1 Introduction
1.1 Purpose of this paper and main theorem
The purpose of this paper is to study certain varieties
$X_n$
that live in the half-spin representations of the even spin groups
$\operatorname {\mathrm {Spin}}(2n)$
with n varying. In particular, we will show that these varieties are defined, for all n, by pulling back the equations for a single
$X_{n_0}$
along suitable contraction maps. The simplest instance of such a variety is the Grassmannian of n-dimensional isotropic spaces in a
$2n$
-dimensional orthogonal space. In this case, we use earlier work [Reference Seynnaeve and Tairi16] by the last two authors to show that
$n_0$
can be taken equal to
$4$
; see Theorem 6.1.
But the half-spin varieties that we introduce go far beyond the maximal isotropic Grassmannian. Indeed, this class of varieties is preserved under linear operations such as joins and tangential varieties, and under finite unions and arbitrary intersections. Consequently, any variety obtained from several copies of the maximal isotropic Grassmannian by such operations is defined by equations of some degree bounded independently of n. We stress, though, that these results are of a purely topological/set-theoretic nature. It is not true, for instance, that one gets the entire ideal of the maximal isotropic Grassmannian of n-spaces in a
$2n$
-space by pulling back equations for
$X_4$
along the maps that we define.
Our main results about half-spin varieties are Theorem 5.6, which establishes a descending chain condition for these, and Corollary 5.8, which implies the results mentioned above. These results follow from a companion result in infinite dimensions, which is a little easier to state here. We will construct a direct limit
$\operatorname {\mathrm {Spin}}(V_\infty )$
of all spin groups; here,
$V_\infty =\bigcup _n V_n$
is a countable-dimensional vector space with basis
$e_1,f_1,e_2,f_2,e_3,f_3,\ldots $
and a bilinear form determined by
$(e_i|e_j)=(f_i|f_j)=0$
and
$(e_i|f_j)=\delta _{ij}$
. Furthermore, we will construct a direct limit
$\bigwedge \nolimits _\infty ^+ E_\infty $
of all even half-spin representations. This space has as basis all formal infinite products

where
$\{i_1<i_2<\ldots \}$
is a cofinite subset of the positive integers. The group
$\operatorname {\mathrm {Spin}}(V_\infty )$
acts naturally on this space, and hence on its dual
$(\bigwedge \nolimits _\infty ^+ E_\infty )^*$
, which we regard as the spectrum of the symmetric algebra on
$\bigwedge \nolimits _\infty ^+ E_\infty $
. Our main theorem is as follows.
Theorem 1.1. The scheme
$(\bigwedge \nolimits _\infty ^+ E_\infty )^*$
is topologically
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherian. That is, every chain

of
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable reduced closed subschemes stabilises.
1.2 Relations to the literature
Our work is primarily motivated by earlier work by the second and third author on Plücker varieties, which live in exterior powers
$\bigwedge \nolimits ^n K^{p+n}$
with both p and n varying. The results in [Reference Draisma and Eggermont6] on Plücker varieties are analoguous to the results we establish here for half-spin varieties, and the main result in [Reference Nekrasov12] is an exact analogue of Theorem 1.1 for the dual infinite wedge, acted upon by the infinite general linear group.
On the one hand, we now have much better tools available to study these kind of questions than we had at the time of [Reference Draisma and Eggermont6] – notably the topological Noetherianity of polynomial functors [Reference Draisma5] and their generalisation to algebraic representations [Reference Eggermont and Snowden7]. But on the other hand, spin representations are much more intricate than polynomial functors, and a part of the current paper will be devoted to establishing the precise relationship between the infinite half-spin representation and algebraic representations of the infinite general linear group, so as to use those tools.
This paper fits in a general programme that asks for which sequences of representations of increasing groups one can expect Noetherianity results. This seems to be an extremely delicate question. Indeed, while Theorem 1.1 establishes Noetherianity of the dual infinite half-spin representation, we do not know whether the dual infinite spin representation is
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherian; see Remark 4.9. Similarly, we do not know whether a suitable inverse limit of exterior powers
$\bigwedge \nolimits ^n V_n$
is
$\operatorname {\mathrm {SO}}(V_\infty )$
-Noetherian – and there are many more natural sequences of representations for which we do not yet have satisfactory results.
In the context of secant varieties, we point out the work by Sam on Veronese varieties: the k-th secant variety of the d-th Veronese embedding of
$\mathbb {P}(K^n)$
is defined ideal-theoretically by finitely many types of equations, independently of n – and in particular in bounded degree [Reference Sam14]. Furthermore, a similar statement holds for the p-th syzygies for any fixed p [Reference Sam15]. Similar results for ordinary Grassmannians were established by Laudone in [Reference Laudone9]. It would be very interesting to know whether their techniques apply to secant varieties of the maximal isotropic Grassmannian in its spinor embedding. Our results here give a weaker set-theoretic statement, but for a more general class of varieties.
After establishing Noetherianity, it would be natural to try and study additional geometric properties of
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable subvarieties of the dual infinite half-spin representation. Perhaps there is a theory there analogous to the theory of
$\operatorname{\mathrm{GL}} $
-varietes [Reference Bik, Draisma, Eggermont and Snowden1, Reference Bik, Draisma, Eggermont and Snowden2]. However, we are currently quite far from any such deeper understanding!
1.3 Organisation of this paper
In §2, we recall the construction of the (finite-dimensional) half-spin representations. We mostly do this in a coordinate-free manner, only choosing – as one must – a maximal isotropic subspace of an orthogonal space for the construction. But for the construction of the infinite half-spin representation, we will need explicit formulas, and these are derived in §2, as well.
In §3, we first describe the embedding of the maximal isotropic Grassmannian in the projectivised half-spin representation. Then, we define suitable contraction and multiplication maps, which we show preserve the cones over these isotropic Grassmannians. Finally, we use these maps to construct the infinite-dimensional half-spin representations.
In §4, we prove Theorem 1.1 (see Theorem 4.1); and in §5, we state and prove the main results about half-spin varieties discussed above. Finally, in §6, we prove the universality of the isotropic Grassmannian of
$4$
-spaces in an
$8$
-dimensional space. We do so by relating the half-spin representations via the Cartan map to the exterior power representations and using results from [Reference Seynnaeve and Tairi16].
2 Finite spin representations and the spin group
In this section, we collect some preliminaries on spin groups and their defining representations. Throughout, we will assume that K is an algebraically closed field of characteristic
$0$
. We follow [Reference Manivel11] in our set-up; for more general references on spin groups and their representations, see [Reference Lawson and Michelsohn10, Reference Procesi13].
2.1 The Clifford algebra
Let V be a finite-dimensional vector space over K endowed with a quadratic form q. The Clifford algebra
$\operatorname{\mathrm{Cl}} (V,q)$
of V is the quotient of the tensor algebra
$T(V) = \bigoplus _{d \geq 0 } V^{\otimes d}$
by the two-sided ideal generated by all elements

This is also the two-sided ideal generated by

where
$(\cdot |\cdot )$
denotes the bilinear form associated to q defined by
$(v|w):=\frac {1}{2}(q(v+w)-q(v)-q(w))$
.
The Clifford algebra is a functor from the category of vector spaces equipped with a quadratic form to the category of (unital) associative algebras. That is, any linear map
$\varphi \colon (V,q) \to (V',q')$
with
$q'(\varphi (v))=q(v)$
for all
$v \in V$
induces a homomorphism of associative algebras
$\operatorname{\mathrm{Cl}} (\varphi ) \colon \operatorname{\mathrm{Cl}} (V,q) \to \operatorname{\mathrm{Cl}} (V',q')$
. If
$\varphi $
is an inclusion
$V \subseteq V'$
, then
$\operatorname{\mathrm{Cl}} (\varphi )$
is injective, and hence,
$\operatorname{\mathrm{Cl}} (V,q)$
is a subalgebra of
$\operatorname{\mathrm{Cl}} (V',q')$
.
The decomposition of
$T(V)$
into the even part
$T^+(V):=\bigoplus _{d \text { even}} V^{\otimes d}$
and the odd part
$T^-(V):=\bigoplus _{d \text { odd}} V^{\otimes d}$
induces a decomposition
$\operatorname{\mathrm{Cl}} (V,q) = \operatorname{\mathrm{Cl}} ^+(V,q) \oplus \operatorname{\mathrm{Cl}} ^-(V,q)$
, turning
$\operatorname{\mathrm{Cl}} (V,q)$
into a
${\mathbb {Z}}/2{\mathbb {Z}}$
-graded associative algebra. Note that, via the commutator on
$\operatorname{\mathrm{Cl}} (V,q)$
, the even Clifford algebra
$\operatorname{\mathrm{Cl}} ^+(V,q)$
is a Lie subalgebra of
$\operatorname{\mathrm{Cl}} (V,q)$
.
The anti-automorphism of
$T(V)$
determined by
$v_1 \otimes \cdots \otimes v_d \mapsto v_d \otimes \cdots \otimes v_1$
preserves the ideal in the definition of
$\operatorname{\mathrm{Cl}} (V,q)$
and therefore induces an anti-automorphism
$x \mapsto x^*$
of
$\operatorname{\mathrm{Cl}} (V,q)$
.
2.2 The Grassmann algebra as a
$\operatorname{\mathrm{Cl}} (V)$
-module
From now on, we will write
$\operatorname{\mathrm{Cl}} (V)$
for
$\operatorname{\mathrm{Cl}} (V,q)$
when q is clear from the context. If
$q=0$
, then
$\operatorname{\mathrm{Cl}} (V) =\bigwedge \nolimits V$
, the Grassmann algebra of V. If
$E \subseteq V$
is an isotropic subspace – that is, a subspace for which
$q|_E=0$
– then this fact allows us to identify
$\bigwedge \nolimits E$
with the subalgebra
$\operatorname{\mathrm{Cl}} (E)$
of
$\operatorname{\mathrm{Cl}} (V)$
.
For general q,
$\operatorname{\mathrm{Cl}} (V)$
is not isomorphic as an algebra to
$\bigwedge \nolimits V$
, but
$\bigwedge \nolimits V$
is naturally a
$\operatorname{\mathrm{Cl}} (V)$
-module as follows. For
$v \in V$
, define
$o(v):\bigwedge \nolimits V \to \bigwedge \nolimits V$
(the ‘outer product’) as the linear map

and
$\iota (v):\bigwedge \nolimits V \to \bigwedge \nolimits V$
(the ‘inner product’) as the linear map determined by

Here, and elsewhere in the paper,
$\kern1.2pt\widehat {\cdot }$
indicates a factor that is left out. Now
$v \mapsto \iota (v) + o(v)$
extends to an algebra homomorphism
$\operatorname{\mathrm{Cl}} (V) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits V)$
. To see this, it suffices to consider
$v,w_1,\ldots ,w_k \in V$
and verify

We write
$a \bullet \omega $
for the outcome of
$a \in \operatorname{\mathrm{Cl}} (V)$
acting on
$\omega \in \bigwedge \nolimits V$
. Using induction on the degree of a product, the linear map
$\operatorname{\mathrm{Cl}} (V) \to \bigwedge \nolimits V, a \mapsto a \bullet 1$
is easily seen to be an isomorphism of vector spaces. In particular,
$\operatorname{\mathrm{Cl}} (V)$
has dimension
$2^{\dim V}$
.
2.3 Embedding
$\mathfrak {so}(V)$
into the Clifford algebra
From now on, we assume that q is nondegenerate and write
$\operatorname {\mathrm {SO}}(V)=\operatorname {\mathrm {SO}}(V,q)$
for the special orthogonal group of q. Its Lie algebra
$\mathfrak {so}(V)$
consists of linear maps
$V \to V$
that are skew-symmetric with respect to
$(\cdot |\cdot )$
– that is, those
$A \in \operatorname{\mathrm{End}} (V)$
such that
$(Av|w) = -(v|Aw)$
for all
$v, w \in V$
. We have a unique linear map
$\psi : \bigwedge \nolimits ^2 V \to \operatorname{\mathrm{Cl}} ^+(V)$
with
$\psi (u \wedge v)=uv-vu$
, and
$\psi $
is injective. A straightforward computation shows that the image L of
$\psi $
is closed under the commutator in
$\operatorname{\mathrm{Cl}} (V)$
, and hence a Lie subalgebra. We claim that L is isomorphic to
$\mathfrak {so}(V)$
. Indeed, for
$u,v,w \in V$
, we have

We see, first, that
$V \subseteq \operatorname{\mathrm{Cl}} (V)$
is preserved under the adjoint action of L; and second, that L acts on V via skew-symmetric linear maps, so that L maps into
$\mathfrak {so}(V)$
. Since every map in
$\mathfrak {so}(V)$
is a linear combination of the linear maps above, and since
$\dim (L)=\dim (\mathfrak {so}(V))$
, the map
$L \to \mathfrak {so}(V)$
is an isomorphism. We will identify
$\mathfrak {so}(V)$
with the Lie subalgebra
$L \subseteq \operatorname{\mathrm{Cl}} (V)$
via the inverse of this isomorphism, and we will identify
$\bigwedge \nolimits ^2 V$
with
$\mathfrak {so}(V)$
via the map
$u\wedge v \mapsto ( w \mapsto (v|w) u - (u|w)v)$
. The concatenation of these identifications is the linear map
$\frac {1}{4} \psi $
.
2.4 The half-spin representations
From now on, we assume that
$\dim (V)=2n$
. We believe that all our results hold mutatis mutandis also in the odd-dimensional case, but we have not checked the details. A maximal isotropic subspace U of V is an isotropic subspace which is maximal with respect to inclusion. Since K is algebraically closed, q has maximal Witt index, so that every maximal isotropic subspace of V has dimension n.
The spin representation of
$\mathfrak {so}(V)$
is constructed as follows. Let F be a maximal isotropic subspace of V and let
$f_1, \ldots , f_n$
be a basis of F. Define
$f:=f_1 \cdots f_n \in \operatorname{\mathrm{Cl}} (F)$
; this element in
$\operatorname{\mathrm{Cl}} (F)=\bigwedge \nolimits F$
is well defined up to a scalar. Then the left ideal
$\operatorname{\mathrm{Cl}} (V) \cdot f$
is a left module for the associative algebra
$\operatorname{\mathrm{Cl}} (V)$
, and hence for its Lie subalgebra
$\mathfrak {so}(V)$
. This ideal is called the spin representation of
$\mathfrak {so}(V)$
. As
$\operatorname{\mathrm{Cl}} (V)$
is
${\mathbb {Z}}/2{\mathbb {Z}}$
-graded, the spin representation splits into a direct sum of two subrepresentations for
$\operatorname{\mathrm{Cl}} ^+(V)$
, and hence for
$\mathfrak {so}(V) \subseteq \operatorname{\mathrm{Cl}} ^+(V)$
– namely,
$\operatorname{\mathrm{Cl}} ^+(V) \cdot f$
and
$\operatorname{\mathrm{Cl}} ^-(V) \cdot f$
. These representations are called the half-spin representations of
$\mathfrak {so}(V)$
.
2.5 Explicit formulas
We will need more explicit formulas for the action of
$\mathfrak {so}(V)$
on the half-spin representations. To this end, let E be another isotropic n-dimensional subspace of V such that
$V=E \oplus F$
. Then the map

is a linear isomorphism, and we use it to identify
$\bigwedge \nolimits E$
with the spin representation. We write
$\rho :\mathfrak {so}(V) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits E)$
for the corresponding representation. It splits as a direct sum of the half-spin representations
$\rho _+:\mathfrak {so}(V) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits ^+ E)$
and
$\rho _-:\mathfrak {so}(V) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits ^- E)$
, where
$\bigwedge \nolimits ^+E=\bigoplus _{d \text { even}} \bigwedge \nolimits ^d E$
and
$\bigwedge \nolimits ^-E=\bigoplus _{d \text { odd}} \bigwedge \nolimits ^d E$
.
In this model of the spin representation, the action of
$v \in E \subseteq \operatorname{\mathrm{Cl}} (V)$
on the spin representation
$\bigwedge \nolimits E$
is just the outer product on
$\bigwedge E: o(v):\bigwedge \nolimits E \to \bigwedge \nolimits E,\ \omega \mapsto v \wedge \omega $
, while the action of
$v \in F \subseteq \operatorname{\mathrm{Cl}} (V)$
is twice the inner product on
$\bigwedge E$
:

The factor
$2$
and the alternating signs come from the following identity in
$\operatorname{\mathrm{Cl}} (V)$
:

For a general
$v \in V$
, we write
$v = v' + v"$
with
$v' \in E$
,
$v" \in F$
. Then the action of V on
$\bigwedge \nolimits E$
is given by

We now compute the linear maps by means of which
$\mathfrak {so}(V)$
acts on
$\bigwedge \nolimits E$
. To this end, recall that a pair
$e,f \in V$
is called hyperbolic if e, f are isotropic and
$(e|f) = 1$
. Given the basis
$f_1,\ldots ,f_n$
of F, there is a unique basis
$e_1,\ldots ,e_n$
of E so that
$(e_i|f_j)=\delta _{ij}$
; then
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
is called a hyperbolic basis of V. Now the element
$e_i \wedge e_j \in \mathfrak {so}(V)$
acts on
$\bigwedge \nolimits E \simeq \operatorname{\mathrm{Cl}} (V)f$
via the linear map

the element
$f_i \wedge f_j$
acts via the linear map

and the element
$e_i \wedge f_j$
acts via the linear map

In particular,
$\omega _0:=e_1 \wedge \cdots \wedge e_n \in \bigwedge \nolimits E$
is mapped to
$0$
by all elements
$e_i \wedge e_j$
and all elements
$e_i \wedge f_j$
with
$i \neq j$
, and it is mapped to
$\frac {1}{2} \omega _0$
by all
$e_i \wedge f_i$
.
2.6 Highest weights of the half-spin representations
Recall, for example, from [Reference Jacobson8, Chapter IV, pages 140–141], that in the basis
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
, matrices in
$\mathfrak {so}(V)$
have the form

Here, the
$(e_i,e_j)$
-entry of A is the coefficient of
$e_i \wedge f_j$
, the
$(e_i,f_j)$
-entry of B is the coefficient of
$e_i \wedge e_j$
, and the
$(f_i,e_j)$
-entry of C is the coefficient of
$f_i \wedge f_j$
.
The diagonal matrices
$e_i \wedge f_i$
span a Cartan subalgebra of
$\mathfrak {so}(V)$
with standard basis consisting of
$h_i:=e_i \wedge f_i - e_{i+1} \wedge f_{i+1}$
for
$i=1,\ldots ,n-1$
and
$h_n:=e_{n-1} \wedge f_{n-1} + e_n \wedge f_n$
(this last element is forgotten in the basis of the Cartan algebra on [Reference Jacobson8, page 141]).
Now
$(e_i \wedge e_j) \omega _0=(e_i \wedge f_j) \omega _0=0$
for all
$i \neq j$
. Furthermore, the elements
$h_1,\ldots ,h_{n-1}$
map
$\omega _0$
to
$0$
, while
$h_n$
maps
$\omega _0$
to
$\omega _0$
. Thus, the Borel subalgebra maps the line
$K \omega _0$
into itself, and
$\omega _0$
is a highest weight vector of the fundamental weight
$\lambda _0:=(0,\ldots ,0,1)$
relative to the standard basis. Summarising,
$\omega _0 \in \bigwedge \nolimits E$
generates a copy of the irreducible
$\mathfrak {so}(V)$
-module
$V_{\lambda _0}$
with highest weight
$\lambda _0$
. Clearly, the
$\mathfrak {so}(V)$
-module generated by
$\omega _0$
is contained in
$\bigwedge \nolimits ^+ E$
if n is even, and contained in
$\bigwedge \nolimits ^- E$
when n is odd. One can also show that both half-spin representations are irreducible; hence, one of them is a copy of
$V_{\lambda _0}$
. For the other half-spin representation, consider the element

This element is mapped to zero by
$e_i \wedge e_j$
for all
$i \neq j$
and by
$e_i \wedge f_j$
for all
$i < j$
. It is further mapped to
$0$
by
$h_1,\ldots ,h_{n-2},h_n$
, and to
$\omega _1$
by
$h_{n-1}$
. For example, we have

Hence,
$\omega _1$
generates a copy of
$V_{\lambda _1}$
, the irreducible
$\mathfrak {so}(V)$
-module of highest weight
$\lambda _1 := (0,\ldots ,0,1,0)$
; this is the other half-spin representation.
2.7 The spin group
Let
$\rho : \mathfrak {so}(V) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits E)$
be the spin representation. The spin group
$\operatorname {\mathrm {Spin}}(V)$
can be defined as the subgroup of
$\operatorname{\mathrm{GL}} (\bigwedge \nolimits E)$
generated by the one-parameter subgroups
$t \mapsto \exp (t \rho (X))$
, where X runs over the root vectors
$e_i \wedge e_j, f_i \wedge f_j$
and
$e_i \wedge f_j$
with
$i \neq j$
. Note that
$\rho (X)$
is nilpotent for each of these root vectors, so that
$t \mapsto \exp (t \rho (X))$
is an algebraic group homomorphism
$K \to \operatorname{\mathrm{GL}} (\bigwedge \nolimits E)$
. It is a standard fact that the subgroup generated by irreducible curves through the identity in an algebraic group is itself a connected algebraic group; see [Reference Borel3, Proposition 2.2]. So
$\operatorname {\mathrm {Spin}}(V)$
is a connected algebraic group, and one verifies that its Lie algebra is isomorphic to the Lie algebra generated by the root vectors X (i.e., to
$\mathfrak {so}(V)$
).
By construction, the (half-)spin representations
$\bigwedge \nolimits E$
,
$\bigwedge \nolimits ^+ E$
and
$\bigwedge \nolimits ^- E$
are representations of
$\operatorname {\mathrm {Spin}}(V)$
. We use the same notation
$ \rho : \operatorname {\mathrm {Spin}}(V) \to \operatorname{\mathrm{GL}} (\bigwedge \nolimits E), \ \rho _+ \colon \operatorname {\mathrm {Spin}}(V) \to \operatorname{\mathrm{GL}} (\bigwedge \nolimits ^+ E), \text { and } \rho _- \colon \operatorname {\mathrm {Spin}}(V) \to \operatorname{\mathrm{GL}} (\bigwedge \nolimits ^- E) $
for these as for the corresponding Lie algebra representations.
Remark 2.1. The algebraic group
$\operatorname {\mathrm {Spin}}(V)$
is usually constructed as a subgroup of the unit group
$\operatorname{\mathrm{Cl}} ^{\times }(V)$
as follows: consider first

sometimes called the Clifford group. Then
$\operatorname {\mathrm {Spin}}(V)$
is the subgroup of
$\Gamma (V)$
of elements of spinor norm
$1$
; that is,
$x x^* = 1$
, where
$x^*$
denotes the involution defined in Section 2.1. In this model of the spin group, it is easy to see that it admits a
$2:1$
covering
$\operatorname {\mathrm {Spin}}(V) \to \operatorname {\mathrm {SO}}(V)$
– namely, the restriction of the homomorphism
$\Gamma (V) \to \mathrm {O}(V)$
given by associating to
$x \in \Gamma (V)$
the orthogonal transformation
$w \mapsto x w x^{-1}$
. For more details, see [Reference Procesi13]. Since our later computations involve the Lie algebra
$\mathfrak {so}(V)$
only, the definition of
$\operatorname {\mathrm {Spin}}(V)$
above suffices for our purposes.
The half-spin representations are not representations of the group
$\operatorname {\mathrm {SO}}(V)$
; this can be checked, for example, by showing that the highest weights
$\lambda _0$
and
$\lambda _1$
are not in the weight lattice of
$\operatorname {\mathrm {SO}}(V)$
.
2.8 Two actions of
$\mathfrak{gl}(E) $
on
$\bigwedge\nolimits E $
The definition of the (half-)spin representation(s) of
$\mathfrak {so}(V)$
and
$\operatorname {\mathrm {Spin}}(V)$
as
$\operatorname{\mathrm{Cl}} ^{(\pm )}(V) f$
involves only the quadratic form q and the choice of a maximal isotropic space
$F \subseteq V$
. Consequently, any linear automorphism of V that preserves q and maps F into itself also acts on
$\operatorname{\mathrm{Cl}} ^{(\pm )}(V) f$
. These linear automorphisms form the stabiliser of F in
$\operatorname {\mathrm {SO}}(V)$
, which is the parabolic subgroup whose Lie algebra consists of the matrices in
$\operatorname {\mathrm {SO}}(V)$
that are block lower triangular in the basis
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
. So, while
$\operatorname {\mathrm {SO}}(V)$
does not act naturally on the (half-)spin representation(s), this stabiliser does.
In particular, in our model
$\bigwedge \nolimits ^{(\pm )} E$
of the (half-)spin representation(s), the group
$\operatorname{\mathrm{GL}} (E)$
, embedded into
$\operatorname {\mathrm {SO}}(V)$
as the subgroup of block diagonal matrices

acts on
$\bigwedge \nolimits E$
in the natural manner. We stress that this is not the action obtained by integrating the action of
$\mathfrak {gl}(E) \subseteq \mathfrak {so}(V)$
on
$\bigwedge \nolimits E$
regarded as the spin representation. Indeed, the standard action of
$e_i \wedge f_j \in \mathfrak {gl}(E)$
on
$\omega :=e_{i_1} \wedge \cdots \wedge e_{i_k} \in \bigwedge \nolimits ^k E$
yields

However, in the spin representation, the action is given by the linear map
$\frac {1}{2}(o(e_i) \iota (f_j) - \iota (f_j) o(e_i))$
. If
$j \neq i$
and
$j \not \in \{i_1,\ldots ,i_k\}$
, then

If
$j \neq i$
and
$j=i_l$
, then

We conclude that for
$i \neq j$
, the action of
$e_i \wedge f_j$
is the same in both representations. However, if
$i=j$
, then

We conclude that if
$\tilde {\rho }:\mathfrak {gl}(E) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits E)$
is the standard representation of
$\mathfrak {gl}(E)$
, then the restriction of the spin representation
$\rho :\mathfrak {so}(V) \to \operatorname{\mathrm{End}} (\bigwedge \nolimits E)$
to
$\mathfrak {gl}(E)$
as a subalgebra of
$\mathfrak {so}(V)$
satisfies

At the group level, this is to be understood as follows. The pre-image of
$\operatorname{\mathrm{GL}} (E) \subseteq \operatorname {\mathrm {SO}}(V)$
in
$\operatorname {\mathrm {Spin}}(V)$
is isomorphic to the connected algebraic group

for which
$(g,t) \mapsto g$
is a
$2:1$
cover of
$\operatorname{\mathrm{GL}} (E)$
, and the restriction of
$\rho $
to H satisfies
$\rho (g,t)=\tilde {\rho }(g) \cdot t^{-1}$
– a ‘twist of the standard representation by the inverse square root of the determinant’.
3 The isotropic Grassmannian and infinite spin representations
3.1 The isotropic Grassmannian in its spinor embedding
As before, let V be a
$2n$
-dimensional vector space over K endowed with a nondegenerate quadratic form. The (maximal) isotropic Grassmannian
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V,q)$
parametrizes all maximal isotropic subspaces of V. It has two connected components, denoted
$\operatorname{\mathrm{Gr}} ^+_{\operatorname {\mathrm {iso}}}(V)$
and
$\operatorname{\mathrm{Gr}} ^-_{\operatorname {\mathrm {iso}}}(V)$
. The goal of this subsection is to introduce the isotropic Grassmann cone, which is an affine cone over
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V,q)$
in the spin representation.
Fix a maximal isotropic subspace
$F \subseteq V$
and as before, set
, where
$f_1, \dots , f_n$
is any basis of F. Now let
$H \subseteq V$
be another maximal isotropic space. Then we claim that the space

is
$1$
-dimensional. Indeed, we may find a hyperbolic basis
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
of V such that
$f_1,\ldots ,f_k$
span
$H \cap F$
,
$f_1,\ldots ,f_n$
span F, and
$e_{k+1},\ldots ,e_n,f_{1},\ldots ,f_k$
span H. We call this hyperbolic basis adapted to H and F. Then the element

lies in
$S_H$
since
$e_i \omega _H=f_j \omega _H = 0$
for all
$i> k$
and
$j \leq k$
. Conversely, if
$\mu \in S_H$
, then write

If
$c_I \neq 0$
for some I with
$I \not \supseteq \{k+1,\ldots ,n\}$
, then for any
$j \in \{k+1,\ldots ,n\} \setminus I$
, we find that
$e_j \mu \neq 0$
. So all I with
$c_I \neq 0$
contain
$\{k+1,\ldots ,n\}$
. If some I with
$c_I \neq 0$
further contains an
$i \leq k$
, then
$f_i \mu $
is nonzero. Hence,
$S_H$
is spanned by
$\omega _H$
, as claimed. In what follows, by slight abuse of notation, we will write
$\omega _H$
for any nonzero vector in
$S_H$
.
The space H can be uniquely recovered from
$\omega _H$
via

