1 Introduction
1.1
The main objective of this paper is to construct regular integral models for some Shimura varieties over places of bad reduction. Results of this type have already been obtained by Rapoport-Zink [Reference Rapoport and Zink24], Pappas [Reference Pappas16] and He-Pappas-Rapoport [Reference He, Pappas and Rapoport12] in specific instances. Here, we consider Shimura varieties associated to unitary similitude groups of signature
$(r,s)$
over an imaginary quadratic field K. These Shimura varieties are moduli spaces of abelian varieties with polarization, endomorphisms and level structure (Shimura varieties of PEL type). Shimura varieties have canonical models over the ‘reflex’ number field E, and in the cases we consider, the reflex field is the field of rational numbers
$\mathbb {Q}$
if
$r = s$
and
$E = K$
otherwise [Reference Pappas15, §3].
Constructing such well-behaved integral models is an interesting and challenging problem, with several applications to number theory. The behavior of these depends very much on the ‘level subgroup’. Here, the level subgroup is the stabilizer of a
$\pi $
-modular lattice in the hermitian space. (This is a lattice which is equal to its dual times a uniformizer). For this level subgroup, and the signatures
$(n-s,s)$
with n even and
$s \leq 3$
, we show that there is a good (semi-stable) model.
By using work of Rapoport-Zink [Reference Rapoport and Zink24] and Pappas [Reference Pappas15], we first construct p-adic integral models, which have simple and explicit moduli descriptions but are not always flat. These models are étale locally around each point isomorphic to certain simpler schemes the naive local models. Inspired by the work of Pappas-Rapoport [Reference Pappas and Rapoport17] and Krämer [Reference Krämer14], we consider a variation of the above moduli problem where we add in the moduli problem an additional subspace in the deRham filtration
$ \mathrm {Fil}^0 (A) \subset H_{dR}^1(A)$
of the universal abelian variety A, which satisfies certain conditions. This is essentially an instance of the notion of a ‘linear modification’ introduced in [Reference Pappas15]. Then for the signatures
$(n-s,s)$
, with
$s \leq 3$
, we calculate the flat closure of these models, and we show that they are smooth when
$s=1$
and have semi-stable reduction when
$s=2$
or
$s=3$
(i.e., they are regular and the irreducible components of the special fiber are smooth divisors crossing normally). Moreover, we want to mention that we obtain a moduli description of this flat closure. We anticipate that our construction will have applications to the study of arithmetic intersections of special cycles and Kudla’s program. (See, for example, [Reference Zhang29], [Reference Bruinier, Howard, Kudla, Rapoport and Yang6] and [Reference He, Li, Shi and Yang11], for some works in this direction.)
1.2
Let us introduce some notation first. Recall that K is an imaginary quadratic field and we fix an embedding
$\varepsilon : K\rightarrow \mathbb {C}$
. Let W be a n-dimensional K-vector space, equipped with a nondegenerate hermitian form
$\phi $
. Consider the group
$G = \mathrm {GU}_n$
of unitary similitudes for
$(W,\phi )$
of dimension
$n\geq 3$
over K. We fix a conjugacy class of homomorphisms
$h: \mathrm {Res}_{\mathbb {C}/\mathbb {R}}\mathbb {G}_{m,\mathbb {C}}\rightarrow \mathrm {GU}_n$
corresponding to a Shimura datum
$(G,X_h)$
of signature
$(r,s)$
. Set
$X = X_h$
. The pair
$(G,X)$
gives rise to a Shimura variety
$Sh(G,X)$
over the reflex field E (see [Reference Pappas and Rapoport18, §1.1] for more details). Let p be an odd prime number which ramifies in K. Set
$K_1=K\otimes _{\mathbb {Q}} \mathbb {Q}_p$
with a uniformizer
$\pi $
, and
$V=W\otimes _{\mathbb {Q}} \mathbb {Q}_p$
. We assume that the hermitian form
$\phi $
is split on V (i.e., there is a basis
$e_1, \dots , e_n$
such that
$\phi (e_i, e_{n+1-j})=\delta _{ij}$
for
$1\leq i, j\leq n$
). We denote by

the standard lattices in V. We can complete this into a self dual lattice chain by setting
$\Lambda _{i+kn}:=\pi ^{-k}\Lambda _i$
(see §2.1).
By [Reference Pappas and Rapoport18], there are 3 different types of the special maximal parahoric subgroups of
$\mathrm {GU}_n$
depending on the parity of n. More precisely, consider the stabilizer subgroup

where I is a nonempty subset of
$\{0,\dots , \lfloor n/2\rfloor \}$
. When
$n=2m+1$
is odd, the special maximal parahoric subgroups are conjugate to the stabilizer subgroup
$P_{\{0\}}$
or
$P_{\{m\}}$
. When
$n=2m$
is even, the special maximal parahoric subgroups are conjugate to the stabilizer subgroup
$P_{\{m\}}$
. In this article, we consider the special maximal parahoric subgroup
$P_{\{m\}}$
when
$n=2m$
is even. We intend to take up the case
$n=2m+1, I=\{m\}$
in a subsequent work.
To explain our results, we assume
$n=2m$
, and
$(r,s)=(n-1,1)$
or
$(n-2,2)$
or
$(n-3,3)$
. Denote by
$P_{\{m\}}$
the stabilizer of
$\Lambda _m$
in
$G(\mathbb {Q}_p)$
. We let
$\mathcal {L}$
be the self-dual multichain consisting of lattices
$\{\Lambda _j\}_{j\in n\mathbb {Z} \pm m}$
. Here,
$\mathcal {G} = \underline {\mathrm {Aut}}(\mathcal {L})$
is the (smooth) group scheme over
$\mathbb {Z}_p$
with
$P_{\{m\}} = \mathcal {G}(\mathbb {Z}_p)$
the subgroup of
$G(\mathbb {Q}_p)$
fixing the lattice chain
$\mathcal {L}$
.
Choose also a sufficiently small compact open subgroup
$K^p$
of the prime-to-p finite adelic points
$G({\mathbb A}_{f}^p)$
of G and set
$\mathbf {K}=K^pP_{\{m\}}$
. The Shimura variety
$\mathrm { Sh}_{\mathbf {K}}(G, X)$
with complex points

is of PEL type and has a canonical model over the reflex field E. We set
$\mathcal {O} = O_{E_v}$
, where v the unique prime ideal of E above
$(p)$
.
We consider the moduli functor
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
over
$\mathrm {Spec \, } \mathcal {O} $
given in [Reference Rapoport and Zink24, Definition 6.9]:
A point of
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
with values in the
$\mathcal {O} $
-scheme S is the isomorphism class of the following set of data
$(A,\iota ,\bar {\lambda }, \bar {\eta })$
:
-
(1) An object
$(A,\iota )$ , where A is an abelian scheme with relative dimension n over S (terminology of [Reference Rapoport and Zink24]), compatibly endowed with an action of
${\mathcal O}$ :
$$\begin{align*}\iota: {\mathcal O} \rightarrow \text{End} \,A \otimes \mathbb{Z}_p.\end{align*}$$
-
(2) A
$\mathbb {Q}$ -homogeneous principal polarization
$\bar {\lambda }$ of the
$\mathcal {L}$ -set A.
-
(3) A
$K^p$ -level structure
$$\begin{align*}\bar{\eta} : H_1 (A, {\mathbb A}_{f}^p) \simeq W \otimes {\mathbb A}_{f}^p \, \text{ mod} \, K^p \end{align*}$$
$({\mathbb A}_{f}^p)^{\times }$ (see loc. cit. for details). The set A should satisfy the determinant condition (i) of loc. cit. which depends on
$(r,s)$ .
For the definitions of the terms employed here, we refer to loc.cit., 6.3–6.8 and [Reference Pappas15, §3]. The functor
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
is representable by a quasi-projective scheme over
$\mathcal {O}$
. Since the Hasse principle is satisfied for the unitary group, we can see as in loc. cit. that there is a natural isomorphism

The moduli scheme
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
is connected to the naive local model
$\mathrm {M}^{\mathrm {naive}}$
(see §3 for the explicit definition). As is explained in [Reference Rapoport and Zink24] and [Reference Pappas15], the naive local model is connected to the moduli scheme
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
via the local model diagram

where the morphism
$\psi _1$
is a
$\mathcal {G}$
-torsor and
$\psi _2$
is a smooth and
$\mathcal {G}$
-equivariant morphism. In particular, since
${\mathcal G}$
is smooth, the above imply that
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}} $
is étale locally isomorphic to
$\mathrm {M}^{\mathrm {naive}}$
. From [Reference Pappas, Rapoport and Smithling20, Remark 2.6.10], we have that the naive local model is never flat, and by the above, the same is true for
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}} $
. Denote by
$ \mathcal {A}^{\mathrm {flat}}_{\mathbf {K}} $
the flat closure of
$\mathrm {Sh}_{\mathbf {K}}(G, X)\otimes _{E} E_v$
in
$ \mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
. As in [Reference Pappas and Rapoport18], there is a relatively representable smooth morphism of relative dimension
$\mathrm {dim} (G)$
,

where the local model
$\mathrm {M}^{\mathrm {loc}} $
is defined as the flat closure of
$ \mathrm {M}^{\mathrm {naive}} \otimes _{\mathcal {O}} E_v$
in
$\mathrm {M}^{\mathrm {naive}}$
. This of course implies that
$\mathcal {A}^{\mathrm {flat}}_{\mathbf {K}}$
is étale locally isomorphic to the local model
$\mathrm {M}^{\mathrm {loc}}$
.
We now consider a variation of the moduli of abelian schemes
$\mathcal {A}_{\mathbf {K}}$
where we add in the moduli problem an additional subspace in the Hodge filtration
$ \mathrm {Fil}^0 (A) \subset H_{dR}^1(A)$
of the universal abelian variety A (see [Reference Haines7, §6.3] for more details) with certain conditions to imitate the definition of the naive splitting model
$\mathcal {M}$
(we refer to §3 for the explicit definition). There is a projective forgetful morphism

which induces an isomorphism over the generic fibers (see §6). Moreover,
$\mathcal {A}_{\mathbf {K}}$
has the same étale local structure as
$\mathcal {M}$
; it is a ‘linear modification’ of
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}\otimes _{\mathcal {O}} O_{K_1}$
in the sense of [Reference Pappas15, §2] (see also [Reference Pappas and Rapoport17, §15]). Note that there is also a corresponding projective forgetful morphism

which induces an isomorphism over the generic fibers (see §3). In §5, we show that
$\mathcal {M}$
is not flat for any signature
$(r,s)$
, and the same is true for
$ \mathcal {A}_{\mathbf {K}}$
. In fact,
$\mathcal {M}$
is not topologically flat since there exists a point which cannot lift to characteristic zero (see Proposition 5.1). This type of phenomenon is similar to what already has been observed for this level subgroup for the wedge local models in [Reference Pappas and Rapoport18, Remark 5.3] (see §3 for the definition of wedge local models). Consider the closed subscheme
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}} \subset \mathcal {A}_{\mathbf {K}} $
defined as
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}} :=\tau _1^{-1} (\mathcal {A}^{\mathrm {flat}}_{\mathbf {K}}) $
. It is a linear modification of
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}\otimes _{\mathcal {O}} O_{K_1}$
.
1.3
One of the main results of the present paper is the following theorem.
Theorem 1.1. Assume that
$ (r,s) = (n-1,1)$
or
$(n-2,2)$
or
$(n-3,3)$
. For every
$K^p$
as above, there is a scheme
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$
, flat over
$\mathrm {Spec \, } (O_{K_1})$
, with

and which supports a local model diagram

such that:
-
a)
$\pi ^{\mathrm {reg}}_{\mathbf {K}}$ is a
$\mathcal {G}$ -torsor for the parahoric group scheme
${\mathcal G}$ that corresponds to
$P_{\{m\}}$ .
-
b)
$q^{\mathrm {reg}}_{\mathbf {K}}$ is smooth and
${\mathcal G}$ -equivariant.
-
c) When
$(r,s) = (n-1,1)$ ,
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$ is a smooth scheme.
-
c’) When
$(r,s) = (n-2,2)$ or
$(r,s) =(n-3,3)$ ,
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$ has semi-stable reduction. In particular,
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$ is regular and has special fiber which is a reduced divisor with normal crossings.
In the above, the splitting model
$\mathrm {M}^{\mathrm {spl}}$
is defined as
$ \mathrm {M}^{\mathrm {spl}}:=\tau ^{-1}(\mathrm {M}^{\mathrm {loc}})$
. Since every point of
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}} $
has an étale neighborhood which is also étale over
$\mathrm {M}^{\mathrm {spl}}$
, it is enough to show that
$\mathrm {M}^{\mathrm {spl}}$
has the above nice properties. To show this, we explicitly calculate an affine chart
$\mathcal {U}$
of
$\mathrm {M}^{\mathrm {spl}}$
in a neighbourhood of
$ \tau ^{-1}(*)$
, where
$*$
is a point from the unique closed
$\mathcal {G}$
-orbit supported in the special fiber of
$\mathrm {M}^{\mathrm {loc}}$
(see [Reference Pappas15, §4] for more details). Note that the unique closed
$\mathcal {G}$
-orbit depends on the parity of s (see §3). We treat these two cases (i.e., when s is even and odd) separately in §4.1 and §4.2.
Moreover, we give a moduli-theoretic description of
$\mathrm {M}^{\mathrm {spl}}$
and so by the above for
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$
. Inspired by the work of Pappas-Rapoport [Reference Pappas and Rapoport18], we define the closed subscheme
$ \mathcal {M}^{\mathrm {spin}} \subset \mathcal {M}$
by adding the ‘spin condition’ (see §7) and we show the following.
Theorem 1.2. For the signatures
$(n-s,s)$
with
$s\leq 3$
, we have
$ \mathcal {M}^{\mathrm {spin}}=\mathrm {M}^{\mathrm {spl}}$
.
Here, we note that at the level of local models, a moduli description of
$\mathrm {M}^{\mathrm {loc}}$
for the signature
$(n-1,1)$
can be obtained by adding the ‘wedge and spin condition’ to the moduli problem of
$\mathrm {M}^{\mathrm {naive}}$
(see [Reference Rapoport, Smithling and Zhang22]). For higher signatures, it has been conjectured that we can get the moduli description of
$\mathrm {M}^{\mathrm {loc}}$
by adding the ‘wedge and spin conditions’ to
$\mathrm {M}^{\mathrm {naive}}$
; see the survey paper [Reference Pappas, Rapoport and Smithling20] for more details.
Before we move on to the description of
$\mathcal {U}$
, we want to give a brief historical outline about the splitting models. Splitting models were first described for associated Shimura varieties of type A and type C by Pappas-Rapoport [Reference Pappas and Rapoport17]. For unitary groups, Krämer [Reference Krämer14] first showed that the splitting model has semi-stable reduction for the signature
$(r,s)=(n-1,1)$
when
$n=2m+1, I=\{0\}$
. When the signature is
$(r,s)=(n-2,2)$
, the first author [Reference Zachos28] showed that the corresponding splitting model
$\mathcal {M}$
does not have semi-stable reduction. Then, he resolved the singularities by giving an explicit blow-up of
$\mathcal {M}$
along a smooth divisor. For more information on the geometry of the special fiber of
$\mathcal {M}$
for general signature
$(r,s)$
, we refer the reader to the very recent work of Bijakowski and Hernandez [Reference Bijakowski and Hernandez4]. Lastly, we want to mention the work of the second author where he considered the splitting model for triality groups in [Reference Zhao30] (see also [Reference Zhao31]).
For the rest of this subsection, we fix
$n=2m$
and
$I=\{m\}$
. Define

