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

Geometric Complexity Theory - Lie Algebraic Methods for Projective Limits of Stable Points

Bharat Adsul Computer Science Department, Indian Institute of Technology, Mumbai, India. [email protected]    Milind Sohoni Computer Science Department, Indian Institute of Technology, Mumbai, and Indian Institute of Technology, Goa, India. [email protected]    K. V. Subrahmanyam Chennai Mathematical Institute, Chennai, India. [email protected]
(31st December, 2021)

Abstract

Let GG be a connected reductive algebraic group over {\mathbb{C}}, with Lie algebra 𝒢{\cal G}, acting rationally on a complex vector space VV, and the corresponding projective space V\mathbb{P}V. Let xVx\in V and let HGH\subseteq G be its stabilizer and 𝒢{\cal H}\subseteq{\cal G}, its Lie algebra. Our primary objective is to understand the points [y][y], and their stabilizers, which occur in the vicinity of [x][x] in V\mathbb{P}V.

Towards this, we construct an explicit Lie algebra action of 𝒢{\cal G} on a suitably parametrized neighbourhood of xx. As a consequence, we show that the Lie algebras of the stabilizers of points in the neighbourhood of xx are parameterized by subspaces of {\cal H}. When {\cal H} is reductive, our results imply that these are in fact, Lie subalgebras of {\cal H}. If the orbit of xx were closed this would also follow from a celebrated theorem of Luna [Lun73].

To construct our Lie algebra action we proceed as follows. We identify the tangent space to the orbit OxO_{x} of xx with a complement of {\cal H} in 𝒢{\cal G}. Let NN be any linear subspace of VV that is complementary to TOxTO_{x}, the tangent space to the orbit OxO_{x}. We construct an explicit map θ:V×NV\theta:V\times N\rightarrow V which captures the action of the nilpotent part of {\cal H} on VV. We show that the map θ\theta is intimately connected with the local curvature form of the orbit at that point. We call this data, of a neighbourhood of xx with the Lie algebra morphism from 𝒢{\cal G} to vector fields in this neighbourhood, a local model at xx.

The action of GG on VV extends to an action on (V){\mathbb{P}}(V). We illustrate the utility of the local model in understanding when [x]V[x]\in{\mathbb{P}}V is in the projective orbit closure of [y]V[y]\in{\mathbb{P}}V via two applications.

The first application is when V=Symd(X)V=Sym^{d}(X), the space of forms in the variables X:={x1,,xk}X:=\{x_{1},\ldots,x_{k}\}. We consider a form ff (as the “yy” above) and elements in its orbit given by f(t)=A(t)ff(t)=A(t)\cdot f, where A(t)A(t) is an invertible family in GL(X)GL(X), with tt\in{\mathbb{C}}. We express f(t)f(t) as f(t)=tag+tbfb+ higher termsf(t)=t^{a}g+t^{b}f_{b}+\mbox{ higher terms}, with a<ba<b. We call gSymd(X)g\in Sym^{d}(X), the leading term as the limit point (the ”xx” above) and fbf_{b} as the direction of approach. Let 𝒦{\cal K} be the Lie algebra of the stabilizer of ff and {\cal H} the Lie algebra of the stabilizer of gg. The local analysis gives us a flattening 𝒦0{\cal K}_{0} of 𝒦{\cal K} as a subalgebra of {\cal H}, thereby connecting the two stabilizers. There is a natural action of {\cal H} on N¯=V/TOg\overline{N}=V/TO_{g}. We show that 𝒦0{\cal K}_{0} also stabilizes fb¯N¯\overline{f_{b}}\in\overline{N}. We show that there is an ϵ\epsilon-extension Lie algebra 𝒦(ϵ){\cal K}(\epsilon) of 𝒦0{\cal K}_{0} whose structure constants are incrementally closer to 𝒦{\cal K}. The extendability of 𝒦0{\cal K}_{0} to 𝒦(ϵ){\cal K}(\epsilon) translates into certain Lie algebra cohomology conditions.

We then specialize to the projective orbit closures O(f)¯\overline{O(f)} of forms ff whose SL(X)SL(X)-orbits are affine, and where the form gg is such that O(g)¯\overline{O(g)} is of co-dimension 11 in O(f)¯\overline{O(f)}. When gg is obtained as the limit of a 1-PS λ(t)\lambda(t), we examine the tangent of exit, i.e., the direction along which the path λ(t)f\lambda(t)\cdot f leaves ff. We show that (i) either {\cal H} has a very simple structure, or (ii) conjugates of the elements of 𝒦{\cal K}, the stabilizer of ff also stabilize gg and the tangent of exit. This is the triple stabilizer condition.