Indeed, we have already seen
$\subseteq $
. For the converse, observe that the vectors
$e_i \omega _H, f_j \omega _H$
with
$i \leq k$
and
$j> k$
are linearly independent.
The map that sends
$H \in \operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V,q)$
to the projective point representing it, that is,

is therefore injective, and it is called the spinor embedding of the isotropic Grassmannian (see [Reference Manivel11]). The isotropic Grassmann cone is defined as

where the union is taken over all maximal isotropic subspaces
$H\subseteq V$
. We denote by
the cones over the connected components of the isotropic Grassmannian in its spinor embedding.
3.2 Contraction with an isotropic vector
Let
$e \in V$
be a nonzero isotropic vector. Then
$V_e:=e^\perp / \langle e \rangle $
is equipped with a natural nondegenerate quadratic form, and there is a rational map
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V) \to \operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V_e)$
that maps an n-dimensional isotropic space H to the image in
$V_e$
of the
$(n-1)$
-dimensional isotropic space
$H \cap e^\perp $
(this is defined if
$e \not \in H$
, which by maximality of H is equivalent to
$H \not \subseteq e^\perp $
). This map is the restriction to
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V)$
of the rational map
$\mathbb {P}(\bigwedge \nolimits ^n V) \to \mathbb {P}(\bigwedge \nolimits ^{n-1} V_e)$
induced by the linear map (‘contraction with e’):

where
$\overline {v_i}$
is the image of
$v_i$
in
$V/\langle e \rangle $
. Note first that this map is the inner product
$\iota (e)$
followed by a projection. Furthermore, a priori, the codomain of this map is the larger space
$\bigwedge \nolimits ^{n-1} (V/\langle e \rangle )$
, but one may choose
$v_1,\ldots ,v_n$
such that
$(e|v_i)=0$
for
$i>1$
, and then it is evident that the image is indeed in
$\bigwedge \nolimits ^{n-1} V_e$
.
We want to construct a similar contraction map at the level of the spin representation. For reasons that will become clear in a moment, we restrict our attention first to a map between two half-spin representations, as follows. Assume that
$e \notin F$
, and choose a basis
$f_1,\ldots ,f_n$
of F such that
$(e|f_i)=\delta _{in}$
. As usual, write
$f:=f_1 \cdots f_n$
, and write
$\overline {f}:=\overline {f}_1 \cdots \overline {f}_{n-1}$
, so that
$\operatorname{\mathrm{Cl}} ^+(V_e)\overline {f}$
is a half-spin representation of
$\mathfrak {so}(V_e)$
.
Then we define the map

where the implicit claim is that the expression on the right lies in
$\operatorname{\mathrm{Cl}} (e^\perp )f_1\cdots f_{n-1}$
, so that its image in
$\operatorname{\mathrm{Cl}} (V_e) \overline {f}$
is well defined (note that the projection
$e^\perp \to V_e$
induces a homomorphism of Clifford algebras), and that this image lies in the left ideal generated by
$\overline {f}$
. To verify this claim, and to derive a more explicit formula for the map above, let
$e_1,\ldots ,e_{n}=e$
be a basis of an isotropic space E complementary to F. Then it suffices to consider the case where
$a=e_{i_1} \cdots e_{i_k}$
for some
$i_1<\ldots <i_k$
. We then have

Multiplying by
$(-1)^{n-1}$
and using that k is even, the latter expression becomes

Hence, we conclude that

In short, in our models
$\bigwedge \nolimits ^+ E$
and
$\bigwedge \nolimits ^+(E/\langle e \rangle )$
for the half-spin representations of
$\mathfrak {so}(V)$
and
$\mathfrak {so}(V_e)$
,
$\pi _e$
is just the reduction-mod-e map. We leave it to the reader to check that the reduction-mod-e map
$\bigwedge \nolimits ^- E \to \bigwedge \nolimits ^-(E/\langle e \rangle )$
arises in a similar fashion from the map

We will informally call the maps
$\pi _e$
‘contraction with e’. Together, they define a map on
$\operatorname{\mathrm{Cl}} (V)f$
which we also denote by
$\pi _e$
.
Proposition 3.1. The contraction map
$\pi _e:\operatorname{\mathrm{Cl}} (V)f \to \operatorname{\mathrm{Cl}} (V)\overline {f}$
is a homomorphism of
$\operatorname{\mathrm{Cl}} (e^\perp )$
-representations.
Proof. Let
$v \in e^\perp $
and consider
$a \in \operatorname{\mathrm{Cl}} ^-(V)$
. Then
$va \in \operatorname{\mathrm{Cl}} ^+(V)$
, and hence,
$\pi _e(vaf)$
is the image in
$\operatorname{\mathrm{Cl}} (V_e)\overline {f}$
of

where we have used
$(v|e)=0$
in the first equality. The right-hand side clearly equals
$\overline {v}$
times the image of
$\pi _e(af)$
in
$\operatorname{\mathrm{Cl}} (V_e) \overline {f}$
.
3.3 Multiplying with an isotropic vector
In a sense dual to the contraction maps,
$c_e:\bigwedge \nolimits ^n V \to \bigwedge \nolimits ^{n-1} V_e$
are multiplication maps defined as follows. Let
$e,h \in V$
be isotropic with
$(e|h)=1$
; such a pair is called a hyperbolic pair. We then have
$V=\langle e,h \rangle \oplus \langle e,h \rangle ^\perp $
, and the map from the second summand to
$V_e=e^\perp /\langle e \rangle $
is an isometry. We use this isometry to identify
$V_e$
with the subspace
$\langle e,h \rangle ^\perp $
of V and write
$s_e$
for the corresponding inclusion map. Then we define

which is just the outer product
$o(h)$
. The projectivisation of this map sends
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V_e)$
isomorphically to the closed subset of
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V)$
consisting of all H containing h. We further observe that

We define a corresponding multiplication map at the level of spin representations as follows: first, we assume that
$h \in F$
, and choose a basis
$f_1,\ldots ,f_n=h$
of F such that
$(e|f_i)=\delta _{in}$
. As usual, we set
$f=f_1 \cdots f_n$
and
$\overline {f}=\overline {f}_1 \cdots \overline {f}_{n-1}$
. Then we define

Note that, for
$a \in \operatorname{\mathrm{Cl}} (V_e)$
, we have

where the last identity can be seen verified in the model
$\bigwedge \nolimits E$
for the spin representation, where
$\pi _e$
is the reduction-mod-e map, and
$\tau _h$
is just the inclusion
$\bigwedge \nolimits E/\langle e \rangle \to \bigwedge \nolimits E$
corresponding to the inclusion
$V_e \to V$
. So
$\pi _e \circ \tau _h=\operatorname{\mathrm{id}} _{\operatorname{\mathrm{Cl}} (V_e)\overline {f}}$
. We will informally call
$\tau _h$
the multiplication map with h.
Proposition 3.2. The multiplication map
$\tau _h:\operatorname{\mathrm{Cl}} (V_e)\overline {f} \to \operatorname{\mathrm{Cl}} (V)f$
is a homomorphism of
$\operatorname{\mathrm{Cl}} (V_e)$
-representations, where
$\operatorname{\mathrm{Cl}} (V_e)$
is regarded a subalgebra of
$\operatorname{\mathrm{Cl}} (V)$
via the section
$s_e:V_e \to V$
.
Proof. Let
$v \in V_e$
and let
$a \in \operatorname{\mathrm{Cl}} (V_e)$
. Then

as desired.
Corollary 3.3. Both the map
$\pi _e: \operatorname{\mathrm{Cl}} (V)f \to \operatorname{\mathrm{Cl}} (V_e)\overline {f}$
and the map
$\tau _h :\operatorname{\mathrm{Cl}} (V_e)\overline {f} \to \operatorname{\mathrm{Cl}} (V)f$
are
$\operatorname {\mathrm {Spin}}(V_e)$
-equivariant, where
$\operatorname {\mathrm {Spin}}(V_e)$
is regarded as a subgroup of
$\operatorname {\mathrm {Spin}}(V)$
via the orthogonal decomposition
$V=V_e \oplus \langle e,h \rangle $
.
Proof. Propositions 3.1 and 3.2 imply that both maps are homomorphisms of
$\mathfrak {so}(V_e)$
-representations. Since
$\operatorname {\mathrm {Spin}}(V_e)$
is generated by one-parameter subgroups corresponding to nilpotent elements of
$\mathfrak {so}(V_e)$
,
$\pi _e$
and
$\tau _h$
are
$\operatorname {\mathrm {Spin}}(V_e)$
-equivariant.
3.4 Properties of the isotropic Grassmannian
The goal of this subsection is to collect properties of the isotropic Grassmann cone that will later motivate the definition of a (half-)spin variety (see Section 5). We fix a maximal isotropic subspace
$F \subseteq V$
and a hyperbolic pair
$(e,h)$
with
$h \in F$
and
$e \not \in F$
and identify
$V_e=e^\perp /\langle e \rangle $
with the subspace
$\langle e,h \rangle ^\perp $
of V. We choose any basis
$f_1,\ldots ,f_n$
of F with
$f_n=h$
and
$(e|f_i)=0$
for
$i<n$
and write
$f:=f_1 \cdots f_n \in \operatorname{\mathrm{Cl}} (V)$
and
$\overline {f}:=\overline {f_1} \cdots \overline {f_{n-1}} \in \operatorname{\mathrm{Cl}} (V_e)$
.
Proposition 3.4. The isotropic Grassmann cone in
$\operatorname{\mathrm{Cl}} (V)f$
has the following properties:
-
1.
$\widehat {\operatorname{\mathrm{Gr}} }_{\operatorname {\mathrm {iso}}}(V) \subseteq \operatorname{\mathrm{Cl}} (V)f$ is Zariski-closed and
$\operatorname {\mathrm {Spin}}(V)$ -stable.
-
2. Let
$\pi _e: \operatorname{\mathrm{Cl}} (V) f \to \operatorname{\mathrm{Cl}} (V_e) \overline {f}$ be the contraction defined in §3.2. Then for every maximal isotropic subspace
$H \subseteq V$ , we have
$$\begin{align*}\pi_e(S_H) \subseteq S_{H_e}, \end{align*}$$
$H_e \subseteq V_e$ is the image of
$e^\perp \cap H$ in
$V_e$ .
-
3. Let
$\tau _h: \operatorname{\mathrm{Cl}} (V_e) \overline {f} \to \operatorname{\mathrm{Cl}} (V) f$ be the map defined in §3.3. Then for every maximal isotropic
$H' \subseteq V_e$ , we have
$$\begin{align*}\tau_h(S_{H'}) = S_{{H'} \oplus \langle h \rangle}. \end{align*}$$
In particular, the contraction and multiplication map
$\pi _e$
and
$\tau _h$
preserve the isotropic Grassmann cones – that is,

Proof of Proposition 3.4.
-
1. This is well known. Indeed, the isotropic Grassmann cone is the union of the cones over the two connected components, and these cones are the union of
$\{0\}$ with the orbits of the highest weight vectors
$\omega _0$ and
$\omega _1$ . These minimal orbits are always Zariski closed. For more detail, see [Reference Procesi13, Theorem 1, p.428].
-
2. Let
$\omega _H$ be a spanning element of
$S_H$ . Then for all
$v \in e^\perp \cap H$ , we have
$$\begin{align*}\overline{v} \cdot \pi_e(\omega_H) = \pi_e(v \cdot \omega_H) = \pi_e(0)=0, \end{align*}$$
$\pi _e(\omega _H)$ lies in
$S_{H_e}$ .
-
3. Let
$\omega _{H'}$ be a spanning element of
$S_{H'}$ . Then for all
$v \in H'$ , we have
$$\begin{align*}v \cdot \tau_h(\omega_{H'}) = \tau_h (v \cdot \omega_{H'}) = \tau_h (0)=0, \end{align*}$$
$$\begin{align*}h \cdot \tau_h(\omega_{H'}) = h \cdot \omega_{H'} f_n = 0, \end{align*}$$
$\tau _h$ and
$h=f_n$ . Thus,
$\tau _h(\omega _{H'})$ lies in
$S_{{H'} \oplus \langle h \rangle }$ . The equality now follows from the fact that
$\tau _h$ is injective.
Remark 3.5. If
$h \in H$
, then
$H=H_e \oplus \langle h \rangle $
, and since
$\pi _e \circ \tau _h$
is the identity on
$\operatorname{\mathrm{Cl}} (V_e) \overline {f}$
, we find that

(i.e., equality holds in (2) of Proposition 3.4). Later, we will see that equality holds under the weaker condition that
$e \not \in H$
, while
$\pi _e(S_H)=\{0\}$
when
$e \in H$
. These statements can also be checked by direct computations, but some care is needed since for
$e,H,F$
in general position, one cannot construct a hyperbolic basis adapted to H and F that moreover contains e.
3.5 The dual of contraction
Let
$e \not \in F \subseteq V$
be an isotropic vector. We want to compute the dual of the contraction map
${\pi _e: \operatorname{\mathrm{Cl}} (V)f \to \operatorname{\mathrm{Cl}} (V_e)\overline {f}}$
; indeed, we claim that this is essentially the map

defined by its restriction
$\operatorname{\mathrm{Cl}} ^\pm (V_e) \overline {f} \to \operatorname{\mathrm{Cl}} ^\mp (V) f$
as

where the sign is
$+$
on
$\operatorname{\mathrm{Cl}} ^+(V_e) \overline {f}$
and
$-$
on
$\operatorname{\mathrm{Cl}} ^-(V_e) \overline {f}$
. The reason for the ‘flip’ of the choice of half-spin representation in the dual will become obvious below. Observe that
$\psi _e$
is well defined and, given a basis
$e_1,\ldots ,e_n=e$
of an isotropic space complementary to F such that
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
is a hyperbolic basis, maps
$\overline {e_J} \overline {f}$
to
$e_{J \cup \{n\}} f$
.
Proposition 3.6. The following diagram

can be made commuting via a
$\operatorname {\mathrm {Spin}}(V_e)$
-module isomorphism on the left vertical arrow and a
$\operatorname {\mathrm {Spin}}(V)$
-module isomorphism on the right vertical arrow.
Remark 3.7. The statement of Proposition 3.6 holds true when replacing
$\operatorname{\mathrm{Cl}} (V) f$
by either one of the two half-spin representations by considering the correct ‘flip’. For example, if
$n = \dim F$
is even, and
$e_1,\dots , e_n, f_1, \dots , f_n$
is a hyperbolic basis as above, then in the
$\bigwedge \nolimits E$
-model, the correct grading is