where
$H_m$
is the unit antidiagonal matrix (of size m). For a general signature
$(r,s)$
, we first obtain the following:
Theorem 1.3.
(1) When s is even, there is an affine chart
$\mathcal {U} \subset \mathcal {M}$
which is isomorphic to
$\mathrm {Spec \, } O_{K_1}[X,Y]/I $
, where

for some choice of
$Y'$
. Here, X, Y are matrices of sizes
$s\times s$
and
$n\times s$
, respectively, with indeterminates as entries and
$Y'$
is a submatrix of Y of size
$ s \times s$
(i.e., it is composed of s rows from Y along with the corresponding s columns).
(2a) When
$s=1$
, there is an affine chart
$\mathcal {U} \subset \mathcal {M}$
which is isomorphic to
$\mathbb {A}_{O_{K_1}}^{n-1}$
.
(2b) When s is odd and
$s\geq 3$
, there is an affine chart
$\mathcal {U} \subset \mathcal {M}$
which is isomorphic to
$\mathbb {A}_{O_{K_1}}^{s-1}\times \mathrm {Spec \, } O_{K_1}[X, Y]/I$
, where

for some choice of
$Y'$
. Here, X, Y are matrices of sizes
$(s-1)\times (s-1)$
and
$(n-2)\times s$
, respectively, with indeterminates as entries and
$Y'$
is a submatrix of Y of size
$ (s-1)\times s$
(i.e., it is composed of
$(s-1)$
rows from Y along with the corresponding s columns).
By using the explicit description of
$\mathcal {U}$
above, we prove that
$\mathcal {G}$
-translates of
$\mathcal {U}$
cover
$\mathrm {M}^{\mathrm {spl}}$
and we deduce the following:
Theorem 1.4.
-
a) When
$s=1$ ,
$\mathrm {M}^{\mathrm {spl}}$ is a smooth scheme.
-
b) When
$s=2$ or
$s=3$ ,
$\mathrm {M}^{\mathrm {spl}}$ has semi-stable reduction. In particular,
$\mathrm {M}^{\mathrm {spl}}$ is regular and has special fiber a reduced divisor with two smooth irreducible components intersecting transversely.
When
$s=2$
or
$s=3$
, we show, by using Theorem 1.2, that one of the two components of the special fiber of
$\mathrm {M}^{\mathrm {spl}}$
maps birationally to the special fiber of
$\mathrm { M}^{\mathrm {loc}}$
while the other irreducible component is the inverse image of the unique closed
${\mathcal G}$
-orbit.
Moreover, for
$(r,s)=(n-1,1) $
, we deduce that
$ \mathrm {M}^{\mathrm {spl}}$
is equal to the local model
$\mathrm {M}^{\mathrm {loc}} \otimes _{\mathcal {O}} O_{K_1}$
(see Remark 5.12). In [Reference Pappas and Rapoport18, §5.3], the authors showed that
$\mathrm {M}^{\mathrm {loc}}$
is smooth. This is a case of ‘exotic’ good reduction. We also refer the reader to the result of [Reference Haines and Richarz8] where the authors give an alternative explanation for the smoothness of
$\mathrm {M}^{\mathrm {loc}}$
by identifying its special fiber with a Schubert variety attached to a minuscule cocharacter in the twisted affine Grassmannian corresponding to
$P_{\{m\}}$
.
Lastly, we want to mention that we can apply these results to obtain regular (formal) models of the corresponding Rapoport-Zink spaces.
Let us now outline the contents of the paper: In §2, we recall the parahoric subgroups of the unitary similitude group when
$n=2m$
. In §3, we recall the definition of certain variants of local models for ramified unitary groups. The explicit equations of
$\mathcal {U}$
are given in §4. In §5.1, we show that the naive splitting model
$\mathcal {M}$
is not flat and prove that
${\mathcal G}$
-translates of
$\mathcal {U}$
cover
$\mathrm {M}^{\mathrm {spl}}$
for the signature
$(r,s)$
(
$s\leq 3$
) in §5.2 and §5.3, respectively. In §6, we use the above results to construct smooth integral models for the signature
$(r,s)=(n-1,1)$
, and integral models with semi-stable reduction for
$(r,s)=(n-2,2)$
and
$(r,s)=(n-3,3)$
. In §7, we give a moduli-theoretic description of
$\mathrm {M}^{\mathrm {spl}}$
for
$s\leq 3$
by considering the spin condition.
2 Preliminaries
2.1 Pairings and Standard Lattices
We use the notation of [Reference Pappas and Rapoport18]. Let
$F_0$
be a complete discretely valued field with ring of integers
$O_{F_0}$
, perfect residue field k of characteristic
$\neq 2$
, and uniformizer
$\pi _0$
. Let
$F/F_0$
be a ramified quadratic extension and
$\pi \in F$
a uniformizer with
$\pi ^2 = \pi _0$
. Let V be a F-vector space of dimension
$n =2m> 3$
and let

be an
$F/F_0$
-hermitian form. We assume that
$\phi $
is split. This means that there exists a basis
$e_1, \dots , e_n$
of V such that

We attach to
$\phi $
the respective alternating and symmetric
$F_0$
-bilinear forms
$V \times V \rightarrow F_0$

For any
$O_F$
-lattice
$\Lambda $
in V, we denote by

the dual lattice with respect to the alternating form and by

the dual lattice with respect to the symmetric form. We have
$ \hat {\Lambda }^s = \pi ^{-1}\hat {\Lambda }.$
Both
$\hat {\Lambda }$
and
$\hat {\Lambda }^s$
are
$O_F$
-lattices in V, and the forms
$\langle \, , \, \rangle $
and
$ (\, , \,) $
induce perfect
$O_{F_0}$
-bilinear pairings

for all
$\Lambda $
. Also, the uniformizing element
$\pi $
induces a
$O_{F_0}$
- linear mapping on
$\Lambda $
which we denote by t.
For
$i= 0, \dots , n- 1$
, we define the standard lattices

We consider nonempty subsets
$I \subset \{0, \dots , m\}$
with the property

(see [Reference Pappas and Rapoport18, §1.2.3(b)] for more details). We complete the
$\Lambda _i$
with
$i \in I$
to a self-dual periodic lattice chain by first including the duals
$ \Lambda _{n-i}:=\hat {\Lambda }^s_i $
for
$i \neq 0$
and then all the
$\pi $
-multiples: For
$j \in \mathbb {Z}$
of the form
$j = k \cdot n \pm i$
with
$i \in I$
, we put
$\Lambda _{j} = \pi ^{-k}\Lambda _i$
. Then
$ \{\Lambda _j\}_j$
form a periodic lattice chain
$\Lambda _I$
(with
$\pi \Lambda _j = \Lambda _{j-n}$
) which satisfies
$ \hat {\Lambda }_j = \Lambda _{-j}$
. We denote by
$\mathcal {L}$
such a self-dual multichain. Observe that the lattice
$\Lambda _0$
is self-dual for the alternating form
$\langle \,,\,\rangle $
and
$\Lambda _m$
is self-dual for the symmetric form
$(\,,\,)$
.
2.2 Unitary Similitude Group and Parahoric Subgroups
We consider the unitary similitude group

and we choose a partition
$n = r + s$
; we refer to the pair
$(r, s)$
as the signature. By replacing
$\phi $
by
$-\phi $
if needed, we can make sure that
$s\leq r$
, and so we assume that
$s \leq r$
(see [Reference Pappas and Rapoport18, §1.1] for more details). Identifying
$G \otimes F \simeq GL_{n,F} \times \mathbb {G}_{m,F}$
, we define the cocharacter
$ \mu _{r,s}$
as
$ (1^{(s)}, 0^{(r)}, 1)$
of
$D \times \mathbb {G}_m$
, where D is the standard maximal torus of diagonal matrices in
$GL_n$
; for more details, we refer the reader to [Reference Smithling26]. We denote by E the reflex field of
$\{ \mu _{r,s}\}$
; then
$E = F_0$
if
$r = s$
and
$E = F$
otherwise (see [Reference Pappas and Rapoport18, §1.1]). We set
$O := O_E$
.
We next recall the description of the parahoric subgroups of G from [Reference Pappas and Rapoport18, §1.2], which actually follows from the results on parahoric subgroups of
$\mathrm {SU}(V,\phi )$
in [Reference Pappas and Rapoport19, §1.2]. Recall that
$n=2m$
is even and I is a nonempty subset of
$\{0,\dots ,m\}$
satisfying (2.1.2). Consider the subgroup

The subgroup
$P_I$
is not a parahoric subgroup in general since it may contain elements with nontrivial Kottwitz invariant. Consider the kernel of the Kottwitz homomorphism:

where
$\kappa _G : G(F_0) \twoheadrightarrow \mathbb {Z} \oplus (\mathbb {Z}/2\mathbb {Z})$
(see also [Reference Smithling26, §3] for more details). We have the following (see [Reference Pappas and Rapoport18, §1.2.3(b)]):
Proposition 2.1. The subgroup
$P_I^0$
is a parahoric subgroup, and every parahoric subgroup of G is conjugate to
$P_I^0$
for a unique nonempty
$I \subset \{0,\dots ,m\}$
satisfying (2.1.2). For such I, we have
$P_I^0 = P_I$
exactly when
$m \in I$
. The special maximal parahoric subgroups are exactly those conjugate to
$P^0_{\{m\}} = P_{\{m\}}$
.
In this paper, we will focus on the special maximal parahoric subgroup when n is even. So, we fix
$n=2m$
,
$I=\{m\}$
, and we let
$\mathcal {L}$
be the multichain defined in 2.1 for this choice of I. Denote by
${\mathcal G}$
the (smooth) group scheme of automorphisms of the polarized chain
$\mathcal {L}$
over
$O_{F_0}$
; then
${\mathcal G}$
is the parahoric group model of G in the sense of Bruhat-Tits [Reference Bruhat and Tits5] with
${\mathcal G}(O_{F_0}) =P_{\{m\}}$
. It has connected fibers (see [Reference Pappas and Rapoport18, §1.5]).
3 Rapoport-Zink local models and variants
We briefly recall the definition of certain variants of local models for ramified unitary groups that correspond to the local model triples
$(G, \mu _{r,s}, P_{\{m\}} )$
.
Let
$\mathrm {M}^{\mathrm {naive}}$
be the functor which associates to each scheme S over
$\mathrm {Spec \, } O$
the set of subsheaves
$\mathcal {F}$
of
$O \otimes \mathcal {O}_S $
-modules of
$ \Lambda _m \otimes \mathcal {O}_S $
such that
-
(1)
$\mathcal {F}$ as an
$\mathcal {O}_S $ -module is Zariski locally on S a direct summand of rank n;
-
(2)
$\mathcal {F}$ is totally isotropic for
$( \, , \, ) \otimes \mathcal {O}_S$ ;
-
(3) (Kottwitz condition)
$\text {char}_{t | \mathcal {F} } (X)= (X + \pi )^r(X - \pi )^s $ .
Remarks 3.1. In [Reference Pappas and Rapoport18, §1.5.1], the authors define the naive local model
$\mathrm {M}^{\mathrm {naive}}$
that sends each O-scheme S to the families of
$O\otimes {\mathcal O}_S$
-modules
$(\mathcal {F}_i\subset \Lambda _i\otimes {\mathcal O}_S)_{i \in n\mathbb {Z}\pm I}$
that satisfy the conditions (a)-(d) of loc. cit. We want to explain how we get the isotropic condition (2) from the condition (c) of loc. cit. When
$I=\{m\}$
, the complete lattice chain gives:

where the isomorphism
$t:\Lambda _{m}\otimes {\mathcal O}_S\rightarrow \Lambda _{-m}\otimes {\mathcal O}_S$
induces an isomorphism
$\mathcal {F}_{m}$
with
$\mathcal {F}_{-m}$
. Hence,
$\mathcal {F}_{-m}$
is determined by
$\mathcal {F}_m$
and we have
$\mathcal {F}_{-m}=t\mathcal {F}_{m}$
. The perfect bilinear pairing

induced by (2.1.1) identifies
$\mathcal {F}_m^\bot \subset \Lambda _{-m}\otimes {\mathcal O}_S$
with
$\mathcal {F}_{-m}$
where
$\mathcal {F}_m^\bot $
is the orthogonal complement of
$\mathcal {F}_m$
under the above perfect pairing. Thus,
$\langle \mathcal {F}_{-m},\mathcal {F}_m\rangle =0$
(condition (c) of loc. cit.) is equivalent to
$(\mathcal {F}_m,\mathcal {F}_m)=0$
since

The functor
$\mathrm {M}^{\mathrm {naive}}$
is represented by a closed subscheme, which we again denote
$\mathrm {M}^{\mathrm {naive}}$
, of
$\mathrm {Gr}(n, 2n) \otimes \mathrm {Spec \, } O$
; hence,
$\mathrm {M}^{\mathrm {naive}}$
is a projective
$ O$
-scheme. (Here, we denote by
$\mathrm {Gr}(n, d)$
the Grassmannian scheme parameterizing locally direct summands of rank n of a free module of rank d.) The scheme
$\mathrm {M}^{\mathrm {naive}}$
is the naive local model of Rapoport-Zink [Reference Rapoport and Zink24]. Also,
$\mathrm {M}^{\mathrm {naive}}$
supports an action of
$\mathcal {G}$
.
Proposition 3.2. We have

In particular, the generic fiber of
$\mathrm {M}^{\mathrm {naive}}$
is smooth and geometrically irreducible of dimension
$rs$
.
Proof. See [Reference Pappas and Rapoport18, §1.5.3].
Next, we consider the closed subscheme
$\mathrm {M}^{\wedge }\subset \mathrm {M}^{\mathrm {naive}}$
by imposing the following additional condition:

More precisely,
$\mathrm {M}^{\wedge }$
is the closed subscheme of
$\mathrm {M}^{\mathrm {naive}}$
that classifies points given by
$\mathcal {F}$
which satisfy the wedge condition. The scheme
$\mathrm { M}^{\wedge }$
supports an action of
$\mathcal {G}$
and the immersion
$i : \mathrm {M}^{\wedge } \rightarrow \mathrm {M}^{\mathrm {naive}}$
is
$\mathcal {G}$
-equivariant. The scheme
$\mathrm {M}^{\wedge }$
is called the wedge local model.
Proposition 3.3. The generic fibers of
$\mathrm {M}^{\mathrm {naive}} $
and
$\mathrm {M}^{\wedge } $
coincide; in particular, the generic fiber of
$\mathrm {M}^{\wedge }$
is a smooth, geometrically irreducible variety of dimension rs.
Proof. See [Reference Pappas and Rapoport18, §1.5.6].
For the special maximal parahoric subgroup
$P_{\{0\}}$
and signature
$(n-1,1)$
, Pappas proved that the wedge local model
$\mathrm {M}^{\wedge }$
is flat [Reference Pappas15]. But for more general parahoric subgroups, the wedge condition turns out to be insufficient (see [Reference Pappas and Rapoport18, Remark 5.3, 7.4]). There is a further variant: let
$\mathrm {M}^{\mathrm {loc}}$
be the scheme theoretic closure of the generic fiber
$ \mathrm {M}^{\mathrm {naive}} \otimes _{O} E$
in
$\mathrm {M}^{\mathrm {naive}}$
. The scheme
$\mathrm {M}^{\mathrm {loc}}$
is called the local model. We have closed immersions of projective schemes

which are equalities on generic fibers (see [Reference Pappas and Rapoport18, §1.5] for more details).
Proposition 3.4.
a) For any signature
$(r, s)$
, the special fiber of
$\mathrm {M}^{\mathrm {loc}}$
is integral, normal and Cohen-Macaulay and has only rational singularities.
b) For
$(r, s)= (n-1,1)$
,
$\mathrm {M}^{\mathrm {loc}}$
is smooth.
Proof. See [Reference Pappas and Rapoport18, §5] and [Reference Haines and Richarz9].
Except in the case (b) above (up to switching
$(r, s)$
and
$(s,r)$
), the local models
$\mathrm {M}^{\mathrm {loc}}$
are never smooth; see [Reference Pappas, Rapoport and Smithling20, Remark 2.6.8] for more details. As in [Reference Pappas and Rapoport18, §2.4.2, §5.5], there is a unique closed
${\mathcal G}$
-orbit in the special fiber of
$\mathrm {M}^{\mathrm {loc}}$
. When s is even, the closed orbit is the
$ k$
-valued point

and when s is odd, the closed orbit is the orbit of the k-valued point

Next, we consider the moduli scheme
$\mathcal {M}$
over
$O_F$
, the naive splitting (or Krämer) model as in [Reference Pappas and Rapoport17, Remark 14.2] and [Reference Krämer14, Definition 4.1], whose points for an
$O_F $
-scheme S are Zariski locally
$\mathcal {O}_S$
-direct summands
$ \mathcal {F}_0 , \mathcal {F}_1 $
of ranks s, n, respectively, such that
-
(1)
$ (0) \subset \mathcal {F}_0 \subset \mathcal {F}_1 \subset \Lambda _m \otimes \mathcal {O}_S$ ;
-
(2)
$\mathcal {F}_1 = \mathcal {F}^{\bot }_1$ , i.e.
$\mathcal {F}_1$ is totally isotropic for
$( \, , \, ) \otimes \mathcal {O}_S$ ;
-
(3)
$(t + \pi ) \mathcal {F}_1 \subset \mathcal {F}_0$ ;
-
(4)
$(t - \pi )\mathcal {F}_0 = (0)$ .
The functor is represented by a projective
$O_F $
-scheme
$ \mathcal {M}$
. The scheme
$ \mathcal {M}$
supports an action of
${\mathcal G}$
, and there is a
${\mathcal G}$
-equivariant projective morphism

which is given by
$(\mathcal {F}_0,\mathcal {F}_1) \mapsto \mathcal {F}_1$
on S-valued points. (Indeed, we can easily see, as in [Reference Krämer14, Definition 4.1], that
$\tau $
is well defined.) The morphism
$\tau : \mathcal {M} \rightarrow \mathrm {M}^{\wedge }\otimes _{O} O_F $
induces an isomorphism on the generic fibers (see [Reference Krämer14, Remark 4.2]). In §5, we will prove that
$ \mathcal {M} $
is not flat for any signature
$(r,s)$
, and we will study the following variation:

The closed subscheme
$\mathrm {M}^{\mathrm {spl}} $
is the splitting model. We will show that this model is smooth for the signature
$(n-1,1) $
and has semi-stable reduction for the signatures
$(n-2,2)$
and
$(n-3,3)$
. Moreover, for these signatures, we will show that
$\mathrm {M}^{\mathrm {spl}}$
has an explicit moduli-theoretic description.
4 Two affine charts
The goals of this section are to write down the equations that define the naive splitting model
$\mathcal {M}$
in two open neighborhoods
$_1{\mathcal U}_{r,s}$
and
$_2{\mathcal U}_{r,s}$
for general signature
$(r,s)$
. Here,
$_1{\mathcal U}_{r,s}$
is an affine chart around
$(\mathcal {F}_0, t\Lambda _m)$
(see §4.1) and
$_2{\mathcal U}_{r,s}$
is an affine chart around
$(\mathcal {F}_0, \Lambda ')$
, where
$\Lambda '$
is defined in §4.2.
4.1 An Affine Chart
$_1{\mathcal U}_{r,s}$
Recall that
$n=2m= r+s$
with
$s\leq r$
. In this subsection, we are going to write down the equations that define
$\mathcal {M}$
in a neighborhood
$_1\mathcal {U}_{r,s}$
of
$(\mathcal {F}_0, t\Lambda _m)$
, where


The matrix of
$(\,,\, )$
in the latter basis is

where

and
$H_m$
is the unit antidiagonal matrix (of size m). Observe that
$ J^2_n = -I_n.$
Also, since
$ s \leq r$
and
$ n =2m =r+s$
, we get that
$s\leq m$
.
In order to find the explicit equations that describe
$_1{\mathcal U}_{r,s}$
, we use similar arguments as in the proof of [Reference Krämer14, Theorem 4.5] (see also [Reference Zachos28, §3]). In our case, we consider

where A is of size
$ n \times n$
, X is of size
$2n \times s$
and
$X_1,X_2$
are of size
$n\times s$
; with the additional condition that
$\mathcal {F}_0$
has rank s and so X has an invertible
$s\times s$
-minor. We also ask that
$(\mathcal {F}_0,\mathcal {F}_1)$
satisfy the following four conditions:
-
(1)
$\mathcal {F}_1^{\bot } = \mathcal {F}_1$ ,
-
(2)
$ \mathcal {F}_0 \subset \mathcal {F}_1$ ,
-
(3)
$(t - \pi )\mathcal {F}_0 = (0)$ ,
-
(4)
$(t + \pi ) \mathcal {F}_1 \subset \mathcal {F}_0$ .
Observe that

of size
$2n \times 2n $
is the matrix giving multiplication by t.
Condition (1) (i.e.,
$\mathcal {F}_1$
is isotropic) translates to
$A^{t} = -J_nAJ_n$
. Condition (2) (i.e.,
$ \mathcal {F}_0 \subset \mathcal {F}_1$
) implies that the generators of
$\mathcal {F}_0$
can be expressed as a linear combination of the generators in
$\mathcal {F}_1$
, which translates to

where Y is of size
$ n\times s$
. Thus, we have

and so
$ X_1 = AY$
and
$X_2 = Y.$
Condition (3) (i.e.,
$(t - \pi )\mathcal {F}_0 = (0)$
) is equivalent to

which amounts to

Thus,
$X_1 = \pi X_2$
, which translates to
$A Y = \pi Y $
by condition
$(2)$
. The last condition
$(t + \pi ) \mathcal {F}_1 \subset \mathcal {F}_0$
translates to

where Z is of size
$n\times s$
. This amounts to

From the above, we get
$A + \pi I_n = X_2 Z^t$
, which by condition (2) translates to
$ A = Y Z^t-\pi I_n. $
Thus, A can be expressed in terms of
$Y,Z$
. In addition, by condition (2) and in particular by the relations
$ X_1 = AY$
and
$X_2 = Y$
, we deduce that the matrix X is given in terms of
$Y,Z$
. Conversely, from
$Y=X_2$
, we get that the matrix Y is given in terms of
$A,X$
(Below we will also show that Z can be expressed in terms of
$A,X$
).
Observe from
$X_1 = \pi X_2$
that all the entries of
$X_1$
are in the maximal ideal and thus a minor involving entries of
$X_1$
cannot be a unit. Recall that the matrix X has an invertible
$s\times s$
-minor and
$X_2=Y$
from condition (2). Thus, the matrix Y has a
$s\times s$
-minor. Let

and

where
$1 \leq i_k \leq n$
for
$k = 1, \dots ,s$
. Here, we want to mention how we can express Z in terms of
$A,X$
. From the above, we have
$Y Z^t=A+ \pi I_n$
and

By using (4.1.1), for any
$1 \leq k \leq s$
, the
$i_k$
-th row of the matrix
$Y Z^t $
gives
$\mathbf {z}^t_k $
. Since
$Y Z^t=A+ \pi I_n$
, we can easily see that the matrix Z can be expressed in terms of
$A,X$
.
We replace A by
$Y Z^t-\pi I_n$
. Hence, conditions (
$1$
) and (
$3$
) are equivalent to


From the above, we deduce the following.
Lemma 4.1. An affine chart
$_1{\mathcal U}_{r,s}$
of
$\mathcal {M}$
around
$(\mathcal {F}_0, t \Lambda _m)$
is given by
$\mathrm {Spec \, } O_F[Y , Z ]/\mathscr {I} $
, where

and
$Y'$
is a submatrix composed of s rows of Y (see the relation (4.1.1)).
Note that the affine chart
$_1{\mathcal U}_{r,s}$
of
$\mathcal {M}$
depends on the matrix
$Y'$
we fixed (see Remark 4.4). Our next goal is to give a simplification of the above equations (see Theorem 4.2). We first study the relation
$Z Y^t =-J_nY Z^tJ_n$
. Consider the matrix
$P = (p_{i,j})$
of size
$n=2m$
with
$P^t=-J_nPJ_n$
. By a direct calculation, we get

for
$1\leq i, j\leq m$
or
$m+1\leq i,j\leq n$
, and

for
$1\leq i\leq m$
and
$m+1\leq j\leq n$
, or
$1\leq j\leq m$
and
$m+1\leq i\leq n$
. By using this observation and by setting
$P = Y Z^t$
, the relation
$ Z Y^t=-J_nY Z^tJ_n$
gives

for
$1\leq i,j \leq n$
, where the different signs
$\pm $
depends on the position of
$i,j$
as above. Consider the following 2 cases:
Case 1: Assume that
$ 1\leq i_k \leq m$
for all
$k \in \{1,\dots ,s\}.$
By setting
$i = n+1-i_q $
and
$j= i_p$
, the equation (4.1.4) gives
$ z_{n+1-i_q,p} = -z_{n+1-i_p,q} .$
In particular, if
$p=q$
, we get
$z_{n+1-i_p,p} = 0 $
for all
$p \in \{1,\dots ,s\}$
. By setting
$ d_{p,q} = z_{n+1-i_p,q}$
, we get the matrix
$D = (d_{p,q})_{p,q}$
of size
$s \times s$
such that

The matrix D is skew-symmetric (i.e.,
$D = -D^t).$
Next, by setting
$j = i_k$
in (4.1.4), we obtain

Thus, we have
$ Z = \widetilde {Y} \cdot D$
, where

Therefore, in this case, the equations (4.1.4) give
$ Z = \widetilde {Y} \cdot D $
.
Case 2: In general, we let
$ 1 \leq t \leq s$
and assume that

By setting
$d_{p,q} = z_{n+1-i_p,q}$
and using the same arguments as above, we get the matrix
$D = (d_{p,q})_{p,q}$
of size
$s \times s$
such that

for
$1\leq p, q \leq t$
or
$t+1\leq p,q\leq n$
, and

for
$1\leq p\leq t$
and
$t+1\leq q\leq n$
, or
$1\leq q\leq t$
and
$t+1\leq p\leq n$
. Next, by setting
$j = i_k$
in (4.1.4) for
$1\leq k \leq t$
, we obtain

Also, by setting
$j = i_k$
for
$t+1\leq k \leq s$
, we get

From the above, we have
$ Z = \widetilde {Y} \cdot \widetilde {D}$
, where
$\widetilde {D} = (\widetilde {d}_{i,j})_{i,j}$
is the
$s \times s$
matrix defined as

It is easy to see that the matrix
$ \widetilde {D}$
is skew symmetric (i.e.,
$\widetilde {d}_{i,j} = - \widetilde {d}_{j,i}).$
In particular, we have the following 4 cases: when
$1 \leq i,j \leq t$
, we have
$ \widetilde {d}_{i,j} = d_{i,j} = -d_{j,i} = - \widetilde {d}_{j,i}$
. For
$1 \leq i \leq t$
and
$t+1 \leq j \leq s $
, we have
$ \widetilde {d}_{i,j} = -d_{i,j} = -d_{j,i} = - \widetilde {d}_{j,i}$
. When
$1 \leq j \leq t$
and
$t+1 \leq i \leq s $
, we get
$ \widetilde {d}_{i,j} = d_{i,j} = d_{j,i} = - \widetilde {d}_{j,i}$
. Lastly, for
$t+1 \leq i,j \leq s$
, we obtain
$ \widetilde {d}_{i,j} = -d_{i,j} = d_{j,i} = - \widetilde {d}_{j,i}$
.
Combining the above cases, for any
$1\leq i_1,\dots , i_s\leq n$
, the equations (4.1.4) show that there exists a skew symmetric matrix
$\mathcal {D}$
of size
$s \times s$
such that

By abuse of notation, write
$ D$
for
$ \mathcal {D}$
. Next, define
$ Q : =\widetilde {Y}^t \cdot Y $
of size
$ s \times s.$
Note that
$ Q^t = -Q.$
We are now ready to show the following:
Theorem 4.2. For any signature
$(r,s)$
, there is an affine chart
$_1{\mathcal U}_{r,s} \subset \mathcal {M}$
which is isomorphic to
$\mathrm {Spec \, } O_F[ D, Y]/I $
, where

for some choice of
$Y'$
.
Proof. Recall from Lemma 4.1 that
${\mathcal U} = \mathrm {Spec \, } O_F[Y , Z ]/\mathscr {I} $
, where

From the above discussion, we have that the relation
$Z Y^t =-J_nY Z^tJ_n $
gives
$ Z = \widetilde {Y} \cdot D$
. Thus, it suffices to focus on the relation
$ Y Z^t Y - 2 \pi Y=0$
. Observe that
$ Y \cdot Z^t \cdot Y =- Y \cdot D \cdot Q.$
Thus,