The second application looks at the space Matn()Mat_{n}(\mathbb{C}) of n×nn\times n-matrices under conjugation by GLnGL_{n}. We show that for a general diagonalizable matrix XX, the signature of nilpotent matrices in the projective closure of its orbit are determined by the multiplicity data of the spectrum of XX.

Finally, we formulate the path problem of finding paths with specific properties from yy to its limit points xx. This connects the algebraic study of the neighborhood of yy, with the local differential geometry of the orbit thereby allowing an optimization formulation.

The questions we study are motivated from Geometric complexity theory, an approach to solve the fundamental lower bound problems in complexity theory. This theory was proposed by the second author and Ketan Mulmuley in [MS01].

1 Introduction

Let GG be a connected reductive algebraic group over {\mathbb{C}} and let ρ:GEnd(V)\rho:G\rightarrow End(V) be a rational representation of GG on a complex vector space VV. We say that GG acts on VV via the representation ρ\rho. Let x,yVx,y\in V. Denote by O(y)O(y) or GyG\cdot y the orbit O(y)={ρ(g)y|gG}O(y)=\{\rho(g)\cdot y|g\in G\} of yy in VV under the given representation. In general the orbit is not Zariski closed, it is only constructible. Clearly GG also acts on projective space (V){\mathbb{P}}(V). For a nonzero vVv\in V, let [v][v] denote the line corresponding to vv in (V){\mathbb{P}}(V).

A fundamental problem of invariant theory is that of obtaining a good description of the GG orbit closures of points yVy\in V and deciding whether a point xx belongs to the orbit closure of yy. These problems go back to Hilbert and are of fundamental importance in the construction of moduli spaces, see [MFK94]. The ubiquitous appearance of this problem in many areas of mathematics is also surveyed in the introduction of [Pop09].

We recall some definitions from Geometric Invariant Theory which we will use throughout the paper. Standard references for this are [MFK94] and [Dol03]. Unless stated otherwise, we work over \mathbb{C}, the field of complex numbers.

We say a point vVv\in V is unstable (or equivalently, is in the nullcone) for the action of GG if Gv¯\overline{Gv} contains 0. A point which is not unstable is said to be semistable. The semistable locus is open in VV. A point is said to be polystable if its orbit is closed in the open set of semistable points. If, furthermore, the stabilizer in GG of a polystable point has dimension 0, we say it is stable. We use the same adjectives, viz., unstable, polystable and stable for the projective point [v][v], based on the status of a representative point v[v]v\in[v].

The central motivation for our investigation arises from computational complexity, which is the study of algorithms to compute functions, typically polynomials and forms. In Appendix  A we give a brief introduction to algebraic complexity theory and the fundamental lower bounds problems therein. We also discuss the geometric complexity theory (GCT) approach to these problems, as proposed by Mulmuley and Sohoni  [MS01],[MS08]. In this formulation, lower bound problems from algebraic complexity translate to determining the membership of special points xx within the projective orbit closure of yy, where yy and xx have distinctive stabilizers. In this setting one of the central problems is to determine the orbit closure in (Symn(n2)){\mathbb{P}}(Sym^{n}(\mathbb{C}^{n^{2}})^{*}) of the determinant, as a polynomial of degree nn in n2n^{2} variables and to determine the stabilizers which arise in this orbit closure. This problem, in turn, is closely connected to determining strata in orbit closures, and their classification.

Determining projective closures is intimately connected to many classical problems. For example, when yy is an unstable point in VV, determining the projective limits of [y][y] requires an analysis of the stratification of the nullcone as a projective variety, [Hes79]. This subsumes, e.g., Hilbert’s work on classifying null forms, and the combinatorial structure of the variety of nilpotent matrices.

Another situation where projective closures arise is in the analysis of leading terms. Consider, for example, a finite dimensional SL(n)SL(n)-module VV, and let λ:SL(n)\lambda:\mathbb{C}^{*}\rightarrow SL(n) be a one parameter subgroup (1-ps) of SL(n)SL(n). Let the action of λ\lambda on a point yVy\in V be given by:

λ(t)y=jtjyj.\lambda(t)\cdot y=\sum_{j}t^{j}y_{j}.