To prove Proposition 3.6, we consider the bilinear form
$\beta $
on the spin representation
$\operatorname{\mathrm{Cl}} (V)f$
defined as in [Reference Procesi13] as follows: for
$af,bf \in \operatorname{\mathrm{Cl}} (V)f$
, it turns out that
$(af)^*bf=f^*a^*bf$
, where
$*$
denotes the anti-automorphism from §2.1, is a scalar multiple of f. The scalar is denoted
$\beta (af,bf)$
. We have the following properties:
Lemma 3.8 [Reference Procesi13, p. 430].
Let
$\beta $
be the bilinear form defined as above.
-
1. The form
$\beta $ is nondegenerate and
$\operatorname {\mathrm {Spin}}(V)$ -invariant.
-
2.
$\beta $ is symmetric if
$n \equiv 0,1 \ \mod 4$ , and it is skew-symmetric if
$n \equiv 2,3 \ \mod 4$ .
-
3. The two half-spin representations are self-dual via
$\beta $ if n is even, and each is the dual of the other if n is odd.
In the proof of Proposition 3.6, we will use a hyperbolic basis
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
with
$e_n=e$
. For a subset
$I=\{i_1<\ldots <i_k\} \subseteq [n]$
, set
$e_I:=e_{i_1} \cdots e_{i_k} \in \operatorname{\mathrm{Cl}} (E) \simeq \bigwedge \nolimits E$
, where E is the span of the
$e_i$
. We have seen in §2.5 that the spin representation has as a basis the elements
$e_I f$
with I running through all subsets of
$[n]$
.
Proof of Proposition 3.6.
Consider the bilinear forms
$\beta $
on
$\operatorname{\mathrm{Cl}} (V)f$
and
$\beta _e$
on
$\operatorname{\mathrm{Cl}} (V_e)\overline {f}$
as defined above. By Lemma 3.8, the spin representations
$\operatorname{\mathrm{Cl}} (V)f$
and
$\operatorname{\mathrm{Cl}} (V_e)\overline {f}$
are self-dual via
$\beta $
and
$\beta _e$
, respectively. Thus, it suffices to prove, for
$a \in \operatorname{\mathrm{Cl}} (V)$
and
$b \in \operatorname{\mathrm{Cl}} (e^\perp )$
, that

We may assume that
$a = e_I$
,
$b = e_J$
with
$I \subseteq [n]$
,
$J \subseteq [n-1]$
.
In the
$\bigwedge \nolimits E$
-model,
$\pi _e$
is the mod-e map, and hence, the left-hand side is zero if
$n \in I$
. If
$n \not \in I$
, then the left-hand side equals the coefficient of
$\overline {f}$
in
$\overline {f}^* \overline {e_I}^* \overline {e_J} \overline {f}$
. This is nonzero if and only if
$[n-1]$
is the disjoint union of I and J, and then it is
$2^{n-1}$
times a sign corresponding to the number of swaps needed to move the factors
$\overline {f_i}$
of
$\overline {f}^*$
to just before the corresponding factor
$\overline {e_i}$
in either
$\overline {e_I}^*$
or
$\overline {e_J}$
.
Apart from the factor
$\frac {(-1)^{n-1}}{2}$
, the right-hand side is the coefficient of f in
$f^* e_I e_J e_n f$
. This is nonzero if and only if
$[n]$
is the disjoint union of the sets
$\{n\},J,I$
, and in that case, it is
$2^n$
times a sign corresponding to the number of swaps needed to move the factors
$f_i$
of
$f^*$
to the corresponding factor
$e_i$
in either
$e_I$
or
$e_J$
or (in the case of
$f_n$
) to just before the factor
$e_n$
. The latter contributes
$(-1)^{n-1}$
, and apart from this factor, the sign is the same as on the left-hand side.
3.6 Two infinite spin representations
Let
$V_\infty $
be the countable-dimensional vector space with basis
$e_1, f_1, e_2, f_2, \dots $
, and equip
$V_\infty $
with the quadratic form for which this is a hyperbolic basis (i.e.,
$(e_i|e_j)=(f_i|f_j)=0$
and
$(e_i|f_j)=\delta _{ij}$
for all
$i,j$
). We write
$E_\infty $
and
$F_\infty $
for the subspaces of
$V_\infty $
spanned by the
$e_i$
and the
$f_i$
, respectively.
Let
$V_n$
be the subspace of
$V_\infty $
spanned by
$e_1, f_1, e_2, f_2, \dots , e_n, f_n$
, with the restricted quadratic form. We further set
$E_n:=V_n \cap E_\infty $
and
$F_n:=V_n \cap F_\infty $
. We define the infinite spin group as
$\operatorname {\mathrm {Spin}}(V_\infty ):=\varinjlim _n \operatorname {\mathrm {Spin}}(V_n)$
, where
$\operatorname {\mathrm {Spin}}(V_{n-1})$
is embedded into
$\operatorname {\mathrm {Spin}}(V_n)$
as the subgroup that fixes
$\langle e_n,f_n \rangle $
element-wise. Similarly, we write
$\operatorname{\mathrm{GL}} (E_\infty ):=\varinjlim _n \operatorname{\mathrm{GL}} (E_n)$
and H for the preimage of
$\operatorname{\mathrm{GL}} (E_\infty )$
in
$\operatorname {\mathrm {Spin}}(V_\infty )$
. We use the notation
$\mathfrak {so}(V_\infty )$
and
$\mathfrak {gl}(E_\infty )$
for the corresponding direct limits of the Lie algebras
$\mathfrak {so}(V_n)$
and
$\mathfrak {gl}(E_n)$
. Here, the direct limits are taken in the categories of abstract groups and Lie algebras, respectively.
The previous paragraphs give rise to various
$\operatorname {\mathrm {Spin}}(V_{n-1})$
-equivariant maps between the spin representations of
$\operatorname {\mathrm {Spin}}(V_{n-1})$
and
$\operatorname {\mathrm {Spin}}(V_n)$
. First, contraction with
$e_n$
,

and second, multiplication with
$f_n$
,

We have that these satisfy
$\pi _{e_n} \circ \tau _{f_n} =\operatorname{\mathrm{id}} $
. Third, the map

is dual to
$\pi _{e_n}$
in the sense of Proposition 3.6.
Definition 3.9. The direct (infinite) spin representation is the direct limit of all spaces
$\operatorname{\mathrm{Cl}} (V_n) f_1 \cdots f_n$
along the maps
$\psi _{e_n}$
. The inverse (infinite) spin representation is the inverse limit of all spaces
$\operatorname{\mathrm{Cl}} (V_n) f_1 \cdots f_n$
along the maps
$\pi _{e_n}$
.
Since the maps
$\psi _{e_n},\pi _{e_n}$
are
$\operatorname {\mathrm {Spin}}(V_{n-1})$
-equivariant, both of these spaces are
$\operatorname {\mathrm {Spin}}(V_\infty )$
-modules. As the dual of a direct limit is the inverse limit of the duals, and since the maps
$\psi _{e_n}$
and
$\pi _{e_n}$
are dual to each other by Proposition 3.6, the inverse spin representation is the dual space of the direct spin representation.
In our model
$\bigwedge \nolimits E_n$
of
$\operatorname{\mathrm{Cl}} (V_n) f_1 \cdots f_n$
, the map
$\psi _{e_n}$
is just the right multiplication

Hence, the direct spin representation has as a basis formal infinite products

where
$I=\{i_1<i_2<\ldots \}$
is a cofinite subset of
${\mathbb {N}}$
. We will write
$\bigwedge \nolimits _\infty E_\infty $
for this countable-dimensional vector space. The action of the Lie algebra
$\mathfrak {so}(V_\infty )$
of
$\operatorname {\mathrm {Spin}}(V_\infty )$
on this space is given via the explicit formulas from §2.5. In particular, the span of the
$e_I$
with
$|{\mathbb {N}} \setminus I|$
even (respectively, odd) is a
$\operatorname {\mathrm {Spin}}(V_\infty )$
-submodule, and
$\bigwedge \nolimits _\infty E_\infty $
is the direct sum of these (irreducible) modules.
Remark 3.10. The reader may wonder why we do not introduce the direct spin representation as the direct limit of all
$\operatorname{\mathrm{Cl}} (V)f_1\cdots f_n$
along the maps
$\tau _{f_n}$
. This would make the ordinary Grassmann algebra
$\bigwedge \nolimits E_\infty $
a model for the direct spin representation, instead of the slightly more complicated-looking space
$\bigwedge \nolimits _\infty E_\infty $
. However, the maps dual to the
$\tau _{f_n}$
correspond to contraction maps with
$f_n \in F$
, which we have not discussed and which interchange even and odd half-spin representations. We believe that our theorem below goes through for this different setting, as well, but we have not checked the details.
3.7 Four infinite half-spin representations
Keeping in mind that the maps
$\psi _{e_n}$
interchange the even and odd subrepresentations, we define the direct (infinite) half-spin representations
$\bigwedge \nolimits _\infty ^\pm E_\infty $
to be the direct limit

along the maps
$\psi _{e_n}$
. For the sake of readability, we will abbreviate this by

where
$\pm (-1)^n$
denotes
$\pm $
if n is even and
$\mp $
if n is odd. In terms of the basis
$e_I$
introduced in §3.6, the half-spin representation
$\bigwedge \nolimits _\infty ^+ E_\infty $
is spanned by all
$e_I$
with
$|{\mathbb {N}} \setminus I|$
even, and
$\bigwedge \nolimits _\infty ^- E_\infty $
by those with
$|{\mathbb {N}} \setminus I|$
odd. The inverse (infinite) half-spin representations are defined as the duals of the direct (infinite) half-spin representations. Using the isomorphisms from Remark 3.7, we observe

So the inverse (infinite) half-spin representations can be identified with the inverse limits of the half-spin representations
$\bigwedge \nolimits ^\pm E_n$
along the projections
$\pi _{e_n}$
.
We can enrich the inverse spin representation to an affine scheme whose coordinate ring is the symmetric algebra on
$\bigwedge \nolimits _\infty E_\infty $
, recalling the following remark.
Remark 3.11. Let K be any field (not necessarily algebraically closed) and W any K-vector space (not necessarily finite dimensional). Then there are canonical identifications

So
$\textrm { Spec}\big ( \textrm { Sym}(W)\big )$
can be seen as an enrichment of
$W^\ast $
to an affine scheme. If W is a linear representation for a group G, then G acts via K-algebra automorphisms on
$\operatorname{\mathrm{Sym}} W$
and hence via K-automorphisms on the affine scheme corresponding to
$W^*$
. For
$W = \bigwedge \nolimits ^\pm _\infty E_\infty $
, this construction extends the natural
$\operatorname {\mathrm {Spin}}(V_\infty )$
-action on the vector space
$\varprojlim _n \bigwedge \nolimits ^\pm E_n \simeq W^*$
to the corresponding affine scheme.
By abuse of notation, we will write
$\left (\bigwedge \nolimits _\infty E_\infty \right )^*$
also for the scheme itself, and similarly for the inverse half-spin representations
$\left (\bigwedge \nolimits ^\pm _\infty E_\infty \right )^*$
. Later, we will also write
$\bigwedge \nolimits ^\pm E_n$
for the affine scheme
$\textrm { Spec}\left (\textrm { Sym}\left ( \bigwedge \nolimits ^{\pm (-1)^n}E_n \right ) \right )$
by identifying
$\bigwedge \nolimits ^\pm E_n \cong \left (\bigwedge \nolimits ^{\pm (-1)^n}E_n\right )^*$
as in Equation(3.3).
4 Noetherianity of the inverse half-spin representations
In this section, we prove our main theorem.
Theorem 4.1. The inverse half-spin representation
$(\bigwedge \nolimits _\infty ^+ E_\infty )^*$
is topologically Noetherian with respect to the action of
$\operatorname {\mathrm {Spin}}(V_\infty )$
. That is, every descending chain

of closed, reduced
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable subschemes stabilises, and the same holds for the other inverse half-spin representation.
Recall that the action of
$\operatorname {\mathrm {Spin}}(V_\infty )$
on the inverse half-spin representation (as an affine scheme) is given by K-automorphisms, as described in Remark 3.11. We write R for the symmetric algebra on the direct spin representation
$\bigwedge \nolimits _\infty E_\infty $
, so the inverse spin representation is
$\operatorname{\mathrm{Spec}} (R)$
. Similarly, we write
$R^\pm $
for the symmetric algebras on the direct half-spin representations, so
$R^\pm $
is the coordinate ring of
$\varprojlim _n \bigwedge \nolimits ^\pm E_n$
, respectively.
Let us briefly outline the proof strategy. We will proceed by induction on the minimal degree of an equation defining a closed subset X. Starting with such an equation p, we show that there exists a partial derivative
$q := \frac {\partial p}{\partial e_I}$
such that the principal open
$X[1/q]$
is topologically
$H_n$
-Noetherian, where
$H_n$
is the subgroup of
$\operatorname {\mathrm {Spin}}(V_\infty )$
defined below. For that, we use that the
$H_n$
-action corresponds to a ‘twist’ of the usual
$\operatorname{\mathrm{GL}} (E_\infty )$
-action, as observed in Section 2.8 (for the exact formula see (2.3)); this allows us to apply the main result of [Reference Eggermont and Snowden7]. Finally, for those points which are contained in the vanishing set of the
$\operatorname {\mathrm {Spin}}(V_\infty )$
-orbit of q, we can apply induction, as the minimal degree of a defining equation has been lowered by
$1$
.
4.1 Shifting
Let
$G_n$
be the subgroup of G that fixes
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
element-wise. Note that
$G_n$
is isomorphic to G; at the level of the Lie algebras, the isomorphism from G to
$G_n$
is given by the map

where the widths of the blocks are
$n,\infty ,n,\infty $
, respectively. We write
$H_n$
for
$H \cap G_n$
, where
$H \subseteq \operatorname {\mathrm {Spin}}(V_\infty )$
is the subgroup corresponding to the subalgebra
$\mathfrak {gl}(E_\infty ) \subseteq \mathfrak {so}(V_\infty )$
. Then
$H_n$
is the pre-image in
$\operatorname {\mathrm {Spin}}(V_\infty )$
of the subgroup
$\operatorname{\mathrm{GL}} (E_\infty )_n \subseteq \operatorname{\mathrm{GL}} (E_\infty )$
of all g that fix
$e_1,\ldots ,e_n$
element-wise and maps the span of the
$e_i$
with
$i>n$
into itself. The Lie algebra of
$H_n$
and of
$\operatorname{\mathrm{GL}} (E_\infty )_n$
consists of the matrices above on the right with
$B=C=0$
.
4.2 Acting with the general linear group on E
For every fixed
$k \in {\mathbb {Z}}_{\geq 0}$
, the Lie algebra
$\mathfrak {gl}(E_\infty ) \subseteq \mathfrak {so}(V_\infty )$
preserves the linear space

and hence, so does the corresponding subgroup
$H \subseteq \operatorname {\mathrm {Spin}}(V_\infty )$
. We let
$R_{\leq \ell } \subseteq R$
be the subalgebra generated by the spaces
$(\bigwedge \nolimits _\infty E_\infty )_k$
with
$k \leq \ell $
. Crucial in the proof of Theorem 4.1 is the following result.
Proposition 4.2. For every choice of nonnegative integers
$\ell $
and n,
$\operatorname{\mathrm{Spec}} (R_{\leq \ell })$
is topologically
$H_n$
-Noetherian; that is, every descending chain

of
$H_n$
-stable closed and reduced subschemes stabilizes.
The key ingredient in the proof of Proposition 4.2 is the main result of [Reference Eggermont and Snowden7]. In order to apply their result, we need to do some preparatory work. We will start with the following lemma.
Lemma 4.3. Every
$H_n$
-stable closed subscheme of
$\operatorname{\mathrm{Spec}} (R_{\leq \ell })$
is also stable under the group
$\operatorname{\mathrm{GL}} (E_\infty )_n$
acting in the natural manner on
$\bigwedge \nolimits _\infty E_\infty $
and its dual, and vice versa.
Proof. Equation (2.3) implies that
$\mathfrak {gl}(E_\infty ) \subseteq \mathfrak {so}(V_\infty )$
acts on
$\bigwedge \nolimits _\infty E_\infty $
via