Now, by using the relation with the minors (4.1.1), we can easily see that the relation
$Y Z^t Y - 2 \pi Y=0$
gives
$D \cdot Q + 2 \pi I_s=0.$
Therefore,
$_1{\mathcal U}_{r,s}$
is isomorphic to

In particular, when the signature
$(r,s)=(n-2,2)$
, we have

where
$d=d_{2,1}=-d_{12}$
and
$q=\sum ^m_{j=1}y_{n+1-j,1}y_{j,2} - \sum ^n_{j=m+1}y_{n+1-j,1}y_{j,2}$
.
Corollary 4.3. When
$(r,s)=(n-2,2)$
, there is an affine chart
$_1\mathcal {U}_{r,s} \subset \mathcal {M}$
which is isomorphic to
$V(I) = \mathrm {Spec \, } O_F[d,(y_{i,1})_{1 \leq i \leq n},(y_{i,2})_{1 \leq i \leq n} ]/ I$
, where

for some
$1\leq i_0,j_0 \leq n$
.
Remark 4.4. Recall that
$_1{\mathcal U}_{r,s}$
is an affine chart around
$(\mathcal {F}_0, t\Lambda _m)$
. Note that for different choices of
$Y'$
, we get different equations from
$D\cdot Q =- 2 \pi I_s$
and so different affine charts
$_1{\mathcal U}_{r,s}$
which in general are not isomorphic. For instance, in Corollary 4.3, there are two different
$V(I)$
up to isomorphism when we choose different places for
$i_0, j_0$
(see §5.2 for more details). In Proposition 5.2, we show that for any such a choice, the affine scheme
$V(I)$
has semi-stable reduction over
$O_F$
.
4.2 An Affine Chart
$_2{\mathcal U}_{r,s}$
Keep the same notation as above. As in [Reference Pappas and Rapoport18], we set
$\Lambda _m=\Lambda '\oplus \Lambda "$
, where

In this subsection, we are going to write down the equations that define
$\mathcal {M}$
in a neighborhood
$_2\mathcal {U}_{r,s}$
of
$(\mathcal {F}_0, \Lambda ')$
.
The matrix of the form
$(\ ,\ )$
in this basis is

where S is the skew-symmetric matrix of size n given by

Observe that
$S^2=-I_n$
since
$J_n^2=-I_n$
. For any point
$(\mathcal {F}_0,\mathcal {F}_1)\in \mathcal {M}$
, we ask that
$ (\mathcal {F}_0,\mathcal {F}_1)$
satisfy the following conditions:
-
(1)
$\mathcal {F}_1^{\bot } = \mathcal {F}_1$ ,
-
(2)
$ \mathcal {F}_0 \subset \mathcal {F}_1$ ,
-
(3)
$(t - \pi )\mathcal {F}_0 = (0)$ ,
-
(4)
$(t + \pi ) \mathcal {F}_1 \subset \mathcal {F}_0$ .
Similar to §4.1, we consider

where
$I_n, A$
are
$n\times n$
-matrices, and
$X_1, X_2$
are
$n\times s$
-matrices. Recall that
$\mathcal {F}_0$
has rank s, so X has a no-vanishing
$s\times s$
minor. Write down the matrix A by

where T is a
$2\times 2$
-matrix, B (resp. C) is a
$2\times (n-2)$
-matrix (resp.
$(n-2)\times 2$
-matrix) and Y is of size
$(n-2)\times (n-2)$
. Then condition (1) is equivalent to
$A^t S=SA$
. This translates to

Note that the first equation immediately gives
$T=\text {diag}(x,x)$
for some variable x, and the second equation shows that C is given in terms of B. Condition (2) is equivalent to

where M is a
$n\times s$
-matrix. This amounts to
$X_1=M, X_2=AM$
. Let

where
$M_1$
is of size
$2\times s$
and
$M_2$
is of size
$(n-2)\times s$
. We can see that
$X_1, X_2$
can be expressed in terms of
$A, M_1, M_2$
. More precisely, we get

The map
$t: \Lambda _m\rightarrow \Lambda _m$
is represented by the matrix

where
$t_0$
is the
$2\times 2$
-antidiag matirx
$\left [\begin {array}{cc} & \pi _0\\ 1& \end {array} \right ]$
. Condition (3) translates to

By setting
$B_i$
as the i-th row of B and by using condition (2), the above relation amounts to

where
$M_0$
is the second row of
$M_1$
. Finally, condition (4) is equivalent to

where
$Z_1$
is of size
$2\times s$
and
$Z_2$
is of size
$(n-2)\times s$
. This, in turn, gives

From the above equations, we can see that A is given in terms of
$x, C, Y$
by condition (1). Note that
$C=M_2Z_1^t$
, and
$Y=M_2Z_2^t-\pi I_{n-2}$
by condition (4). Thus, the matrix A is given in terms of
$M, Z_1, Z_2$
and a variable x. Meanwhile, we have
$X_1=M, X_2=AM$
by condition (2), so that
$X_1, X_2$
are also given in terms of
$M, Z_1, Z_2$
. However, we have
$M=X_1$
, and we claim that
$Z_1, Z_2$
can be expressed in terms of
$A, X_1, X_2$
(we will show this later in this section).
Before we move on, let us simplify our equations first. By replacing
$Y=M_2Z_2^t-\pi I_{n-2}$
, the equation
$Y^t=-J_{n-2}YJ_{n-2}$
from condition (1) amounts to
$(M_2Z_2^t)^t=-J_{n-2}(M_2Z_2^t)J_{n-2}$
, and the equations that involve Y from condition (3), (4) translate to

Finally, we replace C by
$M_2Z_1^t$
, and deduce the following.
Lemma 4.5. The affine chart
$_2{\mathcal U}_{r,s}$
of
$\mathcal {M}$
around
$(\mathcal {F}_0, \Lambda ')$
is given by

where
$\mathscr {I}$
is the ideal generated by the entries of following equations:

Here,
$B_i$
is the i-th row of B for
$i=1,2$
.
Remark 4.6. Condition (3) also gives the equations:

In [Reference Pappas and Rapoport18, §5.3], the authors show that these two equations are automatically true if
$B,C,Y$
satisfy

4.2.1 The case
$(r,s)=(n-1,1)$
In this and the next subsection, our goal is to give a simplification of the equations in Lemma 4.5. We first consider the affine chart
$_2\mathcal {U}_{n-1,1} \subset \mathcal {M}$
for the signature
$(n-1,1)$
. Note that
$M_0$
is of size
$1\times 1$
in this case. Since

we get
$M_0\neq 0$
. Without loss of generality, we assume
$M_0=1$
as rank(
$\mathcal {F}_0$
)=1.
Theorem 4.7. When the signature
$(r,s)=(n-1,1)$
, the affine chart
$_2\mathcal {U}_{n-1,1}$
is smooth and isomorphic to
$\mathbb {A}_{O_F}^{n-1}$
.
Proof. From
$M_0Z_1^t=[1\ \pi ]$
,
$M_0Z_2^t=0$
, we obtain

Let
$M_2=[a_1\ \cdots \ a_{n-2}]^t$
be an
$(n-2)\times 1$
vector. Then
$C=M_2Z_1^t$
amounts to

Also, from
$B=J_2C^tJ_{n-2}$
, we get

It is easy to see that the relation
$Y=M_2Z_2^t-\pi I_{n-2}$
gives
$Y=-\pi I_{n-2}$
and
$B_1+\pi B_2=B_2(M_2Z_2^t)=0$
. Finally, we check that

From the above, the conditions on the
$a_i$
are automatically satisfied. We deduce that the affine chart
$_2\mathcal {U}_{n-1,1}$
is isomorphic to

4.2.2 The cases for
$s\geq 2$
In this subsection, we consider the explicit equations of
$_2\mathcal {U}_{r,s}$
for the signature
$(r,s)$
with
$s\geq 2$
. Observe that
$\mathcal {F}_1=\{v+A(v)\mid v\in \Lambda '\otimes {\mathcal O}_S \}$
. In the special fiber (where
${\mathcal O}_S=k$
), we have
$t\mathcal {F}_1\subset \mathcal {F}_0$
, and
$t(\Lambda '\otimes k)=\text {span}_{O_{F_0}}\{-e_1\}$
, which implies
$-e_1+tA(v)\in \mathcal {F}_0$
. Since

we can always assume that
$-e_1+tA(v)$
is the first generating element of
$\mathcal {F}_0$
. In the matrix language, this is equivalent to assume that the second row of X (i.e.,
$M_0$
) is equal to
$[1\ 0 \cdots 0]$
. Thus, we have

By using
$M_0Z_1^t=[1 \ \pi ]$
,
$M_0Z_2^t=0$
, we set

and

We claim that
$Z_1, Z_2$
can be expressed in terms of
$A, X_1, X_2$
. Observe from
$CM_1+YM_2=\pi M_2$
that all the entries of
$CM_1+YM_2$
are in the maximal ideal, and thus a minor involving entries of
$CM_1+YM_2$
cannot be a unit. Similarly, the entries of the first row of
$xM_1+BM_2$
are also in the maximal ideal. By replacing the order of basis
$\{\pi e_m,\dots , \pi e_n\}$
if necessary, we can assume that the matrix
$M_2$
has an invertible
$(s-1)\times (s-1)$
-minor. Let
$M_2'$
be a submatrix of
$M_2$
composing
$(s-1)$
rows along with the corresponding s columns. We set

for
$1\leq i_2, \cdots , i_s\leq n-2$
. Consider the matrix
$C=M_2Z_1^t$
. We have

By setting
$i=i_2, \dots , i_s$
, it is easy to see that
$Z_1$
can be expressed in terms of C. Similarly,
$Z_2$
can be expressed in terms of Y as
$M_2Z_2^t=\pi I_{n-2}+Y$
. Thus,
$Z_1, Z_2$
are given in terms of
$A, X_1, X_2$
.
Set

where
$q_{i,j}=\sum _{k=1}^{m-1}(a_{k,i}a_{n-1-k,j}-a_{n-1-k,i}a_{k,j})$
. The matrix Q depends on
$M_2$
. Note that
$Q^t=-Q$
by definition.
Theorem 4.8. For any signature
$(r,s)$
with
$s\geq 2$
, there is an affine chart
$_2\mathcal {U}_{r,s}\subset \mathcal {M}$
which is isomorphic to

where

for some choice of
$M^{\prime }_2$
.
$[\mathbf {0}\mid I_{s-1}]$
is of size
$(s-1)\times s$
, with the first column
$0$
.
Proof. The proof is similar to the proof of Theorem 4.2. Consider
$(M_2Z_2^t)^t=-J_{n-2}(M_2Z_2^t)J_{n-2}$
first. We claim that
$Z_2^t$
depends on
$M_2$
and D. Since
$M_2Z_2^t=(\sum _{k=2}^{s}a_{i,k}z_{j,k}')_{ij}$
, the relation
$(M_2Z_2^t)^t=-J_{n-2}(M_2Z_2^t)J_{n-2}$
amounts to

where the different sign
$\pm $
depends on the position of
$i_2,\dots , i_s$
. We obtain

by letting
$i=i_k$
,
$j=n-1-i_k$
for
$k=2,\dots ,s$
. To write down
$Z_2^t$
explicitly, we need to consider the positions of
$i_2,\dots ,i_s$
.
Case 1: Assume that
$1\leq i_2,\dots ,i_s\leq m-1$
. In this case, we have
$z_{n-1-i_p,q}'=-z_{n-1-i_q,p}'$
by letting
$i=i_{p}$
,
$j=n-1-i_{q}$
. Set
$d_{p,q}=z_{n-1-i_p,q}'$
. We obtain

where
$d_{p,q}=-d_{q,p}$
for
$2\leq p,q\leq s$
. This implies that D is skew-symmetric.
For
$i=i_2$
, we have

Since
$1\leq i_2\leq m-1$
, the sign
$\pm $
is positive when
$1\leq j\leq m-1$
, and is negative when
$m\leq j\leq n-2$
. Similar to
$i=i_k$
for
$k=3,\dots , s$
. We have

Case 2: In general, we assume that

Set
$d_{p,q}=z_{n-1-i_p,q}'$
. By letting
$i=i_p, j=n-1-i_q$
, we obtain

Thus,
$z_{j,k}'=\pm (a_{n-1-j,k}d_{k,2}+\cdots +a_{n-1-j,s}d_{k,s})$
, where the sign
$\pm $
is positive if
$1\leq j\leq m$
,
$2\leq k\leq t$
, or
$m\leq j\leq n-2$
,
$t+1\leq k\leq s$
, and negative otherwise. Therefore,
$Z_2^t$
can be expressed as

Set

For
$2\leq p,q\leq t$
, we have
$\tilde {d}_{p,q}=d_{p,q}=-d_{q,p}=-\tilde {d}_{q,p}$
. Similarly, it is easy to see that
$\tilde {d}_{p,q}=-\tilde {d}_{q,p}$
for all the other positions of
$p,q$
. Hence, the matrix
$\widetilde {D}$
is skew-symmetric.
Combining the above cases, for any
$1\leq i_2,\dots , i_s\leq n-2$
, there exists a skew-symmetric matrix
$\mathcal {D}$
of size
$(s-1)\times (s-1)$
such that

By abuse of notation, write
$ D$
for
$ \mathcal {D}$
. Our next goal is to show that the second column of
$Z_1^t$
depends on the first column of
$Z_1^t$
and
$M_2$
. Recall that

and
$B=J_2C^tJ_{n-2}$
. By using (4.2.1), the
$(n-1-i_k)$
-th columns (
$k=2,\dots , s$
) of the relation
$B_1+\pi B_2=B_2(M_2Z_2^t)$
give

where
$q_{i,j}=\sum _{k=1}^{m-1}(a_{k,i}a_{n-1-k.j}-a_{n-1-k,i}a_{k,j})$
. It is easy to see that the rest columns in
$B_1+\pi B_2= B_2(M_2Z_2^t)$
are automatically satisfied when (4.2.2) holds.
Finally, consider
$CM_1+(M_2Z_2^t)M_2=2\pi M_2$
. By using the relations (4.2.1) and (4.2.2), it is equivalent to

where

There are two relations we still need to consider:
$B_1M_2=\pi B_2M_2$
, and
$(M_2Z_2^t)^2=2\pi (M_2Z_2^t)$
. We leave it to the reader to verify that these conditions are already satisfied when
$D\cdot Q=-2\pi I_{s-1}$
.
From all the above, we deduce that an affine neighborhood around
$(\mathcal {F}_0,\Lambda ')$
is given by
$ _2\mathcal {U}_{r,s}\cong \mathrm {Spec \, } O_F[x,z_{1,2},\dots ,z_{1,s}, D, M_2]/I$
, where

Set
$\mathbb {A}_{O_F}^{s-1}=\mathrm {Spec \, } O_F[x,z_{1,2},\dots ,z_{1,s}]$
. For
$s\geq 2$
, we have