Let xx be the least degree term i.e., x=ykx=y_{k} where yk0y_{k}\neq 0, yi=0y_{i}=0 for all i<ki<k. We call xx the leading term of yy under λ\lambda. Then it is easily seen that [x][x] is in the orbit closure of [y][y] in projective space (V){\mathbb{P}}(V). Whence, projective orbit closures contain all such limit points which arise by the action of 1-ps on [y][y]. Thus, while the orbit of yy may be closed in VV, the orbit [y]V[y]\in\mathbb{P}V is almost always not closed, and picks up points from the nullcone which are intimately connected to it.

We give a brief review of what is known (to the best of our knowledge) about the affine GG-orbit closures of points in VV and the projective GG-orbit closure of points in (V){\mathbb{P}}(V).

The set of unstable points in VV is described by the Hilbert-Mumford theory, see [MFK94, Kem78]. Every unstable point is driven to 0 by a one parameter subgroup. In [Kem78], Kempf showed that for every unstable point yy, there is a canonical parabolic subgroup and an optimal one parameter subgroup which drives yy to zero. The corresponding computational question remains far from solved, except in a few special cases, see [BGO+18], [GGOW19], [DM17], [IQS17].

Hesselink [Hes79] gave a description of the the strata of the nullcone in terms of Kempf’s optimal 1-PS. These strata are locally closed, irreducible and GG-invariant. Popov [Pop03] gave a combinatorial algorithm to determine the strata.

When xx is polystable it follows from Matsushima [Mat60] that the stabilizer HH of xx is reductive. In this case Luna’s étale slice theorem gives information about stabilizers in a Zariski neighbourhoods of xx. Luna’s theorem asserts the existence of an HH-stable affine variety NxVN_{x}\subset V passing through xx and a GG-invariant open neighbourhood UU of the orbit GxGx in VV, having the following properties. Let G×HNxG\times^{H}N_{x} be the fibre bundle over G/HG/H with fibre NxN_{x}(see Section 3.2 for definitions). Luna’s theorem states that the map ψ:G×HNxU\psi:G\times^{H}N_{x}\rightarrow U, sending [g,v][g,v] to gvgv is strongly étale. It follows from Luna’s theorem that the every point in a Zariski neighbourhood of xx in VV has a stabilizer which is contained in a conjugate of HH. So, if HH does not contain a conjugate of the stabilizer of yy, then xx cannot belong to the affine orbit closure of yy.

Of particular interest in GCT is the study of forms fSymn(W)f\in Sym^{n}(W^{*}) with reductive stabilizers and their orbit closures. In this connection, we have Matsushima’s result [Mat60] that when ff has a reductive stabilizer, the SL(W)SL(W)-orbit of fVf\in V is an affine variety. Whence, the closure in projective space O([y])¯\O([y])\overline{O([y])}\backslash O([y]) is either zero or is pure of codimension one. Thus, there are forms g1,,gkg_{1},\ldots,g_{k} whose orbits constitute the closure.

The action of GL(W)GL(W) on Symn(W)Sym^{n}(W^{*}) can be extended to a natural action of GL(W)((t))GL(W)\otimes\mathbb{C}((t)) on Symn(W)Sym^{n}(W^{*}), where ((t))\mathbb{C}((t)) denotes the Laurent series in tt. Hilbert [Hil93] showed that for any [g][g] in the orbit closure of [f][f] there is a σGL(W)((t))\sigma\in GL(W)\otimes\mathbb{C}((t)) such that P:=[f]σP:=[f]\circ\sigma satisfies (P)t=0=[g](P)_{t=0}=[g]. In [LL89] (see also [BLMW11, Section 9.4], [HL16, Section 7.1]), this was reproved. They show the existence of a σ~End(W)[t]\tilde{\sigma}\in End(W)\otimes\mathbb{C}[t] such that fσ~=tKgmod(tK+1)f\circ\tilde{\sigma}=t^{K}g\>mod\>(t^{K+1}). Thus, leading terms under 1-parameter substitutions do lead to all points in the closure, and therefore also the gig_{i} above.

Although the above results tell us how a form gg arises in the closure of a form ff, how is one to determine these forms? In GCT, however one starts with forms ff which are characterized by their stabilizers and one expects the problem to be tractable. Even so very little is known. For example, determining the codimension one components of the boundary of the determinant form is an important open problem.

The orbit closure in projective space of a form ff as above can in principle be determined by resolving the indeterminacy locus of a certain rational map. This is outlined in [BLMW11] and makes clear the connections to Geometric invariant theory. In [Hüt17], the author works this out in the case of the determinant form of a 3 x 3 matrix and shows that there are precisely two codimension one components, thereby proving a conjecture of Landsberg[Lan15].