where
$\tilde {\rho }$
is the standard representation of
$\mathfrak {gl}(E_\infty )$
on
$\bigwedge \nolimits _\infty E_\infty $
. An
$H_n$
-stable closed subscheme X of
$\operatorname{\mathrm{Spec}} (R_{\leq \ell })$
is given by an
$H_n$
-stable ideal I in the symmetric algebra
$R_{\leq \ell }$
. Such an I is then also stable under the action of the Lie algebra
$\mathfrak {gl}(E_\infty )_n$
of
$H_n$
by derivations that act on variables in
$\bigoplus _{k=0}^\ell (\bigwedge \nolimits _\infty E_\infty )_{k}$
via
$\rho $
.
We claim that I is a homogeneous ideal. Indeed, for
$f \in I$
, choose
$m>n$
such that all variables in f (which are basis elements
$e_I$
) contain the basis element
$e_m$
of
$E_\infty $
. Let
$A \in \mathfrak {gl}(E_\infty )_n$
be the diagonal matrix with
$0$
’s everywhere except a
$1$
on position
$(m,m)$
. Then
$\rho (A)$
maps each variable in f to
$\frac {1}{2}$
times itself. Hence, by the Leibniz rule,
$\rho (A)$
scales the homogeneous part of degree d in f by
$\frac {d}{2}$
. Since I is preserved by
$\rho (A)$
, it follows that I contains all homogeneous components of f, and hence, I is a homogeneous ideal.
Now let
$B \in \mathfrak {gl}(E_\infty )_n$
and
$f \in I$
be arbitrary. By the previous paragraph, we can assume f to be homogeneous of degree d, and we then have

and since I is
$\rho (B)$
-stable, we deduce
$\tilde {\rho }(B) f \in I$
. This completes the proof in one direction. The proof in the opposite direction is identical.
Remark 4.4. By the proof above, any
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable closed subscheme X of
$(\bigwedge \nolimits _\infty E_\infty )^*$
is an affine cone.
Following [Reference Eggermont and Snowden7], the restricted dual
$(E_\infty )_*$
of
$E_\infty $
is defined as the union
$\bigcup _{n\geq 1}(E_n)^*.$
We will denote by
$\varepsilon ^1, \varepsilon ^2, \dots $
the basis of
$(E_\infty )_*$
that is dual to the canonical basis
$e_1, e_2, \dots $
of
$E_\infty $
given by
$\varepsilon ^i(e_j)=\delta _{ij}$
.
Lemma 4.5. There is an
$\operatorname{\mathrm{SL}} (E_\infty )$
-equivariant isomorphism

which restricts to an isomorphism

We will use this isomorphism to regard
$\bigwedge \nolimits _\infty E_\infty $
as the restricted dual of the Grassmann algebra
$\bigwedge \nolimits E_\infty $
. We stress, though, that this isomorphism is not
$GL(E_\infty )$
-equivariant.
Proof. We have a natural bilinear map

If
$I \subseteq {\mathbb {N}}$
is finite and
$J \subseteq {\mathbb {N}}$
is cofinite, then
$e_I \wedge e_J$
is
$0$
if
$I \cap J \neq \emptyset $
and
$\pm e_{I \cup J}$
otherwise, where the sign is determined by the permutation required to order the sequence
$I,J$
. We then define a perfect pairing
$\gamma $
between the two spaces by

The map
$\Phi _\gamma : \bigwedge \nolimits _\infty E_\infty \to \bigwedge \nolimits (E_\infty )_*, \; \omega ' \mapsto \gamma (\cdot , \omega ')$
induced by
$\gamma $
is the isomorphism given by
$e_I \mapsto \pm \varepsilon ^{I^c}$
, where
$I^c \subseteq {\mathbb {N}}$
is the complement of I and
for a finite set
$J=\{j_1, \dots , j_k\}$
. Note that
$\gamma (A\cdot \omega , A\cdot \omega ') = \det (A)\gamma (\omega , \omega ')$
for all
$A \in \operatorname{\mathrm{GL}} (E_\infty )$
, and hence,
$\gamma $
is
$\operatorname{\mathrm{SL}} (E_\infty )$
-invariant. Therefore, the isomorphism
$\Phi _\gamma $
is
$\operatorname{\mathrm{SL}} (E_\infty )$
-equivariant.
Lemma 4.6. An ideal
$I \subseteq \operatorname{\mathrm{Sym}} (\bigwedge (E_\infty )_*)$
is
$\operatorname{\mathrm{SL}} (E_\infty )$
-stable if and only if it is
$\operatorname{\mathrm{GL}} (E_\infty )$
-stable. The same holds for
$\operatorname{\mathrm{SL}} (E_\infty )_n$
and
$\operatorname{\mathrm{GL}} (E_\infty )_n$
.
Proof. Assume that I is
$\operatorname{\mathrm{SL}} (E_\infty )$
-stable. Let
$f \in I$
and
$A \in \operatorname{\mathrm{GL}} (E_\infty )$
be arbitrary. Choose
$m=m(f,A) \in {\mathbb {N}}$
large enough so that
$f \in \operatorname{\mathrm{Sym}} (\bigwedge \nolimits (E_m)^*)$
and A is the image of some
$A_m \in \operatorname{\mathrm{GL}} (E_m)$
. Define
$A_{m+1} \in \operatorname{\mathrm{GL}} (E_{m+1})$
as the map given by
$A_{m+1}(e_i)=A_m(e_i)$
for
$i \leq m$
and
$A_{m+1}(e_{m+1})=(\det (A_m))^{-1}( e_{m+1})$
, and let
$A^\prime $
be the image of
$A_{m+1}$
in
$\operatorname{\mathrm{SL}} (E_\infty )$
. Then the action of
$A_m$
and
$A_{m+1}$
agree on
$(E_m)^*$
. Hence, they also agree on
$\operatorname{\mathrm{Sym}} (\bigwedge \nolimits (E_m)^*)$
. So
$A \cdot f = A^\prime \cdot f \in I$
since I was assumed to be
$\operatorname{\mathrm{SL}} (E_\infty )$
-stable and
$A^\prime \in \operatorname{\mathrm{SL}} (E_\infty )$
. As
$f \in I$
and
$A \in \operatorname{\mathrm{GL}} (E_\infty )$
were arbitrary, this shows that I is
$GL(E_\infty )$
-stable.
Proof of Proposition 4.2.
First, we claim that
$\operatorname{\mathrm{Spec}} \big (\operatorname{\mathrm{Sym}} \big (\bigoplus _{k=0}^\ell \bigwedge \nolimits ^k (E_\infty )_*\big )\big )$
is topologically
$\operatorname{\mathrm{GL}} (E_\infty )_n$
-Noetherian. Indeed, the standard
$\operatorname{\mathrm{GL}} (E_\infty )$
-representation of the space
$\bigoplus _{k=0}^\ell \bigwedge \nolimits ^k (E_\infty )_*$
is an algebraic representation, and this also remains true when we act with
$\operatorname{\mathrm{GL}} (E_\infty )$
via its isomorphism into
$\operatorname{\mathrm{GL}} (E_\infty )_n$
. Hence, the claim follows from [Reference Eggermont and Snowden7, Theorem 2]. Let
$(X_i)_{i\in {\mathbb {N}}} \subseteq \operatorname{\mathrm{Spec}} (R_{\leq \ell })$
be a descending chain of
$H_n$
-stable, closed, reduced subschemes. By Lemma 4.3, every
$X_i$
is also
$\operatorname{\mathrm{GL}} (E_\infty )_n$
-stable. By Lemma 4.5, there is an
$\operatorname{\mathrm{SL}} (E_\infty )_n$
-equivariant isomorphism
$\operatorname{\mathrm{Spec}} (R_{\leq \ell }) \cong \operatorname{\mathrm{Spec}} \big (\operatorname{\mathrm{Sym}} \big (\bigoplus _{k=0}^\ell \bigwedge \nolimits ^k (E_\infty )_*\big )\big )$
. Let
$X_i^\prime \subseteq \operatorname{\mathrm{Spec}} \big (\operatorname{\mathrm{Sym}} \big (\bigoplus _{k=0}^\ell \bigwedge \nolimits ^k (E_\infty )_*\big )\big )$
be the closed, reduced,
$\operatorname{\mathrm{SL}} (E_\infty )$
-stable subscheme corresponding to
$X_i$
under this isomorphism. Using Lemma 4.6, we see that the subschemes
$X_i^\prime $
are also
$\operatorname{\mathrm{GL}} (E_\infty )_n$
-stable. Therefore, the chain
$(X_i^\prime )_{i\in {\mathbb {N}}}$
stabilizes by our first claim. Consequently, also the chain
$(X_i)_{i\in {\mathbb {N}}}$
stabilizes.
Before we come to the proof of Theorem 4.1, let us recall the action of
$f_i \wedge f_j \in \mathfrak {so}(V_\infty )$
on
$\bigwedge \nolimits ^+_\infty E_\infty $
and its symmetric algebra
$R^+$
in explicit terms. Recall from Section 3.6 that a basis for
$\bigwedge \nolimits ^+_\infty E_\infty $
is given by
$e_I = e_{i_1} \wedge e_{i_2} \wedge \cdots $
, where
$I = \{i_1 < i_2 < \cdots \} \subseteq {\mathbb {N}}$
is cofinite and
$|{\mathbb {N}} \setminus I|$
even. Then we have

where
$c_{i,j}(I)$
depends on the position of
$i,j$
in I. (Note that there is no factor
$4$
, since in our identification of
$\bigwedge \nolimits ^2 V$
to the Lie subalgebra L of
$\operatorname{\mathrm{Cl}} (V)$
we had a factor
$\frac {1}{4}$
.) The corresponding action of
$f_i \wedge f_j$
on polynomials in
$R^+$
is as a derivation.
4.3 Proof of Theorem 4.1
Let
$R^+ \subseteq R$
be the symmetric algebra on the direct half-spin representation
$\bigwedge \nolimits ^+_\infty E_\infty $
, so that
$\operatorname{\mathrm{Spec}} (R^+)$
is the inverse half-spin representation
$(\bigwedge \nolimits ^+_\infty E_\infty )^*$
. We prove topological
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherianity of
$\operatorname{\mathrm{Spec}} (R^+)$
; the corresponding statement for
$\operatorname{\mathrm{Spec}} (R^-)$
is proved in exactly the same manner.
For a closed, reduced
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable subscheme X of
$\operatorname{\mathrm{Spec}} (R^+)$
, we denote by
$\delta _X \in \{0,1,2,\ldots ,\infty \}$
the lowest degree of a nonzero polynomial in the ideal
$I(X) \subseteq R^+$
of X. Here, we consider the natural grading on
$R^+ = \operatorname{\mathrm{Sym}} (\bigwedge \nolimits ^+_\infty E_\infty )$
, where the elements of
$\bigwedge \nolimits ^+_\infty E_\infty $
all have degree
$1$
.
We proceed by induction on
$\delta _X$
to show that X is topologically Noetherian; we may, therefore, assume that this is true for all Y with
$\delta _Y<\delta _X$
. We have
$\delta _X=\infty $
if and only if
$X=\operatorname{\mathrm{Spec}} (R^+)$
. Then a chain

of
$\operatorname {\mathrm {Spin}}(V_\infty )$
-closed subsets is either constant or else there exists an i with
$\delta _{X_i}<\infty $
. Hence, it suffices to prove that X is Noetherian under the additional assumption that
$\delta _X<\infty $
. At the other extreme, if
$\delta _X=0$
, then X is empty and there is nothing to prove. So we assume that
$0<\delta _X<\infty $
and that all Y with
$\delta _Y<\delta _X$
are
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherian.
Let
$p \in R^+$
be a nonzero polynomial in the ideal of X of degree
$\delta _X$
. By Remark 4.4, since X is a cone, p is in fact homogeneous of degree
$\delta _X$
. Let
$e_I$
be a variable appearing in p such that
$k := |I^c|$
is maximal among all variables in p; note that k is even. Then choose
$n \geq k+2$
even such that all variables of p are contained in
$\bigwedge \nolimits ^+ E_n$
(i.e., they are of the form
$e_J$
with
$J \supseteq \{n+1,n+2,\ldots \}$
).
Now act on p with the element
$f_{i_1} \wedge f_{i_2} \in \mathfrak {so}(V_\infty )$
with
$i_1<i_2$
the two smallest elements in I. Since X is
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable, the result
$p_1$
is again in the ideal of X. Furthermore,
$p_1$
has the form

where
$q=\frac {\partial p}{\partial e_{I}}$
contains only variables
$e_J$
with
$|J^c| \leq k$
and where
$r_1$
does not contain
$e_{I \setminus \{i_1,i_2\}}$
but may contain other variables
$e_J$
with
$|J^c|=k+2$
(namely, those with
$i_1,i_2 \not \in J$
for which
$e_{J \cup \{i_1,i_2\}}$
appears in p).
If
$n=k+2$
, then
$I \setminus \{i_1,i_2\}=\{n+1,n+2,\ldots \}$
, and since all variables
$e_J$
in
$p_1$
satisfy
${J \supseteq \{n+1,n+2,\ldots \}}$
,
$e_{I \setminus \{i_1,i_2\}}$
is the only variable
$e_J$
in
$p_1$
with
$|J^c|=k+2$
. If
$n>k+2$
, then we continue in the same manner, now acting with
$f_{i_3} \wedge f_{i_4}$
on
$p_1$
, where
$i_3<i_4$
are the two smallest elements in
$I \setminus \{i_1,i_2\}$
. We write
$p_2$
for the result, which is now of the form

where q is the same polynomial as before and
$r_2$
does not contain the variable
$e_{I \setminus \{i_1,i_2,i_3,i_4\}}$
but may contain other variables
$e_J$
with
$|J^c|=k+4$
.
Iterating this construction, we find the polynomial

in the ideal of X, where
$\ell =(n-k)/2$
, q is the same polynomial as before and
$r_\ell $
only contains variables
$e_J$
with
$|J^c|<n$
. Let
$Z:=X[1/q]$
be the open subset of X where q is nonzero.
Lemma 4.7. For every variable
$e_J$
with
$|J^c| \geq n$
, the ideal of Z in the localisation
$R^+[1/q]$
contains a polynomial of the form
$e_J - s/q^d$
for some
$d \in {\mathbb {Z}}_{\geq 0}$
and some
$s \in R^+_{\leq n-2}$
.
Proof. We proceed by induction on
$|J^c|=:m$
. By successively acting on
$p_\ell $
with the elements
${f_n \wedge f_{n+1},f_{n+2} \wedge f_{n+3},\ldots ,f_{m-1}\wedge f_m}$
, we find the polynomial