In particular, for the signature
$(r,s)=(n-3,3)$
, we have

where we set
$d=d_{2,3}=-d_{3,2}, q=q_{2,3}=-q_{3,2}$
. Recall that
$q=\sum _{k=1}^{m-1}(a_{k,2}a_{n-1-k,3}-a_{n-1-k,2}a_{k,3})$
. The equation
$D\cdot Q=-2\pi I_{s-1}$
amounts to
$d\cdot q-2\pi =0$
.
Corollary 4.9. When
$(r,s)=(n-3,3)$
, there is an affine chart
$_2{\mathcal U}_{n-3,3}$
which is isomorphic to

where

for some
$ 1 \leq i_2,i_3 \leq n-2$
.
Remark 4.10. As in Remark 4.4, for different choices of
$M^{\prime }_2$
, we get different equations from
$D\cdot Q = -2 \pi I_s$
and so different affine charts
$_2\mathcal {U}_{r,s}$
which in general are not isomorphic. For instance, in Corollary 4.9, there are two different affine charts up to isomorphism when we choose different positions for
$i_2, i_3$
(see §5.3 for details). In Proposition 5.7, we show that for any such a choice, the affine scheme
$V (I)$
has semi-stable reduction over
$O_F$
.
5 Splitting models
$\mathrm {M}^{\mathrm {spl}}$
In this section, we will first show that the naive splitting model
$\mathcal {M}$
is not flat over
$\mathrm {Spec \, } O_F$
for a general signature
$(r,s)$
. Then we will define the closed subscheme
$\mathrm {M}^{\mathrm {spl}} \subset \mathcal {M} $
. We show that
$\mathrm {M}^{\mathrm {spl}}$
is smooth for the signature
$(r,s)=(n-1,1)$
and has semi-stable reduction for the signature
$(r,s)=(n-2,2)$
and
$(r,s)=(n-3,3)$
.
5.1 Naive Splitting Models
$\mathcal {M}$
By using the explicit description of the two affine charts
$_1{\mathcal U}_{r,s}$
,
$_2{\mathcal U}_{r,s}\subset \mathcal {M}$
, that we obtained in §4, we will show that
$\mathcal {M}$
is not flat over
$\mathrm {Spec \, } O_F$
.
Proposition 5.1. For any signature
$(r,s)$
, the naive splitting model
$\mathcal {M}$
is not flat over
$\mathrm {Spec \, } O_F$
.
Proof. It is enough to show a) When s is odd, the generic fiber
$_1{\mathcal U}_{r,s}\otimes _{O_F}F$
is empty and so the point
$(\mathcal {F}_0, t\Lambda _m)$
does not lift to characteristic zero, and b) When s is even, the generic fiber
$_2{\mathcal U}_{r,s}\otimes _{O_F}F$
is empty and so the point
$(\mathcal {F}_0, \Lambda ')$
does not lift to characteristic zero.
Observe that for a skew-symmetric matrix P of size
$s\times s$
, the determinant of P is equal to
$0$
if s is odd. Recall that we denote by
$_1{\mathcal U}_{r,s}$
the affine chart around the point
$(\mathcal {F}_0, t\Lambda _m)\in \mathcal {M}$
. By Theorem 4.2, the affine chart
$_1{\mathcal U}_{r,s}$
is isomorphic to
$\mathrm {Spec \, } O_F[D,Y]/I$
, where

for any signature
$(r,s)$
. Here, D is a skew-symmetric matrix of size
$s\times s$
. Thus, when s is odd, we have
$\det (D)=0$
. In this case, we can see that the generic fiber
$_1{\mathcal U}_{r,s}\otimes _{O_F}F$
is empty since

It follows that the point
$(\mathcal {F}_0, t\Lambda _m)$
does not lift to characteristic zero when s is odd.
Recall that we denote by
$_2{\mathcal U}_{r,s}$
the affine chart around the point
$(\mathcal {F}_0, \Lambda ')\in \mathcal {M}$
. By Theorem 4.8, the affine chart
$_2{\mathcal U}_{r,s}$
is isomorphic to
$\mathbb {A}^{s-1}_{O_F}\times \mathrm {Spec \, } O_F[D,M_2]/I$
, where

When s is even, we see that the generic fiber
$_2{\mathcal U}_{r,s}\otimes _{O_F}F$
is empty since D is of size
$(s-1)\times (s-1)$
and
$0=\det (D\cdot Q)=(-2\pi )^{s-1}$
. Therefore, the point
$(\mathcal {F}_0, \Lambda ')$
does not lift to characteristic zero when s is even.
Recall that there is a forgetful morphism
$\tau : \mathcal {M}\rightarrow \mathrm {M}^\wedge \otimes _O O_F$
. From [Reference Pappas and Rapoport18, Remark 5.3], [Reference Pappas, Rapoport and Smithling20, Remark 2.6.10], the authors show that the wedge local model
$\mathrm {M}^\wedge $
is not topological flat for any signature when n is even. Consider the local model
$\mathrm {M}^{\mathrm {loc}}$
to be the scheme-theoretic closure in
$\mathrm {M}^\wedge $
of its generic fiber. We have a closed morphism
$\mathrm {M}^{\mathrm {loc}}\hookrightarrow \mathrm {M}^\wedge $
. Define the splitting model as the inverse image of
$\mathrm {M}^{\mathrm {loc}}$
; that is,

The restriction of the forgetful morphism
$\tau $
on
$\mathrm {M}^{\mathrm {spl}}$
gives

5.2 The Case
$(r,s)=(n-2,2)$
By using the explicit description of
$ _1\mathcal {U}_{n-2,2}$
in §4.1, we show the following.
Proposition 5.2. When
$(r,s)=(n-2,2)$
,
$_1{\mathcal U}_{n-2,2}$
has semi-stable reduction over
$O_F$
. In particular,
$_1{\mathcal U}_{n-2,2}$
is regular and has special fiber a reduced divisor with two smooth irreducible components intersecting transversely.
Proof. From Corollary 4.3, it is enough to show that the scheme
$V(I) $
has semi-stable reduction over
$O_F$
. It suffices to prove that
$V(I) $
is regular and its special fiber is reduced with smooth irreducible components that have smooth intersections with correct dimensions (see [Reference de Jong13, §2.16]).
First, we rename our variables as follows:
$x_i := y_{i,1}$
and
$y_i := y_{i,2}$
for
$1 \leq i \leq n.$
Thus,
$V(I) = \mathrm {Spec \, } O_F[d,\underline x, \underline y]/ I$
, where

By symmetry, as above, there two cases to consider: Case 1 when
$1 \leq i_0,\,j_0 \leq m$
and Case 2 when
$ 1 \leq i_0 \leq m$
and
$ m+1 \leq j_0 \leq n $
. We consider these two cases separately in this proof.
In Case 1, we can assume, for simplicity, that
$i_0=1$
and
$ j_0=2$
. Thus,
$V(I)$
is isomorphic to
$V(\mathcal {I}) = \mathrm {Spec \, } R/\mathcal {I}$
, where

and

Over the special fiber (
$\pi =0$
), we have
$ V(\mathcal {I} _s) \cong \mathrm {Spec \, } \bar {R}/\mathcal {I}_s $
, where

It has two smooth irreducible components
$V(\mathcal {I}_i) = \mathrm {Spec \, } \bar {R}/\mathcal {I}_i$
, where

that are isomorphic to
$\mathbb {A}_k^{2n-4}$
. Their intersection is also smooth and of dimension
$2n-5$
. Next, we prove that
$V(\mathcal {I} _s)$
is reduced by showing that
$\mathcal {I}_s =\mathcal {I}_1 \cap \mathcal {I}_2.$
Clearly,
$\mathcal {I} _s \subset \mathcal {I}_1 \cap \mathcal {I}_2$
. Let
$g \in \mathcal {I}_1 \cap \mathcal {I}_2 .$
Thus,
$g \in \mathcal {I}_1$
and
$g = f_1 d$
for
$f_1 \in \bar {R}$
. Also,
$g \in \mathcal {I}_2$
and so
$f_1 d \in \mathcal {I}_2$
.
$\mathcal {I}_2$
is a prime ideal and
$d \notin \mathcal {I}_2$
. Thus,
$f_1 \in \mathcal {I}_2$
and

for
$f_2 \in \bar {R}$
. Thus,
$g \in \mathcal {I}_s$
and so
$\mathcal {I}_s = \mathcal {I}_1 \cap \mathcal {I}_2 .$
Lastly, we can easily see that the ideals
$ \mathcal {I}_1,\mathcal {I}_2$
are principal over
$V(\mathcal {I})$
. From the above, we deduce that
$V(\mathcal {I}) $
is regular (see [Reference Hartl10, Remark 1.1.1]).
In Case 2, we set
$\Delta =\{1,\dots ,n-2\}/ \{m, m+1\}$
and let
$i_0=m, j_0=m+1$
. Then
$V(I)$
is isomorphic to
$V(\mathcal {J}) = \mathrm {Spec \, } R'/\mathcal {J}$
, where
$R' = O_F[d,(x_i)_{i \in \Delta }, (y_i)_{i \in \Delta }]$
and

To see that
$V(\mathcal {J}) $
has semi-stable reduction over
$O_F$
, one proceeds exactly as in Case 1. Over the special fiber (
$\pi =0$
), we have
$ V(\mathcal {J} _s) \cong \mathrm {Spec \, } \bar {R}'/\mathcal {J}_s $
where
$ \bar {R} = k[d,(x_i)_{i \in \Delta }, (y_i)_{i \in \Delta }]$
and

It has two smooth irreducible components
$V(\mathcal {J}_i) = \mathrm {Spec \, } R/\mathcal {J}_i$
, where

of dimension
$2n-4$
. (Note that in this case, the second component
$V(\mathcal {J}_2)$
is not isomorphic to the affine space
$\mathbb {A}_k^{2n-4}$
as in Case 1.) Their intersection is also smooth and of dimension
$2n-5$
. By using similar arguments as Case 1, we can show that
$V(\mathcal {J}_s)$
is reduced and
$V(\mathcal {J})$
is regular.
Since
$_1{\mathcal U}_{n-2,2}$
has semi-stable reduction, and is therefore flat over
$O_F$
, we have
$_1{\mathcal U}_{n-2,2}\subset \mathrm {M}^{\mathrm {spl}}$
. Our next goal is to prove that
$\mathrm { M}^{\mathrm {spl}}$
has semi-stable reduction over
$O_F$
for the signature
$(r,s)= (n-2,2)$
. From §4.1, it is enough to show that
$\mathcal {G}$
-translates of
$_1{\mathcal U}_{n-2,2}$
cover
$\mathrm {M}^{\mathrm {spl}}$
(see Theorem 5.4).
Assume that s is even. Recall from §4.1 the
${\mathcal G}$
-equivariant morphism
$\tau : \mathrm {M}^{\mathrm {spl}} \rightarrow \mathrm {M}^{\mathrm {loc}}\otimes _{O} O_F$
which is given by
$(\mathcal {F}_0,\mathcal {F}_1) \mapsto \mathcal {F}_1$
. Let
$\tau _s : \mathrm {M}^{\mathrm {spl}}\otimes k \rightarrow \mathrm {M}^{\mathrm {loc}}\otimes k$
be the morphism over the special fiber. Let
$(\mathcal {F}_0,\mathcal {F}_1)$
be point of the special fiber of
$\mathrm {M}^{\mathrm {spl}}$
. Over the special fiber the condition (4) amounts to
$t\mathcal {F}_0 = (0)$
and so
$ (0) \subset \mathcal {F}_0 \subset t \Lambda _m \otimes k$
(see §3 for the definition of the naive splitting models). Hence, we consider the morphism

given by
$ (\mathcal {F}_0,\mathcal {F}_1) \mapsto \mathcal {F}_0.$
Here,
$\mathrm {Gr}(s,n)\otimes k$
is the finite-dimensional Grassmannian of s-dimensional subspaces of
$t \Lambda _m \otimes k.$
This has a section

given by
$ \mathcal {F}_0 \mapsto (\mathcal {F}_0,\mathcal {F}_1)$
with
$\mathcal {F}_1 = t\Lambda _m \otimes k.$
The image of the section
$\phi $
is an irreducible component of
$ \mathrm {M}^{\mathrm { spl}}\otimes _{O_F} k$
which is the fiber
$\tau ^{-1}_s (t\Lambda _m\otimes k)$
over the ‘worst point’ of
$\mathrm {M}^{\mathrm {loc}}\otimes k$
– that is, the unique closed
${\mathcal G}$
-orbit supported in the special fiber (see [Reference Pappas and Rapoport19, §5]). Hence,
$\tau ^{-1}_s(t\Lambda _m)$
is isomorphic to the Grassmannian
$ \mathrm {Gr}(s,n)\otimes k $
of dimension
$s(n-s)$
.
We equip
$t \Lambda _m \otimes k$
with the nondegenerate alternating form
$ \langle \, , \, \rangle '$
which is defined as
$\langle tv,tw \rangle ' := ( v,tw ) = v^tJ_n w$
. Note that this is well defined (see [Reference Richarz25, §5.2] for more details). Denote by
$\mathcal {Q}(s,n)$
the closed subscheme of
$\phi (\mathrm {Gr}(s,n)\otimes k) $
of dimension
$ s(2n-3s+1)/2$
which parametrizes all the isotropic s-subspaces
$\mathcal {F}_0$
in the n-space
$t\Lambda _m \otimes k$
.
Next, we give a couple of useful observations concerning the special fiber of the affine chart
$_1{\mathcal U}_{r,s}$
(see §4.1 for the notation), which will be used in the proof of the following theorem.
Remark 5.3. Recall from Theorem 4.2 that over the special fiber of
$_1{\mathcal U}_{r,s}$
, we have
$D=-D^t$
,
$D\cdot Q=0$
and the minor relations from (4.1.1).
a) Observe that if
$D =0$
, then
$Z=0$
, which gives
$A=0$
since
$A = Y Z^t$
over the special fiber. If
$A=0$
, then
$\mathcal {F}_1 = t \Lambda _m$
. Hence,
$_1{\mathcal U}_{r,s} \cap \tau ^{-1}(t\Lambda _m)$
is a smooth irreducible component of dimension
$s(n-s)$
.
b) Recall that

