2016 ): 451 – 532 . Google Scholar CrossRef Search ADS [4] Cho S. “Group Schemes and Local Densities of Ramified Hermitian Lattices in Residue Characteristic 2 Part II.” Preprint , available at https://sites.google.com/site/sungmuncho12/. [5] Conway J. H. and Sloane. J. A. “Low-dimensional lattices, IV: The mass formula.” Proceedings of the Royal Society of London. Series A 419 ( 1988 ): 259 – 86 . Google Scholar CrossRef Search ADS [6] Gan W. T. and Yu. J.-K. “Group schemes and local densities.” Duke Mathematical Journal 105 ( 2000 ): 497 – 524 . Google Scholar CrossRef Search ADS [7] Hartshorne R. Algebraic Geometry. Grad. Texts in Math . 52 . New York : Springer , 1977 . Google Scholar CrossRef Search ADS [8] Hironaka Y. and Sato. F. “Local densities of representations of quadratic forms over $$p$$-adic integers (the non-dyadic case).” Jounal of Number Theory 83 ( 2000 ): 106 – 36 . Google Scholar CrossRef Search ADS [9] Jacobowitz R. “Hermitian forms over local fields.” American Journal of Mathematics 84 ( 1962 ): 441 – 65 . Google Scholar CrossRef Search ADS [10] Kitaoka Y. Arithmetic of Quadratic Forms. Cambridge Tracts in Math . 106 . Cambridge : Cambridge Univ. Press , 1993 . Google Scholar CrossRef Search ADS [11] Knus M.-A. Quadratic and Hermitian Forms Over Rings. Berlin : Springer , 1991 Google Scholar CrossRef Search ADS [12] O’Meara O. T. “Quadratic forms over local fields.” American Journal of Mathematics 77 ( 1955 ): 87 – 116 . Google Scholar CrossRef Search ADS [13] O’Meara O. T. Introduction to Quadratic Forms , reprint of 1973 ed. Classics Math . Berlin : Springer , 2000 . © 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 - A Uniform Construction of Smooth Integral Models and a Conjectural Recipe for Computing Local Densities JO - International Mathematics Research Notices DO - 10.1093/imrn/rnw340 DA - 2017-02-26 UR - https://www.deepdyve.com/lp/oxford-university-press/a-uniform-construction-of-smooth-integral-models-and-a-conjectural-m3vEzrZrzg SP - 1 EP - 3907 VL - Advance Article IS - 12 DP - DeepDyve ER -