in the ideal of X, where r contains only variables
$e_L$
with
$|L^c|<m$
. Now act with elements of
$\mathfrak {gl}(E_\infty )$
to obtain an element

where
$\tilde {r}$
still contains only variables
$e_L$
with
$|L^c|<m$
. Inverting q, this can be used to express
$e_J$
in such variables
$e_L$
. By the induction hypothesis, all those
$e_L$
admit an expression, on Z, as a polynomial in
$R^+_{\leq n-2}$
times a negative power of q. Then the same holds for
$e_J$
.
Lemma 4.8. The open subscheme
$Z=X[1/q]$
is stable under the group
$H_n$
and
$H_n$
-Noetherian.
Proof. By Lemma 4.3, X is stable under
$\operatorname{\mathrm{GL}} (E_\infty )_n$
. The polynomial q is homogeneous and contains only variables
$e_J$
with
$J \supseteq \{n+1,n+2,\ldots \}$
. Every
$g \in \operatorname{\mathrm{GL}} (E_\infty )_n$
scales each such variable with
$\det (g)$
, and hence, maps q to a scalar multiple of itself. We conclude that Z is stable under
$\operatorname{\mathrm{GL}} (E_\infty )_n$
, and hence by (a slight variant of) Lemma 4.3 also under
$H_n$
.
By Lemma 4.7, the projection dual to the inclusion
$R^+_{\leq n-2}[1/q] \subseteq R^+[1/q]$
restricts on Z to a closed embedding, and this embedding is
$H_n$
-equivariant. By Proposition 4.2, the image of Z is
$H_n$
-Noetherian, and hence, so is Z itself.
Proof of Theorem 4.1.
Let

be a chain of reduced,
$\operatorname {\mathrm {Spin}}(V_{\infty })$
-stable closed subschemes. Let
$Y \subseteq X$
be the reduced closed subscheme defined by the orbit
$\operatorname {\mathrm {Spin}}(V_{\infty }) \cdot q$
. Since q has degree
$\delta _X-1$
, we have
$\delta _Y < \delta _X$
, and hence, Y is
$\operatorname {\mathrm {Spin}}(V_{\infty })$
-Noetherian by the induction hypothesis. It follows that the chain

is eventually stable. However, the chain

consists of reduced,
$H_n$
-stable closed subschemes of Z; hence, it is eventually stable by Lemma 4.8.
Now pick a (not necessarily closed) point
$P \in X_i$
for
$i \gg 0$
. If
$P \in Y \cap X_i$
, then
$P \in Y \cap X_{i-1}$
by the first stabilisation. However, if
$P \not \in Y \cap X_i$
, then there exists a
$g \in \operatorname {\mathrm {Spin}}(V_\infty )$
such that
$gP \in Z$
. Then
$gP$
lies in
$X_i \cap Z$
, which by the second stabilisation equals
$X_{i-1} \cap Z$
; hence,
$P=g^{-1}(gP)$
lies in
$X_{i-1}$
, as well. We conclude that the chain
$(X_i)_i$
of closed, reduced subschemes of X stabilises. Hence, the inverse half-spin representation
$(\bigwedge \nolimits ^+_\infty E_\infty )^*$
is topologically
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherian.
Remark 4.9. While the proof of Theorem 4.1 for the even half-spin case is easily adapted to a proof for the odd half-spin case, we do not know whether the spin representation
$(\bigwedge \nolimits _\infty E_\infty )^*$
itself is topologically
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherian! Also, despite much effort, we have not succeeded in proving that the inverse limit
$\varprojlim _n \bigwedge \nolimits ^n V_n$
along the contraction maps
$c_{e_n}$
is topologically
$\operatorname {\mathrm {SO}}(V_\infty )$
-Noetherian. Indeed, the situation is worse for this question: like the inverse spin representation, this limit is the dual of a countable-dimensional module that splits as a direct sum of two
$\operatorname {\mathrm {SO}}(V_\infty )$
-modules – and here, we do not even know whether the dual of one of these modules is topologically Noetherian!
5 Half-spin varieties and applications
In this section, we introduce the notion of half-spin varieties and reformulate our main result Theorem 4.1 in this language. We start by fixing the necessary data determining the half-spin representations of
$\operatorname {\mathrm {Spin}}(V)$
.
Notation 5.1. As shorthand, we write
$\mathcal {V}=(V,q,F) \in \mathcal {Q}$
to refer to a triple where
-
1. V is an even-dimensional vector space over K,
-
2. q is a nondegenerate symmetric quadratic form on V, and
-
3. F is a maximal isotropic subspace of V.
An isomorphism
$\mathcal {V} \to \mathcal {V}'=(V',q',F')$
of such triples is a linear bijection
$\varphi : V \to V'$
with
$q'(\varphi (v))=q(v)$
and
$\varphi (F)=F'$
.
Given a triple
$\mathcal {V}$
, we have half-spin representations
$\operatorname{\mathrm{Cl}} ^{\pm }(V)f$
, where
$f=f_1 \cdots f_n$
with
$f_1,\ldots ,f_n$
a basis of F (recall that the left ideal
$\operatorname{\mathrm{Cl}} (V)f$
does not depend on this basis). Half-spin varieties are
$\operatorname {\mathrm {Spin}}(V)$
-invariant subvarieties of these half-spin representations that are preserved by the contraction maps
$\pi _e$
from §3.2 and the multiplication maps
$\tau _h$
from §3.3. The precise definition below is inspired by that of a Plücker variety in [Reference Draisma and Eggermont6]. It involves a uniform choice of either even or odd half-spin representations. For convenience of notation, we will only explicitly work with the even half-spin representations, but all further results are valid for the odd counterparts as well.
Definition 5.2 (Half-spin variety).
A half-spin variety is a rule X that assigns to each triple
${\mathcal {V}=(V,q,F) \in \mathcal {Q}}$
a closed, reduced subscheme
$X(\mathcal {V}) \subseteq \operatorname{\mathrm{Cl}} ^+(V)f$
such that
-
1.
$X(\mathcal {V})$ is
$\operatorname {\mathrm {Spin}}(V)$ -stable;
-
2. for any isomorphism
$\varphi :\mathcal {V} \to \mathcal {V}'$ , the map
$\operatorname{\mathrm{Cl}} ^+(\varphi )$ maps
$X(\mathcal {V})$ into
$X(\mathcal {V}')$ ;
-
3. for any isotropic
$e \in V$ with
$e \not \in F$ , if we set
$V':=e^\perp /\langle e \rangle $ ,
$q'$ the induced form on
$V'$ ,
$F'$ the image of
$F \cap e^\perp $ in
$V'$ , and
$\mathcal {V}':=(V',q',F')$ , then the contraction map
$\pi _e:\operatorname{\mathrm{Cl}} ^+(V)f \to \operatorname{\mathrm{Cl}} ^+(V')f'$ maps
$X(\mathcal {V})$ into
$X(\mathcal {V}')$ ; and
-
4. for any
$\mathcal {V}=(V,q,F)$ , if we construct a triple
$\mathcal {V}'$ by setting
$V':=V \oplus \langle e,h \rangle $ ,
$q'$ the quadratic form that restricts to q on V, makes the direct sum orthogonal, and makes
$e,h$ a hyperbolic basis, if we set
$f':=f \cdot h$ , then the map
$\tau _h: \operatorname{\mathrm{Cl}} ^+(V)f \to \operatorname{\mathrm{Cl}} ^+(V')f'$ maps
$X(\mathcal {V})$ into
$X(\mathcal {V}')$ .
Example 5.3. The following are examples of half-spin varieties.
-
1. Trivially,
$X(\mathcal {V}) := \operatorname{\mathrm{Cl}} ^+(V)f$ ,
$X(\mathcal {V}) := \{0\}$ and
$X(\mathcal {V}) := \emptyset $ define half-spin varieties.
-
2. The even component of the cone over the isotropic Grassmannian,
$X(\mathcal {V}):=\widehat {\operatorname{\mathrm{Gr}} }_{\operatorname {\mathrm {iso}}}^+(V,q)$ is a half-spin variety by Proposition 3.4.
-
3. For two half-spin varieties X and
$X'$ , their join
$X + X'$ defined by
$$\begin{align*}(X + X')(\mathcal{V}) := \overline{\{x+x' \mid x \in X(\mathcal{V}), x' \in X'(\mathcal{V})\}} \end{align*}$$
-
4. The intersection of two half-spin varieties X and
$X'$ is a half-spin variety, which is defined by
$(X \cap X')(\mathcal {V}) := X(\mathcal {V}) \cap X'(\mathcal {V})$ .
Similar as in §3.6, we will use the following notation: for every
$n \in {\mathbb {N}}$
, we consider the vector space
$V_n = \langle e_1,\dots , e_n,f_1,\dots ,f_n\rangle $
together with the quadratic form
$q_n$
whose corresponding bilinear form
$(\cdot |\cdot )$
satisfies

Furthermore, let
$E_n=\langle e_1,\dots ,e_n\rangle $
and
$F_n=\langle f_1,\dots ,f_n\rangle $
; these are maximal isotropic subspaces of
$V_n$
. We will denote the associated tuple by
$\mathcal {V}_n=(V_n,q_n,F_n)$
.
Remark 5.4. A half-spin variety X is completely determined by the values
$X(\mathcal {V}_n)$
; that is, if X and
$X'$
are half-spin varieties such that
$X(\mathcal {V}_n) = X'(\mathcal {V}_n)$
for all
$n \in {\mathbb {N}}$
, then
$X(\mathcal {V}) = X'(\mathcal {V})$
for all tuples
$\mathcal {V}$
.
We now want to associate to each half-spin variety X an infinite-dimensional scheme
$X_\infty $
embedded inside the inverse half-spin representation
$(\bigwedge \nolimits ^+_\infty E_\infty )^*$
as follows. Since
$V_n = E_n \oplus F_n$
, we can use the isomorphism from §2.5 to embed
$X(\mathcal {V}_n)$
as a reduced subscheme of
$\bigwedge ^+ E_n$
(recall from §3.7 that we view
$\bigwedge ^+ E_n$
as the affine scheme with coordinate ring
$\textrm { Sym}(\bigwedge \nolimits ^{+(-1)^n}E_n)$
). We abbreviate
.
For
$N \geq n$
, let
$\pi _{N,n}: \bigwedge \nolimits ^+E_N \to \bigwedge \nolimits ^+E_n$
, resp.
$\tau _{n,N}: \bigwedge \nolimits ^+E_n \to \bigwedge \nolimits ^+E_N$
be the maps induced by canonical projection
$E_N \to E_n$
, resp. by the injection
$E_n \hookrightarrow E_N$
. Note that
$\tau _{n,N}$
is a section of
$\pi _{N,n}$
. Recall that
$(\bigwedge \nolimits ^+_\infty E_\infty )^* = \varprojlim _n \bigwedge ^+E_n$
. We denote the structure maps by
$\pi _{\infty ,n}: (\bigwedge \nolimits ^+_\infty E_\infty )^* \to \bigwedge \nolimits ^+E_n$
and by
$\tau _{n,\infty }: \bigwedge \nolimits ^+E_n \to (\bigwedge \nolimits ^+_\infty E_\infty )^*$
the inclusion maps induced by
$\tau _{n,N}$
.
From the definition of a half-spin variety, it follows that

Hence, the inverse limit

is well defined, and a closed, reduced,
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable subscheme of
$(\bigwedge \nolimits ^+_\infty E_\infty )^*$
. In order to see this, write
and
. Let
$I_n \subseteq R_n$
be the radical ideal associated to
$X_n \subseteq \operatorname{\mathrm{Spec}} (R_n)$
(i.e.,
$X_n = V(I_n) = \operatorname{\mathrm{Spec}} (R_n/I_n)$
). As
$\operatorname{\mathrm{Spec}} (\cdot )$
is a contravariant equivalence of categories, it holds that

So
$X_\infty $
corresponds to the ideal
. As all
$I_n \subseteq R_n$
are radical, so is
$I_\infty \subseteq R_\infty $
, and therefore,
$X_\infty $
is a reduced subscheme.
It follows from Equation (5.1) that

Lemma 5.5. The mapping

is injective. That is, if X and
$X'$
are half-spin varieties such that
$X_\infty = X^{\prime }_\infty $
, then
$X = X'$
(i.e.,
$X(\mathcal {V}) = X'(\mathcal {V})$
for all tuples
$\mathcal {V}$
).
Proof. Note that, for all
$n \in {\mathbb {N}}$
, it holds that

Indeed, the inclusion
$\supseteq $
is contained in Equation (5.2), and the other direction
$\subseteq $
follows from the fact that
$\tau _{n,\infty }:X_n \to X_\infty $
is a section of
$\pi _{\infty ,n}$
. Hence, if
$X_\infty = X^{\prime }_\infty $
, then

By Remark 5.4, this shows that
$X = X'$
.
For two half-spin varieties X and
$X'$
, we will write
$X \subseteq X'$
if
$X(\mathcal {V}) \subseteq X'(\mathcal {V})$
for all
$\mathcal {V} = (V,q,F)$
. Theorem 4.1 then implies the following.
Theorem 5.6 (Noetherianity of half-spin varieties).
Every descending chain of half-spin varieties

stabilizes; that is, there exists
$m_0 \in {\mathbb {N}}$
such that
$X^{(m)} = X^{(m_0)}$
for all
$m \geq m_0$
.
Proof. Note that the mapping
$X \mapsto X_\infty $
is order preserving; that is, if
$X \subseteq X'$
, then
$X_\infty \subseteq X^{\prime }_\infty $
. Hence, a chain

of half-spin varieties induces a chain

of closed, reduced,
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable subschemes in
$(\bigwedge \nolimits ^+_\infty E_\infty )^*$
. By Theorem 4.1, we know that
$(\bigwedge \nolimits ^+_\infty E_\infty )^*$
is topologically
$\operatorname {\mathrm {Spin}}(V_\infty )$
-Noetherian. Hence, the chain
$X^{(m)}_\infty $
stabilizes. But then, by Lemma 5.5, also the chain of half-spin varieties
$X^{(m)}$
stabilizes. This completes the proof.
As a consequence, we obtain the next results, which state how
$X_\infty $
is determined by the data coming from some finite level of X.
Theorem 5.7. Let X be a half-spin variety. Then there exists
$n_0 \in {\mathbb {N}}$
such that

where
$\operatorname{\mathrm{rad}} ( \operatorname {\mathrm {Spin}}(V_\infty ) \cdot I_{n_0} ) \subseteq \operatorname{\mathrm{Sym}} (\bigwedge \nolimits ^+_\infty E_\infty )$
is the radical ideal generated by the
$\operatorname {\mathrm {Spin}}(V_\infty )$
-orbits of the ideal
$I_{n_0} \subseteq \operatorname{\mathrm{Sym}} (\bigwedge ^{+(-1)^{n_0}}E_{n_0})$
defining
$X_{n_0} \subseteq \bigwedge ^+E_{n_0}$
.
Proof. For each
$n \in {\mathbb {N}}$
, set
. We denote by
$I_\infty \subseteq \operatorname{\mathrm{Sym}} (\bigwedge \nolimits ^+_\infty E_\infty )$
the ideal associated to
$X_\infty $
. This ideal is
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable, radical and it holds that
$I_\infty = \varinjlim _n I_n$
. Thus,
$\bigcup _n J_n = I_\infty $
. Since
$(J_n)_{n\in {\mathbb {N}}}$
is an increasing chain of closed
$\operatorname {\mathrm {Spin}}(V_\infty )$
-stable radical ideals, by Theorem 4.1, there exists
$n_0 \in {\mathbb {N}}$
such that
$J_n=J_{n_0}$
for all
$n\geq n_0$
. Therefore,
$I_\infty = \bigcup _n J_n = J_{n_0}$
and hence,
$X_\infty =V(I_\infty )=V(J_{n_0}).$
Corollary 5.8. Let X be a half-spin variety. There exists
$n_0 \in {\mathbb {N}}$
such that for all
$n\geq n_0$
, it holds that