where
$ \mathbf {y}_i$
are the columns of the matrix
$Y.$
By direct calculations, we get that the entries
$q_{i,j}$
of the matrix Q are given by
$q_{i,j} = \langle \mathbf {y}_i,\mathbf {y}_j \rangle '$
. Thus, over the special fiber, we have
$\langle \mathcal {F}_0,\mathcal {F}_0 \rangle ' = Q $
, and so from the rank of
$ Q$
, we read how isotropic
$\mathcal {F}_0$
is with respect to
$\langle \,,\,\rangle '$
. When the rank of the matrix Q is zero, which actually occurs, we have
$\langle \mathcal {F}_0,\mathcal {F}_0 \rangle ' = 0 $
.
c) From (a) and (b), we can easily see that
$_1{\mathcal U}_{r,s}$
contains points
$(\mathcal {F}_0, \mathcal {F}_1)$
where
$ \mathcal {F}_0 \in \mathcal {Q}(s, n)$
and
$\mathcal {F}_1 = t\Lambda _m$
.
Theorem 5.4. When
$(r,s)=(n-2,2)$
,
$\mathcal {G}$
-translates of
$_1{\mathcal U}_{n-2,2}$
cover
$\mathrm {M}^{\mathrm {spl}}$
.
Proof. From §5.1, we have the forgetful
${\mathcal G}$
-equivariant morphism
$\tau : \mathrm {M}^{\mathrm {spl}} \rightarrow \mathrm {M}^{\mathrm {loc}}\otimes _{O} O_F$
given by
$(\mathcal {F}_0,\mathcal {F}_1) \mapsto \mathcal {F}_1$
. As in [Reference Pappas15, §4] and [Reference Pappas and Rapoport18, §2.4.2, 5.5], the worst point of
$\mathrm {M}^{\mathrm {loc}}\otimes _{O} O_F$
is given by the k-valued point
$t\Lambda _m\otimes k$
. From the construction of splitting models and local models, in order to show that
$\mathcal {G}$
-translates of
$_1{\mathcal U}_{n-2,2}$
cover
$\mathrm {M}^{\mathrm {spl}}$
, it is enough to prove that
$\mathcal {G}$
-translates of
$_1{\mathcal U}_{n-2,2}$
cover
$\tau ^{-1}(t\Lambda _m)$
.
Recall that
$ \tau ^{-1}_s(t\Lambda _m\otimes k) \cong \mathrm {Gr}(2,n)\otimes k$
and
$\mathcal {G}_k$
acts via its action by reduction to
$t\Lambda _m\otimes k/t^2\Lambda _m\otimes k$
. This action factors through the symplectic group
$Sp(n)_k$
of the symplectic form
$\langle \,,\,\rangle '$
on
$ t \Lambda _m\otimes k $
and gives the map
$ \mathcal {G}_k \rightarrow Sp(n)_k$
. As in [Reference Pappas and Rapoport19, §4],
$\mathcal {G}_k$
has
$Sp(n)_k$
as its maximal reductive quotient. Therefore, the map
$\mathcal {G}_k \rightarrow Sp(n)_k$
is surjective.
Next, the
$Sp(n)_k$
-action on
$\mathrm {Gr}(2, n)$
has two orbits. More precisely, the orbits are

and

Observe that
$Q(2)$
is contained in the (Zariski) closure of
$Q(0)$
, and so
$Q(2) = \mathcal {Q}(2,n)$
is the unique closed orbit (see, for example, [Reference Barbasch and Evens3, §3.1] and [Reference Arbarello, Cornalba, Griffiths and Harris1]). Thus,
$\mathcal {Q}(2,n)$
is contained in the closure of each orbit
$Q(i)$
,
$i=0,2$
.
Lastly, from Remark 5.3, we have that
$_1{\mathcal U}_{n-2,2}\otimes k$
contains points
$(\mathcal {F}_0,t\Lambda _m)$
with
$\mathcal {F}_0 \in \mathcal {Q}(2,n) $
and so
$_1{\mathcal U}_{n-2,2}$
contains points from all the orbits. Therefore, from all the above, we deduce that the
$\mathcal {G}$
-translates of
$_1{\mathcal U}_{n-2,2}$
cover
$\tau ^{-1}(t\Lambda _m)$
.
Corollary 5.5. When
$(r,s)=(n-2,2)$
, the scheme
$\mathrm {M}^{\mathrm {spl}}$
has semi-stable reduction. In particular,
$\mathrm {M}^{\mathrm {spl}}$
is regular and has special fiber a reduced divisor with two smooth irreducible components intersecting transversely.
Remark 5.6. Note that the construction of
$\phi : \mathrm {Gr}(s,n)\otimes k\cong \tau ^{-1}_s(t\Lambda _m\otimes k)$
and the observations in Remark 5.3 are true for any even s. In fact, for any even s, if we could show that the affine chart
$_1{\mathcal U}_{r,s}$
is flat, then
$_1{\mathcal U}_{r,s}\subset \mathrm {M}^{\mathrm {spl}}$
. By using a similar argument as above, we then could prove that
${\mathcal G}$
-translates of
$_1{\mathcal U}_{r,s}$
cover
$\mathrm {M}^{\mathrm {spl}}$
.
5.3 The Case
$(r,s)=(n-1,1)$
and
$(r,s)=(n-3,3)$
By using the explicit description of
$_2{\mathcal U}_{n-3,3}$
in 4.2, we show
Proposition 5.7. When
$(r,s)=(n-3,3)$
, the affine chart
$_2\mathcal {U}_{n-3,3}$
has semi-stable reduction over
$O_F$
. In particular,
$_2\mathcal {U}_{n-3,3}$
is regular and has special fiber a reduced divisor with two smooth irreducible components intersecting transversely.
Proof. Recall from §4.2.2 that the quadratic polynomial
$\sum _{k=1}^{m-1}(a_{k,2}a_{n-1-k,3}-a_{n-1-k,2}a_{k,3})$
depends on the positions of
$i_2, i_3$
. By symmetry, we only need to consider the case when
$1\leq i_2, i_3\leq m-1$
and the case when
$1\leq i_2\leq m-1, m\leq i_3\leq n-2$
.
We consider these two cases separately in this proof. In the first case, for simplicity, we can set
$i_2=1, i_3=2$
. Thus,
$_2\mathcal {U}_{n-3,3}$
is isomorphic to
$V(\mathcal {I})=\mathrm {Spec \, } R/\mathcal {I}$
, where

and

It is easy to see that we have two irreducible components in the special fiber, where

Both irreducible components are isomorphic to
$\mathbb {A}_k^{3(n-3)}$
, so they are smooth with the correct dimension. Their intersection is isomorphic to
$\mathbb {A}_k^{3(n-3)-1}$
, which is also smooth. Since
$I_1, I_2$
are prime ideals, it is enough to prove that the special fiber
$V(\mathcal {I}_s)$
is reduced by showing
$\mathcal {I}_s=I_1\cap I_2$
. By using similar arguments as in Proposition 5.2, we get that
$V(\mathcal {I}_s)$
is reduced and
$V(\mathcal {I})$
is regular.
In the second case, we can set
$\Delta =\{1,\dots ,n-2\}/ \{m-1, m\}$
, and let
$i_2=m-1, i_3=m$
. Then
$_2\mathcal {U}_{n-3,3}$
is isomorphic to
$V(\mathcal {J})=\mathrm {Spec \, } R'/\mathcal {J}$
, where

and

Again, we have two irreducible components over the special fiber:

Both of them are smooth and of dimension
$3(n-3)$
(Note that in this case, the second component
$V(\mathcal {J}_2)$
is not isomorphic to
$\mathbb {A}_k^{3(n-3)}$
). Their intersection is also smooth and of dimension
$3n-10$
. By using the same arguments as above, we can show that
$V(\mathcal {J}_s)$
is reduced and
$V (\mathcal {J} )$
is regular. Thus,
$_2\mathcal {U}_{n-3,3}$
has semi-stable reduction over
$O_F$
.
From Theorem 4.7 and Proposition 5.7, we have that
$_2{\mathcal U}_{n-1,1}$
is smooth and
$_2{\mathcal U}_{n-3,3}$
has semi-stable reduction. Thus, both
$ _2{\mathcal U}_{n-1,1}$
and
$_2{\mathcal U}_{n-3,3} $
are
$O_F$
-flat. As in §5.2, we deduce that these affine charts are open subschemes of the corresponding splitting model
$\mathrm {M}^{\mathrm { spl}}$
. We will prove that
$\mathrm {M}^{\mathrm {spl}}$
is smooth when
$(r,s)=(n-1,1)$
and
$\mathrm {M}^{\mathrm {spl}}$
has semi-stable reduction when
$(r,s)=(n-3,3)$
by showing that
${\mathcal G}$
-translates of
$_2{\mathcal U}_{n-1,1}$
and
$_2{\mathcal U}_{n-3,3}$
cover
$\mathrm {M}^{\mathrm {spl}}$
.
Recall from §3 the
${\mathcal G}$
-equivariant morphism
$\tau : \mathrm {M}^{\mathrm {spl}}\rightarrow \mathrm {M}^{\mathrm {loc}} \otimes _{O}O_F$
which is given by
$(\mathcal {F}_0,\mathcal {F}_1)\mapsto \mathcal {F}_1$
and let
$\tau _s: \mathrm {M}^{\mathrm {spl}}\otimes k\rightarrow \mathrm {M}^{\mathrm {loc}} \otimes k$
be the morphism over the special fiber. When s is odd, the unique closed
${\mathcal G}$
-orbit of
$\mathrm {M}^{\mathrm {loc}} \otimes k$
is the orbit of
$\Lambda '\otimes k$
(see §3). From the construction of the local model, it is enough to show that
${\mathcal G}_k$
-translates of
$_2\mathcal {U}_{n-1,1}\otimes k$
and
$_2\mathcal {U}_{n-3,3}\otimes k$
cover
$\tau _s^{-1}(\Lambda '\otimes k)$
.
When
$(r,s)=(n-1,1)$
, the inverse image
$\tau _s^{-1}(\Lambda '\otimes k)$
contains points
$(\mathcal {F}_0,\Lambda '\otimes k)$
satisfying
$t\Lambda '\subset \mathcal {F}_0$
. Observe that
$t\Lambda '=\text {span}_{k}\{-e_1\}$
and
$\mathcal {F}_0$
has rank
$1$
. Thus,
$\tau _s^{-1}(\Lambda '\otimes k)=(\text {span}_{k}\{-e_1\}, \Lambda '\otimes k)$
is just a unique point (Note that it is different from the case when s is even, see §5.2, where we have
$t\Lambda _m=0$
). From §4.2, we see that
$_2\mathcal {U}_{n-1,1}\otimes k$
contains that point and so
${\mathcal G}_k$
-translates of
$_2\mathcal {U}_{n-1,1}\otimes k$
cover
$\tau _s^{-1}(\Lambda '\otimes k)$
. Hence,
${\mathcal G}$
-translates of
$_2\mathcal {U}_{n-1,1}$
cover
$\mathrm {M}^{\mathrm {spl}}$
.
When s is odd and
$\geq 3$
, we have

Set
$\mathcal {F}_0=\text {span}_k\{e_1, v_2, \dots , v_s\}$
. Then
$t\mathcal {F}_0=0$
and
$\mathcal {F}_0\subset \Lambda '\otimes k$
imply that
$v_2, \dots , v_s$
are a linear combination of
$\{-e_2,\dots , -e_m, \pi e_{m+1},\dots , \pi e_{n-1}\}$
. Thus, we consider the morphism

given by
$(\mathcal {F}_0,\Lambda ')\mapsto \text {span}_k\{v_2, \cdots , v_s\}$
. Here,
$\mathrm {Gr}(s-1,n-2)\otimes k$
is the Grassmannian of
$(s-1)$
-dimensional subspaces of

We also have a section

given by
$W\rightarrow (\text {span}_k(e_1, W),\Lambda ')$
. Similar to §5.2, it is easy to see that the image of the section
$\phi $
is
$\tau _s^{-1}(\Lambda '\otimes k)$
. Hence,
$\tau _s^{-1}(\Lambda '\otimes k)$
is isomorphic to
$\mathrm {Gr}(s-1,n-2)\otimes k$
. As in §5.2, we can equip
$\tau _s^{-1}(\Lambda '\otimes k)$
with a nondegenerate alternating form; that is,
$\langle tv,tw\rangle ':=(tv,w)=v^tJ_{n-2}w$
for
$tv, tw\in \tau _s^{-1}(\Lambda '\otimes k)$
. Denote by
$\mathcal {Q}(s-1,n-2)$
the closed subscheme of
$\mathrm {Gr}(s-1,n-2)\otimes k$
which parametrizes all isotropic
$(s-1)$
-subspaces with respect to
$\langle \ , \ \rangle '$
.
The following observations concern the special fiber of
$_2{\mathcal U}_{r,s}$
when s is odd. We keep the same notation as in §4.2.
Remark 5.8. (1) Set
$\mathbf {a}_i$
the i-th column of the matrix
$M_2$
;
$\mathbf {a}_i$
can be expressed as the element
$tv\in \tau _s^{-1}(\Lambda '\otimes k)$
, where
$tv=[ 0~0~a_{1,i}~\cdots ~a_{n-2,i}\mid ~0~\cdots ~ 0]^t$
. Thus,

(2) For the special fiber
$\mathrm {M}^{\mathrm {spl}}\otimes k$
, consider
$(_2\mathcal {U}_{r,s}\otimes k) \cap \tau _s^{-1}(\Lambda '\otimes k)$
. We get
$C=0$
since
$A=0$
. The first column of C is
$\mathbf {a}_1+z_{1,2}\mathbf {a}_2+\cdots +z_{1,s}\mathbf {a}_s$
. By setting
$i=i_2, \dots , i_s$
, this gives us
$z_{1,2}=\cdots =z_{1,s}=0$
and so
$\mathbf {a}_1=\mathbf {0}$
. Thus,

where

By definition,
$Q=[q_{i,j}]=0$
is equivalent to
$\langle \mathbf {a}_i,\mathbf {a}_j \rangle '=0$
for
$2\leq i, j\leq s$
. Therefore,
$\pi (\mathcal {F}_0)$
is totally isotropic under
$\langle ~,~ \rangle '$
if and only if
$Q=0$
(which actually occurs). Thus,

Theorem 5.9. When
$(r,s)=(n-3,3)$
,
${\mathcal G}$
-translates of
$_2\mathcal {U}_{n-3,3}$
cover
$\mathrm {M}^{\mathrm {spl}}$
.
Proof. It is enough to prove that
${\mathcal G}$
-translates of
$_2\mathcal {U}_{n-3,3}$
cover
$\tau ^{-1}(\Lambda ')$
. We use similar arguments as in the proof of Theorem 5.4. Since
$\mathcal {G}_k$
acts on
$\tau _s^{-1}(\Lambda '\otimes k)\simeq \mathrm {Gr}(2,n-2)\otimes k$
, this action factors through the symplectic group
$Sp(n-2)_k$
of the symplectic form
$\langle \,,\,\rangle '$
on
$ \Lambda '/ \{-\pi ^{-1}e_1, -e_1\} $
, and gives the surjective map
$ \mathcal {G}_k \rightarrow Sp(n-2)_k$
, where

By [Reference Barbasch and Evens3, §3], the
$Sp(n-2)_k$
-action on
$\mathrm {Gr}(2,n-2)\otimes k$
has two orbits:
$Q(0)$
and
$Q(2)$
, where

for
$i=0, 2$
. Similar to the proof of Theorem 5.4, we have
$Q(2)\subset \overline {Q(0)}$
. Thus,
$Q(2)$
is the unique closed orbit. Observe that
$Q(2)$
contains all totally isotropic subspaces
$\pi (\mathcal {F}_0)$
, and so
$Q(2)= \mathcal {Q}(2,n-2)$
.
From Remark 5.8 (2), we have that
$_2{\mathcal U}_{n-3,3}\otimes k$
contains points
$(\mathcal {F}_0,\Lambda '\otimes k)$
with
$\pi (\mathcal {F}_0) \in \mathcal {Q}(2,n-2) $
. By identifying
$\tau _s^{-1}(\Lambda '\otimes k)$
with
$\mathrm {Gr}(2,n-2)\otimes k$
, we get that
$_2{\mathcal U}_{n-3,3}$
contains points from all the orbits. Therefore, from all the above, we deduce that the
$\mathcal {G}$
-translates of
$_2{\mathcal U}_{n-3,3}$
cover
$\tau ^{-1}(\Lambda ')$
.
From Propositions 4.7 and 5.7 and Theorem 5.9, we have the following:
Corollary 5.10.
-
a) When
$(r,s)=(n-1,1)$ ,
$\mathrm {M}^{\mathrm {spl}}$ is a smooth scheme.
-
b) When
$(r,s)=(n-3,3)$ ,
$\mathrm {M}^{\mathrm {spl}}$ has semi-stable reduction over
$O_F$ . In particular,
$\mathrm {M}^{\mathrm {spl}}$ is regular and has special fiber a reduced divisor with two smooth irreducible components intersecting transversely.
Remark 5.11. For s odd and
$\geq 5$
, note that we have

and

If we can show that the affine chart
$_2{\mathcal U}_{r,s}$
is flat, then we have
$_2{\mathcal U}_{r,s}\subset \mathrm {M}^{\mathrm {spl}}$
. By using a similar argument as the proof of Theorem 5.9, we could prove that
${\mathcal G}$
-translates of
$_2{\mathcal U}_{r,s}$
cover
$\mathrm {M}^{\mathrm {spl}}$
.
Remark 5.12. When
$ (r,s) = (n-1,1)$
, the local model
$ \mathrm {M}^{\mathrm {loc}} \subset \mathrm {M}^{\wedge }$
is smooth (see [Reference Pappas and Rapoport18, §5.3]). Consider the morphism
$\tau : \mathrm {M}^{\mathrm {spl}} \longrightarrow \mathrm {M}^{\mathrm {loc}} \otimes _{O} O_F$
. The generic fiber of all of these is the same (see §3). In particular, from all the above discussion in this section, we can see that
$ \mathrm {M}^{\mathrm {spl}} \cong \mathrm {M}^{\mathrm {loc}} \otimes _{O} O_{F}$
.
6 Application to Shimura varieties
In this section, we give the most immediate application of Corollaries 5.5 and 5.10 to unitary Shimura varieties. Indeed, by using work of Rapoport and Zink [Reference Rapoport and Zink24], we first construct p-adic integral models for the signatures
$ (n-s,s)$
with
$ s\leq 3$
where the level subgroup at p is the special maximal parahoric subgroup
$P_{\{m\}}$
defined in §2.2. These models have simple and explicit moduli descriptions and are étale locally around each point isomorphic to naive local models. Then by using the above corollaries and producing a linear modification, we obtain smooth integral models of the corresponding Shimura varieties when
$s=1$
and semi-stable models when
$s=2$
or
$s=3$
(i.e., they are regular and the irreducible components of the special fiber are smooth divisors crossing normally). (See Theorem 6.1.)
We start with an imaginary quadratic field K and we fix an embedding
$\varepsilon : K\rightarrow \mathbb {C}$
. Let W be a n-dimensional K-vector space, equipped with a nondegenerate hermitian form
$\phi $
. Consider the group
$G = \mathrm {GU}_n$
of unitary similitudes for
$(W,\phi )$
of dimension
$n\geq 3$
over K. Assume that
$n=2m$
is even. We fix a conjugacy class of homomorphisms
$h: \mathrm {Res}_{\mathbb {C}/\mathbb {R}}\mathbb {G}_{m,\mathbb {C}}\rightarrow \mathrm {GU}_n$
corresponding to a Shimura datum
$(G,X) = (\mathrm {GU}_n,X_h)$
of signature
$(r,s)$
. The pair
$(G,X)$
gives rise to a Shimura variety
$\mathrm {Sh}(G,X)$
over the reflex field E. (See [Reference Pappas and Rapoport18, §1.1] and [Reference Pappas15, §3] for more details on the description of the unitary Shimura varieties.) Let p be an odd prime number which ramifies in K. Set
$K_1=K\otimes _{\mathbb {Q}} \mathbb {Q}_p$
with a uniformizer
$\pi $
, and
$V=W\otimes _{\mathbb {Q}} \mathbb {Q}_p$
. We assume that the hermitian form
$\phi $
is split on V (i.e., there is a basis
$e_1, \dots , e_n$
such that
$\phi (e_i, e_{n+1-j})=\delta _{ij}$
for
$1\leq i, j\leq n$
). We denote by

the standard lattices in V. We can complete this into a self dual lattice chain by setting
$\Lambda _{i+kn}:=\pi ^{-k}\Lambda _i$
(see §2.1). Denote by
$P_{\{m\}}$
the stabilizer of
$\Lambda _m$
in
$G(\mathbb {Q}_p)$
. We let
$\mathcal {L}$
be the self-dual multichain consisting of lattices
$\{\Lambda _j\}_{j\in n\mathbb {Z} \pm m}$
. Here,
$\mathcal {G} = \underline {\mathrm {Aut}}(\mathcal {L})$
is the (smooth) group scheme over
$\mathbb {Z}_p$
with
$P_{\{m\}} = \mathcal {G}(\mathbb {Z}_p)$
the subgroup of
$G(\mathbb {Q}_p)$
fixing the lattice chain
$\mathcal {L}$
.
Choose also a sufficiently small compact open subgroup
$K^p$
of the prime-to-p finite adelic points
$G({\mathbb A}_{f}^p)$
of G and set
$\mathbf {K}=K^pP_{\{m\}}$
. The Shimura variety
$\mathrm { Sh}_{\mathbf {K}}(G, X)$
with complex points

is of PEL type and has a canonical model over the reflex field E. We set
$\mathcal {O} = O_{E_v}$
, where v the unique prime ideal of E above
$(p)$
.
We consider the moduli functor
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
over
$\mathrm {Spec \, } \mathcal {O} $
given in [Reference Rapoport and Zink24, Definition 6.9]:
A point of
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
with values in the
$\mathcal {O} $
-scheme S is the isomorphism class of the following set of data
$(A,\iota ,\bar {\lambda }, \bar {\eta })$
:
-
(1) An object
$(A,\iota )$ , where A is an abelian scheme with relative dimension n over S (terminology of [Reference Rapoport and Zink24]), compatibly endowed with an action of
${\mathcal O}$ :
$$\begin{align*}\iota: {\mathcal O} \rightarrow \text{End} \,A \otimes \mathbb{Z}_p.\end{align*}$$
-
(2) A
$\mathbb {Q}$ -homogeneous principal polarization
$\bar {\lambda }$ of the
$\mathcal {L}$ -set A.
-
(3) A
$K^p$ -level structure
$$\begin{align*}\bar{\eta} : H_1 (A, {\mathbb A}_{f}^p) \simeq W \otimes {\mathbb A}_{f}^p \, \text{ mod} \, K^p \end{align*}$$
$({\mathbb A}_{f}^p)^{\times }$ (see loc. cit. for details).
The set A should satisfy the determinant condition (i) of loc. cit.
For the definitions of the terms employed here, we refer to loc.cit., 6.3–6.8 and [Reference Pappas15, §3]. The functor
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
is representable by a quasi-projective scheme over
$\mathcal {O}$
. Since the Hasse principle is satisfied for the unitary group, we can see as in loc. cit. that there is a natural isomorphism

As is explained in [Reference Rapoport and Zink24] and [Reference Pappas15], the naive local model
$\mathrm {M}^{\mathrm {naive}}$
is connected to the moduli scheme
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
via the local model diagram

where the morphism
$\psi _1$
is a
$\mathcal {G}$
-torsor and
$\psi _2$
is a smooth and
$\mathcal {G}$
-equivariant morphism. Therefore, there is a relatively representable smooth morphism

where the target is the quotient algebraic stack.
Next, denote by
$ \mathcal {A}^{\mathrm {flat}}_{\mathbf {K}} $
the flat closure of
$\mathrm {Sh}_{\mathbf {K}}(G, X)\otimes _{E} E_v$
in
$ \mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}$
. Recall from §3 that the flat closure of
$ \mathrm {M}^{\mathrm {naive}} \otimes _{\mathcal {O}} E_v$
in
$\mathrm {M}^{\mathrm {naive}}$
is by definition the local model
$\mathrm {M}^{\mathrm {loc}} $
. By the above, we can see, as in [Reference Pappas and Rapoport18], that there is a relatively representable smooth morphism of relative dimension
$\mathrm {dim} (G)$
,

This of course implies that
$\mathcal {A}^{\mathrm {flat}}_{\mathbf {K}}$
is étale locally isomorphic to the local model
$\mathrm {M}^{\mathrm {loc}}$
.
One can now consider a variation of the moduli of abelian schemes
$\mathcal {A}_{\mathbf {K}}$
where we add in the moduli problem an additional subspace in the Hodge filtration
$ \mathrm {Fil}^0 (A) \subset H_{dR}^1(A)$
of the universal abelian variety A (see [Reference Haines7, §6.3] for more details) with certain conditions to imitate the definition of the naive splitting model
$\mathcal {M}$
. (See §3 for the definition of naive splitting models.) More precisely,
$\mathcal {A}_{\mathbf {K}}$
associates to an
$O_{K_1}$
-scheme S the set of isomorphism classes of objects
$(A,\iota ,\bar {\lambda }, \bar {\eta },\mathscr {F}_0) $
. Here,
$(A,\iota ,\bar {\lambda }, \bar {\eta })$
is an object of
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}(S).$
Set
$\mathscr {F}_1 := \mathrm {Fil}^0 (A) $
. The final ingredient
$\mathscr {F}_0$
of an object of
$\mathcal {A}_{\mathbf {K}}$
is the subspace
$ \mathscr {F}_0 \subset \mathscr {F}_1 \subset H_{dR}^1(A) $
of rank s which satisfies the following conditions:

There is a forgetful projective morphism

defined by
$(A,\iota ,\bar {\lambda }, \bar {\eta },\mathscr {F}_0) \mapsto (A,\iota ,\bar {\lambda }, \bar {\eta }) $
. Moreover,
$\mathcal {A}_{\mathbf {K}}$
has the same étale local structure as
$\mathcal {M}$
; it is a ‘linear modification’ of
$\mathcal {A}^{\mathrm {naive}}_{\mathbf {K}}\otimes _{\mathcal {O}} O_{K_1}$
in the sense of [Reference Pappas15, §2] (see also [Reference Pappas and Rapoport17, §15]). As we showed in Proposition 5.1, the scheme
$\mathcal {M}$
is never flat, and by the above, the same is true for
$\mathcal {A}_{\mathbf {K}}$
.
Theorem 6.1. Assume that
$ (r,s) = (n-1,1)$
or
$(n-2,2)$
or
$(n-3,3)$
. For every
$K^p$
as above, there is a scheme
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$
, flat over
$\mathrm {Spec \, } (O_{K_1})$
, with

and which supports a local model diagram

such that
-
a)
$\pi ^{\mathrm {reg}}_{\mathbf {K}}$ is a
$\mathcal {G}$ -torsor for the parahoric group scheme
${\mathcal G}$ that corresponds to
$P_{\{m\}}$ .
-
b)
$q^{\mathrm {reg}}_{\mathbf {K}}$ is smooth and
${\mathcal G}$ -equivariant.
-
c) When
$(r,s) = (n-1,1)$ ,
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$ is a smooth scheme.
-
c’) When
$(r,s) = (n-2,2)$ or
$(r,s) =(n-3,3)$ ,
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$ is regular and has special fiber which is a reduced divisor with normal crossings.
Proof. From the above, we have the local model diagram

where the morphism
$\psi _1$
is a
$\mathcal {G}$
-torsor and
$\psi _2$
is a smooth and
$\mathcal {G}$
-equivariant morphism. Set

which carries a diagonal
$\mathcal {G}$
-action. Since
$\tau : \mathrm {M}^{\mathrm {spl}} \rightarrow \mathrm {M}^{\mathrm {loc}}\otimes _{O} O_F$
is projective (see §5), we can see ([Reference Pappas15, §2]) that the quotient

is represented by a scheme and gives a
${\mathcal G}$
-torsor. (Recall that the morphism
$\tau $
induces an isomorphism on the generic fibers.) This is an example of a linear modification; see [Reference Pappas15, §2]. The projection gives a smooth
${\mathcal G}$
-morphism

which completes the local model diagram. Property (c) (resp. (c’)) follows from Theorem 5.10 (resp. Corollaries 5.5 and 5.10) and properties (a) and (b) which imply that
$ \mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$
and
$ \mathrm {M}^{\mathrm {spl}}$
are locally isomorphic for the étale topology.
Remarks 6.2. Similar results can be obtained for corresponding Rapoport-Zink formal schemes. (See [Reference He, Pappas and Rapoport12, §4] for an example of this parallel treatment.)
7 Moduli description of
$\mathrm {M}^{\mathrm {spl}}$
In this section, we show that by adding the ‘spin condition’ in the naive splitting model
$\mathcal {M}$
, we get the splitting model
$\mathrm {M}^{\mathrm {spl}}$
for signature
$s\leq 3$
.
We use the notation of §2, §3, and we set
$ W= \wedge ^{n}_F (V\otimes _{F_0}F).$
Recall that the symmetric form
$( \, , \, )$
splits over V, and thus, there is a canonical decomposition

of
$W $
as an
$\mathrm {SO}_{2n} \left ((\, ,\,)\right )(F)$
-representation (see [Reference Rapoport, Smithling and Zhang23, §7], [Reference Smithling27, §2]). For a standard lattice
$\Lambda _i$
in V (see §2), we set
$W(\Lambda _i)=\wedge ^n(\Lambda _i\otimes _{O_{F_0}}O_F )$
and
$W(\Lambda _i)_{\pm 1 } = W_{\pm 1} \cap W(\Lambda ).$
For an
$O_F$
-scheme S and
$\epsilon \in \{\pm 1\}$
, we set