Proof. Take
$n_0$
as in Theorem 5.7. Then the statement follows from that theorem and [Reference Draisma4, Lemma 2.1]. To apply that lemma, we must check condition (*) in that paper – namely, that for
$q \geq n \geq n_0$
and
$g \in \operatorname {\mathrm {Spin}}(V_q)$
, we can write

for suitable
$m \leq n_0$
and
$g' \in \operatorname {\mathrm {Spin}}(V_n)$
and
$g" \in \operatorname {\mathrm {Spin}}(V_{n_0})$
. In fact, since half-spin varieties are affine cones, it suffices that this identity holds up to a scalar factor. It also suffices to prove this for g in an open dense subset U of
$\operatorname {\mathrm {Spin}}(V_q)$
, because the equations for
$X_{n_0}$
pulled back along the map on the left for
$g \in U$
imply the equations for all g. We will prove this, with
$m=n_0$
, using the Cartan map in Lemma 6.6 below.
6 Universality of
$\widehat{\mathrm{Gr}} ^+_{{\mathrm {iso}}}(4,8)$
and the Cartan map
6.1 Statement
In [Reference Seynnaeve and Tairi16], the last two authors showed that in even dimension, the isotropic Grassmannian in its Plücker embedding is set-theoretically defined by pulling back equations coming from
$\widehat {\operatorname{\mathrm{Gr}} }_{\operatorname {\mathrm {iso}}}(4,8)$
. Using the Cartan map, we can translate this into a statement about the isotropic Grassmannian in its spinor embedding and prove the following result.
Theorem 6.1. For all
$n\geq 4$
, we have

where
$I_4$
is the ideal of polynomials defining
$\widehat {\operatorname{\mathrm{Gr}} }^+_{\operatorname {\mathrm {iso}}}(V_4) \subseteq \operatorname{\mathrm{Cl}} ^+(V_4)f.$
In other words, the bound
$n_0$
from Corollary 5.8 can be taken equal to
$4$
for the cone over the isotropic Grassmannian. We give the proof of Theorem 6.1 in §6.5 using properties of the Cartan map that will be established in the following sections.
6.2 Definition of the Cartan map
When we regard
$e_1 \wedge \cdots \wedge e_n$
as an element of the n-th exterior power
$\bigwedge \nolimits ^n V$
of the standard representation V of
$\mathfrak {so}(V)$
, then it is a highest weight vector of weight
$(0,\ldots ,0,2)=2\lambda _0$
, where
$\lambda _0$
is the fundamental weight introduced in §2.6 and the highest weight of the half-spin representation
$\operatorname{\mathrm{Cl}} ^{(-1)^n}(V) f$
. Similarly, the element
$e_1 \wedge e_2 \wedge \cdots \wedge e_{n-1} \wedge f_n \in \bigwedge \nolimits ^n V$
is a highest weight vector of weight
$(0,\ldots ,0,2,0)=2 \lambda _1$
, where
$\lambda _1$
is the highest weight of the other half-spin representation. So
$\bigwedge \nolimits ^n V$
contains copies of the irreducible representations
$V_{2\lambda _0},V_{2 \lambda _1}$
of
$\mathfrak {so}(V)$
; in fact, it is well known to be the direct sum of these. To compare our results in this paper about spin representations with earlier work by the last two authors about exterior powers, we will need the following considerations.
Consider any connected, reductive algebraic group G, with maximal torus T and Borel subgroup
$B \supseteq T$
. Let
$\lambda $
be a dominant weight of G, let
$V_{\lambda }$
be the corresponding irreducible representation, and let
$v_\lambda \in V_{\lambda }$
be a nonzero highest-weight vector (which is unique up to scalar multiples). Then the symmetric square
$S^2 V_\lambda $
contains a one-dimensional space of vectors of weight
$2 \lambda $
, spanned by
$v_{2\lambda }:=v_\lambda ^2$
. This vector is itself a highest-weight vector, and hence generates a copy of
$V_{2\lambda }$
; this is sometimes called the Cartan component of
$S^2 V_\lambda $
. By semisimplicity, there is a G-equivariant linear projection
$\pi : S^2 V_\lambda \to V_{2\lambda }$
that restricts to the identity on
$V_{2\lambda }$
. The map

is a nonzero polynomial map, homogeneous of degree
$2$
, and hence induces a rational map
$\nu _2:\mathbb {P} V_{\lambda } \to \mathbb {P} V_{2\lambda }$
. Note that this is the composition of the quadratic Veronese embedding and the projection
$\pi $
. We will refer to
$\nu _2$
and to
$\widehat {\nu _2}$
as the Cartan map.
Lemma 6.2. The rational map
$\nu _2$
is a morphism and injective.
We thank J. M. Landsberg for help with the following proof.
Proof. To show that
$\nu _2$
is a morphism, we need to show that
$\pi (v^2)$
is nonzero whenever v is. Now the set Q of all
$[v] \in \mathbb {P} V_\lambda $
for which
$\pi (v^2)$
is zero is closed and B-stable. Hence, if
$Q \neq \emptyset $
, then by Borel’s fixed point theorem, Q contains a B-fixed point. But the only B-fixed point in
$\mathbb {P} V_\lambda $
is
$[v_\lambda ]$
, and
$v_\lambda $
is mapped to the nonzero vector
$v_{2\lambda }$
. Hence,
$Q=\emptyset $
.
Injectivity is similar but slightly more subtle. Assume that there exist distinct
$[v],[w]$
with
$\nu _2([v])=\nu _2([w])$
. Then
$\{[v],[w]\}$
represents a point in the Hilbert scheme H of two points in
$\mathbb {P} V_\lambda $
. Now the locus Q of points S in H such that
$\nu _2(S)$
is contained in a single reduced point is closed in H, as it is the projection to H of the closed subvariety

and since
$\mathbb {P} V_{2 \lambda }$
is projective. Since H is a projective scheme, Q is projective. Hence, Q contains a B-fixed point
$S_0$
. This scheme
$S_0$
cannot consist of two distinct reduced points: if it did, then either both points would be B-fixed, but there is only one B-fixed point, or else they would be a B-orbit, but this is impossible since B is connected. Therefore, the reduced subscheme of
$S_0$
is
$\{[v_{\lambda }]\}$
, and
$S_0$
represents the point
$[v_\lambda ]$
together with a nonzero tangent direction in
$T_{[v_\lambda ]} \mathbb {P} V_{\lambda } = V_\lambda /K v_\lambda $
, represented by
$w \in V_\lambda $
. Furthermore, B-stability of
$S_0$
implies that the B-module generated by w equals
$\langle w,v_\lambda \rangle _K$
. That
$S_0$
lies in Q means that

We find that
$\pi (v_\lambda w)=0$
, so that the G-module generated by
$v_\lambda w \in S^2 V$
does not contain
$V_{2\lambda }$
. But since
$v_\lambda $
is (up to a scalar) fixed by B, the B-module generated by
$v_\lambda w$
equals
$v_\lambda $
times the B-module generated by w, and hence contains
$v_\lambda ^2=v_{2\lambda }$
, a contradiction.
Observe that
$\nu _2$
maps the unique closed orbit
$G \cdot [v_\lambda ]$
in
$\mathbb {P} V_\lambda $
isomorphically to the unique closed orbit
$G \cdot [v_{2\lambda }]$
– both are isomorphic to
$G/P$
, where
$P \supseteq B$
is the stabiliser of the line
$K v_\lambda $
and of the line
$K v_{2\lambda }$
. In our setting above, where
$G=\operatorname {\mathrm {Spin}}(V)$
and
$\lambda \in \{\lambda _0,\lambda _1\}$
, the closed orbit
$G \cdot [v_{2\lambda }]$
is one of the two connected components of the Grassmannian
$\operatorname{\mathrm{Gr}} _{\operatorname {\mathrm {iso}}}(V)$
of n-dimensional isotropic subspaces of V, in its Plücker embedding; and the closed orbit in the projectivised half-spin representation
$\mathbb {P} V_{\lambda }$
is the same component of the isotropic Grassmannian but now in its spinor embedding.
In what follows, we will need a more explicit understanding both of the embedding of the isotropic Grassmannian in the projectivised (half-)spin representations and of the map
$\widehat {\nu _2}$
. These are treated in the next two paragraphs.
6.3 The map
$\widehat {\nu _2}$
from the spin representation to the exterior power
In §6.2, we argued the existence of
$\operatorname {\mathrm {Spin}}(V)$
-equivariant quadratic maps from the half-spin representations to the two summands of
$\bigwedge \nolimits ^n V$
. In [Reference Manivel11], these two maps are described jointly as

where
$\bullet $
stands for the
$\operatorname{\mathrm{Cl}} (V)$
-module structure of
$\bigwedge \nolimits V$
from §2.2 and
$a^*$
refers to the anti-automorphism of the Clifford algebra from §2.1.
Lemma 6.3. The map
$\hat {\nu }_2$
maps the isotropic Grassmann cone in its spinor embedding to the isotropic Grassmann cone in its Plücker embedding, that is,

where
$\widehat {\operatorname{\mathrm{Gr}} }^{\operatorname {Pl}}_{\operatorname {\mathrm {iso}}}(V)$
is the isotropic Grassmann cone in its Plücker embedding (see [Reference Seynnaeve and Tairi16, Definition 3.7]).
Proof. Let
$H \subseteq V$
be a maximal isotropic subspace that intersects F in a k-dimensional space. Choose a hyperbolic basis
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
adapted to H and F, so that
$H=\langle e_{k+1},\ldots ,e_n,f_1,\ldots ,f_k \rangle $
is represented by the vector
$\omega _H:= e_{k+1} \cdots e_n f \in \widehat {\operatorname{\mathrm{Gr}} }_{\operatorname {\mathrm {iso}}}(V)$
where
$f=f_1 \cdots f_n$
; see §3.1. Set
${a:=e_{k+1} \cdots e_n}$
. Now

where we have used the definition of
$\operatorname{\mathrm{Cl}} (V)$
(in the first step), the fact that the second copy of
$e_n$
is perpendicular to all elements before it and multiplies to zero with the first copy of
$e_n$
(in the second step), and have repeated this another
$n-k-1$
times in the last step. We now find that

so that
$(afa^*) \bullet 1$
lies in one of the two summands of
$\bigwedge \nolimits ^n V$
and spans the line representing the space H in the Plücker embedding. This shows that
$\hat {\nu _2}$
maps the isotropic Grassmann cone in its spinor embedding to the isotropic Grassmann cone in its Plücker embedding, as desired.
Remark 6.4. While the spin representation
$\operatorname{\mathrm{Cl}} (V)f$
depends only on the space F – since F determines f up to a scalar, which does not affect the left ideal
$\operatorname{\mathrm{Cl}} (V)f$
– the map
$\widehat {\nu _2}$
actually depends on f itself: for
$\tilde {f}:=t f$
with
$t \in K^*$
, the map
$\widehat {\nu _2}$
constructed from
$\tilde {f}$
sends
$af= (t^{-1}a)\tilde {f}$
to
$t^{-1}a \tilde {f} t^{-1}a^*=t^{-1}a f a^*$
, so the new
$\widehat {\nu _2}$
is
$t^{-1}$
times the old map.
6.4 Contraction and the Cartan map commute
Recall from §6.2 that we have quadratic maps
$\widehat {\nu _2}$
from the half-spin representations to the two summands of
$\bigwedge \nolimits ^n V$
; together, these form a quadratic map
$\widehat {\nu _2}$
which we discussed in §6.3. By abuse of terminology, we call this, too, the Cartan map. Given an isotropic vector
$e \in V$
that is not in F, we write
$\widehat {\nu _2}$
also for the Cartan map
$\operatorname{\mathrm{Cl}} (V_e)\overline {f} \to \bigwedge \nolimits ^{n-1} V_e$
(notation as in §3.2). Recall from §3.2 the contraction map
$c_e: \bigwedge \nolimits ^n V \to \bigwedge \nolimits ^{n-1} V_e$
and its spin analogue
$\pi _e: \operatorname{\mathrm{Cl}} (V)f \to \operatorname{\mathrm{Cl}} (V_e)\overline {f}$
. Also, for a fixed
$h=f_n \in F$
with
$(e|h)=1$
, recall from §3.3 the multiplication map
$m_h:\bigwedge \nolimits ^{n-1} V_e \to \bigwedge \nolimits ^n V$
and its spin analogue
$\tau _h:\operatorname{\mathrm{Cl}} (V_e) \overline {f} \to \operatorname{\mathrm{Cl}} (V) f$
. The relations between these maps are as follows.
Proposition 6.5. The following diagrams essentially commute:

More precisely, one can rescale the restrictions of
$c_e$
to the two
$\mathfrak {so}(V)$
-submodules of
$\bigwedge \nolimits ^n V$
each by
$\pm 1$
in such a manner that the diagram commutes, and similarly for
$m_h$
.
Naturally, we could have chosen the scalars in the definition of
$c_e$
(or, using a square root of
$-1$
, in that of
$\pi _e$
) such that the diagram literally commutes. However, we have chosen the scalars such that
$c_e$
has the most natural formula and
$\pi _e,\tau _h$
have the most natural formulas in our model
$\bigwedge \nolimits E$
for the spin representation.
Proof. We may choose a hyperbolic basis
$e_1,\ldots ,e_n,f_1,\ldots ,f_n$
of V such that
$e=e_n$
and
$f_1,\ldots ,f_n$
is a basis of F. We write
$f:=f_1 \cdots f_n$
and
$\overline {f}:=\overline {f}_1 \cdots \overline {f}_{n-1}$
.
Since the vertical maps are quadratic, it is not sufficient to show commutativity on a spanning set. We therefore consider

where, for
$I=\{i_1<\ldots <i_k\}$
, we write
$e_I:=e_{i_1} \cdots e_{i_k}$
. We then have

and

Now note that, since
$\overline {f}$
has
$n-1$
factors, if
$I,J$
do not have the same parity, then acting with
$\overline {e_I} \overline {f} \overline {e_J}^*$
on
$1$
yields a zero contribution in
$\bigwedge \nolimits ^{n-1} V_e$
. Hence, the sum above may be split into two sums, one of which is

However, consider

For the same reason as above, this splits into two sums, and we want to compare the following expression to (6.2):

Now recall that the action of
$e=e_n \in V \subseteq \operatorname{\mathrm{Cl}} (V)$
on
$\bigwedge \nolimits V$
is via
$o(e)+\iota (e)$
, while
$c_e$
is
$\iota _e$
followed by projection to
$\bigwedge \nolimits ^{n-1} V_e$
. Hence, to compute (6.3), we may as well compute the summands of

that do not contain a factor e. Terms with
$n \in I$
do not contribute because then
$e e_I=0$
. Terms with
$n \not \in I$
but
$n \in J$
do not contribute because when e gets contracted with
$f_n$
a factor e in
$e_J^*$
survives, and when e does not get contracted with
$f_n$
, we use
$ee_J^*=0$
. So we may restrict attention to the terms with
$n \not \in I \cup J$
. Let
$I,J$
correspond to such a term; that is,
$|I|,|J|$
are even and
$n \not \in I \cup J$
. Write
$I=\{i_1<\ldots <i_k\}$
and
$J=\{j_1<\ldots <j_l\}$
. Then

The second term in the last expression will contribute only terms with a factor e to the final result, and the former term contributes

Comparing this with (6.2), we see that the diagram commutes on terms in
$\operatorname{\mathrm{Cl}} ^+(V)f$
up to the factor
$(-1)^{n-1}$
. A similar computation shows that it commutes on terms in
$\operatorname{\mathrm{Cl}} ^-(V)f$
up to a factor factor
$(-1)^n$
.
We now consider the second diagram, where V is split as the orthogonal direct sum
$V_e \oplus \langle e,h \rangle $
with
$e=e_n,h=f_n$
. Consider
$a \in \operatorname{\mathrm{Cl}} (\langle e_1,\ldots ,e_{n-1} \rangle )$
. By the same argument as above, it suffices to consider the case where all summands of a in the basis
$e_I$
have indices I with
$|I|$
of the same parity, say even. Then
$\widehat \nu _2 \circ \tau _h$
in the diagram sends
$a\overline {f}$
to the component in
$\bigwedge \nolimits ^n V$
of
$afa^* \bullet 1$
. Since the summands
$e_I$
in a all have
$n \not \in I$
, in
$a f a^* \bullet 1$
all summands have a factor
$f_n$
, and indeed,

(when all terms in a have
$|I|$
odd, we get a minus sign). The component in
$\bigwedge \nolimits ^n V$
of this expression is the same as the one obtained via
$m_h \circ \widehat \nu _2$
.
6.5 Proof of Theorem 6.1
In this section, we use the Cartan map to prove Corollary 6.1, and finish the proof of Corollary 5.8 via a similar argument.
Proof of Theorem 6.1.
For a quadratic space of dimension
$2n$
, denote by
$\widehat {\operatorname{\mathrm{Gr}} }^{\operatorname {Pl}}_{\operatorname {\mathrm {iso}}}(V) \subseteq \bigwedge \nolimits ^n V$
the isotropic Grassmann cone over the Plücker embedding. Given a maximal isotropic subspace
$F \subseteq V$
with basis
$f_1,\ldots ,f_n$
and
$f:=f_1 \cdots f_n$
, let
$\hat {\nu }_2:\operatorname{\mathrm{Cl}} ^+(V)f \to \bigwedge \nolimits ^n V$
be the Cartan map defined in §6.3. For any isotropic
$v \in V \setminus F$
, the diagram

commutes up to scalar factor at the bottom by Proposition 6.5, where
$V_v:=v^\perp /\langle v \rangle $
and where
$\overline {f}$
is the image of a product of a basis of
$v^\perp \cap F_n$
.
The proof of [Reference Seynnaeve and Tairi16, Corollary 4.2] shows that for
$\omega \in \bigwedge \nolimits ^n V$
, the following are equivalent:
-
1.
$\omega \in \widehat {\operatorname{\mathrm{Gr}} }^{\textrm { Pl}}_{\operatorname {\mathrm {iso}}}(V)$ ;
-
2. For every sequence of isotropic vectors
$v_1 \in V$ ,
$v_2 \in V_{v_1}$ ,
$v_3 \in (V_{v_1})_{v_2}, \dots , v_{n-4} \in (\cdots ((V_{v_1})_{v_2})_{v_3}\cdots )_{v_{n-3}}$ , it holds
$$\begin{align*}C(\omega) \in \widehat{\operatorname{\mathrm{Gr}}}^{\textrm{ Pl}}_{\operatorname{\mathrm{iso}}}(W), \end{align*}$$
$W:=(\cdots ((V_{v_1})_{v_2})_{v_3}\cdots )_{v_{n-4}}$ and
$C:\bigwedge \nolimits ^n V \to \bigwedge \nolimits ^{4}W$ is the composition
$C:= c_{v_{n-4}} \circ \cdots \circ c_{v_1}$ of the contraction maps
$c_{v_i}$ introduced in Section 3.2.
By slight abuse of notation, we also write
$v_1,\ldots ,v_{n-4}$
for preimages of these vectors in V. These span an
$(n-4)$
-dimensional isotropic subspace U of V (provided that each
$v_i$
chosen above in the successive quotients is nonzero), and W equals
$U^\perp / U$
. For any fixed
$\omega $
, the condition that
$C(\omega )$
lies in
$\widehat {\operatorname{\mathrm{Gr}} }^{\textrm { Pl}}_{\operatorname {\mathrm {iso}}}(W)$
is a closed condition on U, and hence, it suffices to check that condition for U in a dense subset of the Grassmannian of isotropic
$(n-4)$
-dimensional subspaces of V. In particular, it suffices to check this when
$U \cap F_n=\{0\}$
.
Fix
$n \geq 4$
and
$x \in \operatorname{\mathrm{Cl}} (V_n)f_1 \cdots f_n$
such that
$p(g \cdot x)=0$
for all
$g \in {\operatorname {\mathrm {Spin}}}(V_n)$
and all
$p \in I_4$
. This means precisely that
$\pi _{n,4}(g\cdot x) \in \widehat {\operatorname{\mathrm{Gr}} }^{+}_{\operatorname {\mathrm {iso}}}(V_4)$
for all
$g \in {\operatorname {\mathrm {Spin}}}(V_n)$
. We need to show that
$x \in \widehat {\operatorname{\mathrm{Gr}} }^{+}_{\operatorname {\mathrm {iso}}}(V_n)$
. To this end, consider
$\omega :=\hat {\nu }_2(x) \in \bigwedge \nolimits ^nV_n$
. It suffices to show that
$\omega \in \widehat {\operatorname{\mathrm{Gr}} }^{\operatorname {Pl}}_{\operatorname {\mathrm {iso}}}(V_n)$
. Indeed, this follows from the fact that
$\hat {\nu }_2\big ( \widehat {\operatorname{\mathrm{Gr}} }^{+}_{\operatorname {\mathrm {iso}}}(V)\big )$
is one of the two irreducible components of
$\widehat {\operatorname{\mathrm{Gr}} }^{\operatorname {Pl}}_{\operatorname {\mathrm {iso}}}(V)$
(see Example 6.3) and because
$\nu _2$
is an injective morphism by Lemma 6.2. Let
$v_1,v_2,\ldots ,v_{n-4} \in V_n$
as above: linearly independent, and such that the span
$U:=\langle v_1,\ldots ,v_{n-4} \rangle $
is an isotropic space that intersects
$F_n$
trivially. Let
$C:= c_{v_{n-4}} \circ \cdots \circ c_{v_1}$
be the composition of the associated contractions. We need to show that
$ C(\omega ) \in \widehat {\operatorname{\mathrm{Gr}} }^{\operatorname {Pl}}_{\operatorname {\mathrm {iso}}}(W)$
, where
$W:=U^\perp /U$
.
Now
$\hat {\nu }_2\big ( \widehat {\operatorname{\mathrm{Gr}} }^{+}_{\operatorname {\mathrm {iso}}}(W)\big ) \subseteq \widehat {\operatorname{\mathrm{Gr}} }^{\textrm { Pl}}_{\operatorname {\mathrm {iso}}}(W)$
by Example 6.3, and the diagram

where
$\overline {f}$
is the image of the product of a basis of
$U^\perp \cap F_n$
, commutes up to a scalar factor in the bottom map due to Proposition 6.5. Hence, it suffices to check that
$\pi _{v_{n-4}} \circ \cdots \circ \pi _{v_1}(x) \in \widehat {\operatorname{\mathrm{Gr}} }^{+}_{\operatorname {\mathrm {iso}}}(W)$
. Now there exists an element
$g \in \operatorname {\mathrm {Spin}}(V_n)$
that maps
$F_n$
into itself (not with the identity!) and sends
$v_i$
to
$e_{n+1-i}$
for
$i=1,\ldots ,n-4$
. This induces an isometry
$W:=U^\perp / U \to (U')^\perp / U'= V_4 = \langle e_1,\ldots ,e_4,f_1,\ldots ,f_4 \rangle $
, where
$U':=\langle e_5,\ldots ,e_n \rangle $
. This in turn induces a linear isomorphism (unique up to a scalar)
$\operatorname{\mathrm{Cl}} (W)\cdot \overline {f} \to \operatorname{\mathrm{Cl}} (V_4)\cdot f_1 \cdots f_4$
(where f on the left is the product of a basis of
$F_n \cap U^\perp $
) that maps
$\widehat {\operatorname{\mathrm{Gr}} }^+_{\operatorname {\mathrm {iso}}}(W)$
onto
$\widehat {\operatorname{\mathrm{Gr}} }^+_{\operatorname {\mathrm {iso}}}(V_4)$
. Since, by assumption,
$\pi _{n,4}(g \cdot x)=\pi _{e_5} \circ \cdots \circ \pi _{e_n}(g \cdot x)$
lies in the latter isotropic Grassmann cone,
$\pi _{v_{n-4}} \circ \cdots \circ \pi _{v_1}(x)$
lies in the former.
Lemma 6.6. Let
$q \geq n \geq n_0$
. Then for all g in some open dense subset of
$\operatorname {\mathrm {Spin}}(V_q)$
, there exist
$g' \in \operatorname {\mathrm {Spin}}(V_n)$
and
$g" \in \operatorname {\mathrm {Spin}}(V_{n_0})$
such that

holds up to a scalar factor.
Proof. The proof is similar to that above; we just give a sketch. Using the Cartan map, which is equivariant for the relevant spin groups, this lemma follows from a similar statement for the corresponding (halfs of) exterior power representations. Specifically, define

Then the desired identity is

(up to a scalar), where

and the
$c_{e_i}$
and
$m_{f_j}$
are as defined in §3.2 and §3.3, respectively. Furthermore, since the exterior powers are representations of the special orthogonal groups, we may take
$g,g',g"$
to be in
$\operatorname {\mathrm {SO}}(V_q), \operatorname {\mathrm {SO}}(V_n), \operatorname {\mathrm {SO}}(V_{n_0})$
, respectively.
We investigate the effect of the map on the left on (a pure tensor in
$\bigwedge \nolimits ^n V_n$
corresponding to) a maximal (i.e., n-dimensional) isotropic subspace W of
$V_n$
. First, W is extended to
$W':=W \oplus F$
, then g is applied to
$W'$
, and the final contraction map sends
$gW'$
to the image in
$V_{q}/E$
of
$(gW') \cap E^{\perp }$
.
Instead of intersecting
$gW'$
with
$E^\perp $
, we may intersect
$W'=W \oplus F$
with
$(E")^\perp $
, where
$E":=g^{-1} E$
, followed by the isometry
$\overline {g}:(E")^\perp / E" \to E^\perp /E$
induced by g. Accordingly, one can verify that the map on the left-hand side of (6.4) becomes (a scalar multiple of)

where
$c_{E"}:\bigwedge \nolimits ^q V_q \to \bigwedge \nolimits ^{n_0}((E")^\perp / E")$
is the composition of contractions with a basis of
$E"$
, and where we write
$\overline {g}$
also for the map that
$\overline {g}$
induces from
$\bigwedge \nolimits ^{n_0} ((E")^\perp / E")$
to
$\bigwedge \nolimits ^{n_0} (E^\perp /E)$
.
Now consider the space
$E" \cap (V_n \oplus F) \subseteq V_q$
. For g in an open dense subset of
$\operatorname {\mathrm {SO}}(V_q)$
, this intersection has the expected dimension
$(q-n_0) + (2n+q-n)-2q=n-n_0$
, and for g in an open dense subset of
$\operatorname {\mathrm {SO}}(V_q)$
, we also have
$(E")^\perp \cap F=\{0\}$
(because
$(E")^\perp $
has codimension
$q-n_0$
, which is at least the dimension
$q-n$
of F). We restrict ourselves to such g. Then in particular,
$E" \cap F=\{0\}$
, and therefore, the projection
$\widetilde {E} \subseteq V_n$
of
$E" \cap (V_n \oplus F)$
along F has dimension
$n-n_0$
, as well. Note that
$\widetilde {E}$
is isotropic because
$E"$
is and because F is the radical of the bilinear form on
$V_n \oplus F$
.
Furthermore, the projection
$V_n \oplus F \to V_n$
restricts to a linear isomorphism
$(V_n \oplus F) \cap (E")^\perp \to \widetilde {E}^\perp $
, where the latter is the orthogonal complement of
$\widetilde {E}$
inside
$V_n$
. This linear isomorphism induces an isometry

between spaces of dimension
$2n_0$
equipped with a nondegenerate bilinear forms. However, the inclusion
$V_n \oplus F \to V_q$
also induces an isometry

Now a computation shows that, up to a scalar, we have

where
$c_{\widetilde {E}}:\bigwedge \nolimits ^n V_n \to \bigwedge \nolimits ^{n_0}(\widetilde {E}^\perp /\widetilde {E})$
is a composition of contractions with a basis of
$\widetilde {E}$
. Now choose
$g' \in \operatorname {\mathrm {SO}}(V_n)$
such that
$g' \widetilde {E}=E'$
, so that we have

where
$\overline {g'}$
is the isometry
$\widetilde {E}^\perp /\widetilde {E} \to (E')^\perp / E'$
induced by
$g'$
. We then conclude that

and hence, we are done if we set

Competing interest
The authors have no competing interests to declare.
Financial support
CC was supported by Research foundation – Flanders (FWO) – Grant Number 12AZ524N and Swiss National Science Foundation (SNSF) fellowship 217058. JD was partially funded by a Vici grant from the Netherlands Organisation for Scientific Research and Swiss National Science Foundation (SNSF) project grant 200021-227864, and NT was funded by SNSF project grant 200021-191981. RE was supported by NWO Veni grant 016.Veni.192.113. TS is supported by Research foundation – Flanders (FWO) – Grant Number 1219723N.