We now formulate the spin condition:
(5) (Spin condition) the line bundle
$\wedge ^n \mathcal {F}_1 \subset W(\Lambda _m)\otimes _{O_F} \mathcal {O}_S$
is contained in
$ L_{m,(-1)^s}(S)$
.
The spin splitting model
$\mathcal {M}^{\mathrm {spin}}$
is the closed subscheme of
$\mathcal {M}$
defined by imposing the spin condition. We have the following inclusions of closed subschemes
$\mathrm { M}^{\mathrm {spl}}\subset \mathcal {M}^{\mathrm {spin}}\subset \mathcal {M}$
which are all equalities between generic fibers (see [Reference Pappas and Rapoport18, §7] and [Reference Rapoport, Smithling and Zhang23, §9]). The restriction of
$\tau $
on
$\mathcal {M}^{\mathrm {spin}}$
gives us

where
$\mathrm {M}^{\mathrm {spin}}$
is the spin local model defined in [Reference Pappas and Rapoport18]. In particular,
$\mathrm {M}^{\mathrm {spin}}$
is the closed subscheme of
$\mathrm {M}^{\wedge }$
that classifies points given by
$\mathcal {F}_1$
which satisfy the spin condition. We want to mention that Smithling [Reference Smithling26, Theorem 1.3] proved that
$\mathrm {M}^{\mathrm {spin}}$
is topologically flat. Also, from [Reference Rapoport, Smithling and Zhang23, Remark 9.9], we get that the spin condition can be characterized as the following condition:
(5’) The rank of
$ (t+ \pi )$
on
$\mathcal {F}_1$
has the same parity as s.Footnote
1
Remark 7.1. Recall from the proof of Proposition 5.1 that the point
$(\mathcal {F}_0, t\Lambda _m)$
when s is odd and the point
$(\mathcal {F}_0, \Lambda ')$
when s is even does not lift to the generic fiber. However, we can easily see that these points do not satisfy the condition (5’) and so they are not in
$\mathcal {M}^{\mathrm {spin}}$
.
Proposition 7.2.
-
a) When s is even,
$_1{\mathcal U}_{r,s}\subset {\mathcal {M}}^{\mathrm {spin}}$ .
-
b) When s is odd,
$_2{\mathcal U}_{r,s}\subset {\mathcal {M}}^{\mathrm {spin}}$ .
Proof. It is enough to consider the rank
$((t+\pi )\mathcal {F}_1)$
over the special fiber (
$\pi = 0$
) for the following two cases.
Case 1: Consider the affine chart
$_1{\mathcal U}_{r,s}$
around
$(\mathcal {F}_0, \Lambda _m)$
. Recall from §4.1 that

By
$(t + \pi ) \mathcal {F}_1 \subset \mathcal {F}_0$
, we get

where Z is of size
$n\times s$
. Since rank
$(\mathcal {F}_0)=s$
, the rank of
$(t + \pi ) \mathcal {F}_1$
amounts to rank
$(Z^t)$
. Note that
$Z^t = D^t\cdot \widetilde {Y}^t$
, where D is of size
$s\times s$
and rank
$(\widetilde {Y})=s$
. Thus, rank
$(Z^t)=$
rank
$(D^t)$
. Since D is a skew-symmetric matrix, the rank of D is always an even number. Therefore the rank of
$(t+\pi )\mathcal {F}_1$
, where
$(\mathcal {F}_0, \mathcal {F}_1)\in \, _{1}{\mathcal U}_{r,s}$
, is always even.
Case 2: Consider the affine chart
$_2{\mathcal U}_{r,s}$
around
$(\mathcal {F}_0, \Lambda ')$
. In this case, from §4.2, we have

and

where
$Z_1$
is of size
$2\times s$
and
$Z_2$
is of size
$(n-2)\times s$
. Thus, the rank of
$(t + \pi ) \mathcal {F}_1$
amounts to rank
$([Z_1^t\ Z_2^t])$
. Over the special fiber, we have

We treat
$Z_1^t, Z_2^t$
separately. By (4.2.1),
$Z_2^t$
can be expressed as

where D is a skew-symmetric matrix, and
$\tilde {M}_2$
contains an
$I_{s-1}$
-minor. Thus,
$Z_2^t$
is generated by the columns of
$\left [\begin {array}{c}0\cdots 0\\ D\end {array}\right ]$
. For
$Z_1^t$
, we claim that the second column is generated by
$\left [\begin {array}{c}0\cdots 0\\ D\end {array}\right ]$
. From (4.2.2), we have

in the special fiber. Recall that
$D\cdot Q=0$
, where

Then

which can be expressed as

Therefore, the second column is generated by
$\left [\begin {array}{c}0\cdots 0\\ D\end {array}\right ]$
. From all the above, we deduce that the rank of
$[Z_1^t\ Z_2^t]$
amounts to the rank of

Since rank(D) is even, and the first column is linear independent to the rest of the columns, the rank of
$(t+\pi )\mathcal {F}_1$
, where
$(\mathcal {F}_0, \mathcal {F}_1)\in _{2}{\mathcal U}_{r,s}$
, is always odd.
Therefore,
$_1{\mathcal U}_{r,s}\subset {\mathcal {M}}^{\mathrm {spin}}$
if and only if s is even and
$_2{\mathcal U}_{r,s}\subset {\mathcal {M}}^{\mathrm {spin}}$
if and only if s is odd.
In particular, when
$s=2$
, we have
$_{1}{\mathcal U}_{n-2,2}\subset \mathcal {M}^{\mathrm {spin}}$
, and when
$s=3$
, we have
$_{2}{\mathcal U}_{n-3,3}\subset \mathcal {M}^{\mathrm {spin}}$
. Moreover, since
$\mathrm {M}^{\mathrm {spin}}$
is topologically flat, or in other words, the underlying topological spaces of
$\mathrm {M}^{\mathrm {spin}}$
and
$\mathrm {M}^{\mathrm {loc}}$
coincide, we can repeat the
$\mathcal {G}$
-translates argument as in Theorems 5.4 and 5.9 and obtain the following.
Proposition 7.3.
-
a) When
$(r,s)=(n-2,2)$ ,
${\mathcal G}$ -translates of
$_1\mathcal {U}_{n-2,2}$ cover
$\mathcal {M}^{\mathrm {spin}}$ .
-
b) When
$(r,s)=(n-3,3)$ ,
${\mathcal G}$ -translates of
$_2\mathcal {U}_{n-3,3}$ cover
$\mathcal {M}^{\mathrm {spin}}$ .
From Theorems 5.4, 5.9 and Proposition 7.3, we deduce the following.
Theorem 7.4. For the signatures
$(n-s,s)$
with
$s\leq 3$
, we have
$\mathcal {M}^{\mathrm {spin}}=\mathrm {M}^{\mathrm {spl}}$
.
Therefore, we obtain a moduli-theoretic description of
$\mathrm {M}^{\mathrm {spl}}$
and by Theorem 6.1 for the corresponding integral model
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}}$
. In particular, the closed subscheme
$\mathcal {A}^{\mathrm {spl}}_{\mathbf {K}} \subset \mathcal {A}_{\mathbf {K}}$
(see §6 for the explicit definition of
$\mathcal {A}_{\mathbf {K}}$
) is obtained by imposing the following additional condition on the moduli problem of
$\mathcal {A}_{\mathbf {K}}$
:
Spin condition. The rank of
$(\iota (\pi ) + \pi )$
on
$\mathrm {Fil}^0(A)$
has the same parity as s.
Remark 7.5. Recall the
${\mathcal G}$
-equivariant morphism
$\tau : \mathrm {M}^{\mathrm {spl}}\rightarrow \mathrm {M}^{\mathrm {loc}}\otimes _O O_F$
, which is given by
$(\mathcal {F}_0, \mathcal {F}_1)\mapsto \mathcal {F}_1$
. By the above remark, we have
$\mathcal {M}^{\mathrm {spin}}=\mathrm {M}^{\mathrm {spl}}$
for the signatures
$(n-s,s)$
with
$s\leq 3$
. Define
$\tau _s: \mathrm {M}^{\mathrm {spl}}\otimes k\rightarrow \mathrm {M}^{\mathrm {loc}}\otimes k$
over the special fiber. For a point
$ (\mathcal {F}_0, \mathcal {F}_1) \in \mathrm {M}^{\mathrm {spl}}\otimes k$
, we have
$(0) \subset t \mathcal {F}_1\subset \mathcal {F}_0$
and
$t\mathcal {F}_0 = (0)$
which gives
$(0)\subset \mathcal {F}_0 \subset t\Lambda _m \otimes k$
. Also, we have

The spaces
$t^{-1}(\mathcal {F}_0) , \,\mathcal {F}_0^{\bot }$
have rank
$n + s$
,
$2n-s$
, respectively, and we indicate with
$ \bot $
the orthogonal complement with respect to
$(\,,\,)$
. From above, we obtain that

A) When the signature
$(r,s)=(n-2,2)$
, we have a morphism
$\pi : \mathrm {M}^{\mathrm {spl}}\otimes k\rightarrow \mathrm {Gr}(2,n)\otimes k$
given by
$(\mathcal {F}_0, \mathcal {F}_1)\mapsto \mathcal {F}_0$
. Let
$M_1=\tau _s^{-1}(t\Lambda _m\otimes k)$
, and
$M_2=\pi ^{-1}(\mathcal {Q}(2,n))$
. Recall that
$M_1$
is isomorphic to the Grassmannian
$\mathrm {Gr}(2,n)\otimes k$
of dimension
$2n-4$
. The spin condition translates to: rank(
$t\mathcal {F}_1)=$
even. Since
$0\subset t\mathcal {F}_1\subset \mathcal {F}_0$
, we get rank
$(t\mathcal {F}_1)=0$
or
$2$
, (i.e.,
$t\mathcal {F}_1=0$
or
$t\mathcal {F}_1=\mathcal {F}_0$
).
Furthermore, from
$t\mathcal {F}_1=0$
, we get
$\mathcal {F}_1\subset t\Lambda _m\otimes k$
which implies
$\mathcal {F}_1=t\Lambda _m\otimes k$
, so
$(\mathcal {F}_0, \mathcal {F}_1)\in M_1$
. However, if
$t\mathcal {F}_1=\mathcal {F}_0$
, we have rank
$(t(\mathcal {F}_1+t\Lambda _m\otimes k))=$
rank
$(t\mathcal {F}_1)=2$
. Thus, rank
$(\mathcal {F}_1+t\Lambda _m\otimes k)\geq n+2$
. By rank
$(t^{-1}(\mathcal {F}_0))=n+2$
and (7.0.1), we get
$\mathcal {F}_1+t\Lambda _m\otimes k = \mathcal {F}_0^{\bot }\cap t^{-1}(\mathcal {F}_0)= t^{-1}(\mathcal {F}_0)$
, which implies
$ t^{-1}(\mathcal {F}_0) \subset \mathcal {F}_0^{\bot }$
. As in [Reference Zachos28, §2.1], observe that
$ \mathcal {F}_0 \in \mathcal {Q}(2,n)$
(i.e.,
$ \langle \mathcal {F}_0,\mathcal {F}_0 \rangle ' = 0$
) is equivalent to
$\text {rank} \, (t^{-1}(\mathcal {F}_0) \cap \mathcal {F}^{\bot }_0) = n+2 $
(i.e.,
$ t^{-1}(\mathcal {F}_0) \subset \mathcal {F}_0^{\bot })$
. So
$(\mathcal {F}_0, \mathcal {F}_1)\in M_2$
.
Therefore,
$\mathrm {M}^{\mathrm {spl}}\otimes k = M_1 \cup M_2$
. Moreover, from the above, we can easily see that
$\tau ': M_2\setminus M_1\simeq \mathrm {M}^{\mathrm {loc}} \setminus \{t\Lambda _m\otimes k\}$
, so that
$M_2$
maps birationally to the special fiber of
$\mathrm {M}^{\mathrm {loc}}$
.
Lastly, the fiber
$\pi ^{-1}(\{\mathcal {F}_0\}) $
with
$\mathcal {F}_0 \in \mathcal {Q}(2,n) $
contains all
$(\mathcal {F}_0, \mathcal {F}_1) $
with
$\mathcal {F}_1$
satisfying
$\mathcal {F}_0 \subset (t^{-1}(\mathcal {F}_0))^{\bot } \subset \mathcal {F}_1 \subset t^{-1}(\mathcal {F}_0).$
These
$ \{\mathcal {F}_1\}$
correspond to isotropic, with respect to
$( \,,\,)$
, subspaces in the four-dimensional space
$t^{-1}(\mathcal {F}_0)/ (t^{-1}(\mathcal {F}_0))^{\bot }$
, and they are parameterized by the orthogonal Grassmannian
$ \mathrm {OGr}(2,4)$
. Therefore,
$M_2$
is a
$\mathrm {OGr}(2,4)$
-bundle over
$ \mathcal {Q}(2,n)$
with dimension
$2n-4$
, and
$M_1, M_2$
intersect transversally over the smooth scheme
$\mathcal {Q}(2,n)$
of rank
$2n-5$
.
B) When the signature
$(r,s)=(n-3,3)$
, the unique closed
$\mathcal {G}$
-orbit in
$\mathrm {M}^{\mathrm {loc}}\otimes k$
is not the point
$t\Lambda _m\otimes k$
but the orbit of
$\Lambda '\otimes k$
which we denote by
$O_{wt}$
. Let

and

The spin condition translates to rank
$(t\mathcal {F}_1)=$
odd. By
$0\subset t\mathcal {F}_1\subset \mathcal {F}_0$
, we get rank
$(t\mathcal {F}_1)=1$
or
$3$
.
If rank
$(t\mathcal {F}_1)=3$
, we have
$t\mathcal {F}_1=\mathcal {F}_0$
which implies rank
$(\mathcal {F}_1+t\Lambda _m\otimes k)\geq n+3$
. By (7.0.1), we get
$\mathcal {F}_0^{\bot }\cap t^{-1}(\mathcal {F}_0)= t^{-1}(\mathcal {F}_0)$
. Thus,
$\mathcal {F}_0$
is totally isotropic under
$\langle \ ,\ \rangle '$
. Next, assume that rank
$(t\mathcal {F}_1)=1$
, or in other words, consider

where the quotients arising from all slanted inclusions are finitely generated projective k-modules of rank 1. We can easily see that this holds for
$ \mathcal {F}_1 = \Lambda '\otimes k$
. Since
$ t \Lambda _m$
is fixed by the action of
$ {\mathcal G}$
and the
$ {\mathcal G}$
-orbit
$ O_{wt}$
is closed, we can deduce that rank
$(t\mathcal {F}_1)=1$
if and only if
$\mathcal {F}_1 \in O_{wt}$
. Therefore, as in (A), we get
$\mathrm {M}^{\mathrm {spl}}\otimes k = M_1 \cup M_2$
, where
$M_2$
maps birationally to the special fiber of
$\mathrm {M}^{\mathrm {loc}}$
.
Acknowledgements
We thank G. Pappas for several useful comments on a preliminary version of this article. We also thank the referees for their useful suggestions.
Competing interest
The authors have no competing interests to declare.
Financial support
The first author was supported by CRC 1442 Geometry: Deformations and Rigidity and the Excellence Cluster of the Universität Münster.