In Popov [Pop09] the author gives a constructive algorithm to determine if xx is in the orbit closure of yy. He does this in both, the affine as well as the projective setting. This establishes the decidability of the orbit closure membership problem in the sense of computability theory.

1.1 Our results and the organization of the paper.

In the spirit of Luna [Lun73], in this paper we develop a model for a neighbourhood of xx, when xx has a non trivial stabilizer. We aim to study the set of such points yy for which xx may appear as a limit point. Moreover, in the sense of [MS01], we consider points xx and yy which have distinctive stabilizers. While the stabilizer KGK\subseteq G of yy is reductive, the stabilizer HH of xx is typically not reductive.

To describe our approach it will be useful to think of GG as a complex Lie group with Lie algebra 𝒢{\cal G}, see [Lee01]. Let H={g|ρ(g)x=x}H=\{g|\rho(g)\cdot x=x\} be the stabilizer of xx in GG and {\cal H} its Lie algebra. Our main technical construction, Theorem 2.7, is an explicit Lie algebra homomorphism from 𝒢{\cal G} to vector fields at points in a neighbourhood of xx. As a consequence we show that the Lie algebras of the stabilizers of points in this neighbourhood are parameterized by subspaces of {\cal H}. When xx has a reductive stabilizer, our results imply that the Lie algebras of stabilizers of points in the neighbourhood are Lie subalgebras of {\cal H}. Such a conclusion would also follow from Luna’s theorem for polystable xx.

To describe the 𝒢{\cal G}-action, we start with NN, a linear subspace of VV that is complementary to TOxTO_{x}, the tangent space to the orbit OxO_{x}. Let MM be any locally closed submanifold of GG passing through ee that is complementary to HH. Then M×NM\times N is a locally closed submanifold of G×VG\times V containing (e,0)(e,0). Next, the map G×VG×HVG\times V\to G\times^{H}V is 𝒢{\cal G}-equivariant, and the locally closed embedding of M×NG×VM\times N\to G\times V gives us a quotient 𝒢{\cal G} action on M×NM\times N. The map from G×HVG\times^{H}V to VV sending [g,v]g(x+v)[g,v]\rightarrow g(x+v) when restricted to M×NM\times N gives us a 𝒢{\cal G}-equivariant diffeomorphim of a neighbourhood WW of (e,0)M×N(e,0)\in M\times N to a neighbourhood of xx in VV. An important step in this paper is an explicit description of this action of 𝒢{\cal G} on WW in terms of the projections WMW\to M and WNW\to N coming from the product M×NM\times N. An essential component of this construction, described in Section 2.2, is a map θ:V×NV\theta:V\times N\rightarrow V which captures the action of the nilpotent part of {\cal H}. In Remark 2.11 we show that this map is intimately related to the second fundamental form at xx. We call this data, of a neighbourhood of xx with an explicitly defined 𝒢{\cal G}-action, a local model at xx. We develop the local model in Section 2.

We illustrate the utility of the local model in two applications. In Section 3 we apply this to the situation V=Symd(X)V=Sym^{d}(X), the space of degree dd forms in the variables XX. Let ff and gg be two forms with stabilizer (Lie algebras) 𝒦{\cal K} and {\cal H} respectively. We consider a form ff (as the “yy” above) and elements in its orbit, of form f(t)=A(t)ff(t)=A(t)\cdot f, where A(t)A(t) is an invertible family in GL(X)GL(X). We express f(t)f(t) as f(t)=tag+tbfb+ higher termsf(t)=t^{a}g+t^{b}f_{b}+\mbox{ higher terms}, with a<ba<b. We call gSymd(X)g\in Sym^{d}(X), the leading term as the limit point (and “xx” above) and fbf_{b} as the tangent of approach. We construct a basis {𝔨i(t)=𝔥i(t)+𝔰i(t)}i=1dim(𝒦)\{\mathfrak{k}_{i}(t)=\mathfrak{h}_{i}(t)+\mathfrak{s}_{i}(t)\}_{i=1}^{dim({\cal K})} of f(t)f(t) for generic tt, with the coefficients of 𝔥i(t)\mathfrak{h}_{i}(t) belonging to {\cal H}. In Theorem 3.13 we relate the stabilizer {\cal H} of gg with 𝒦{\cal K}, the stabilizer of ff as follows. The above data gives us an action of {\cal H} on N¯=V/TOg\overline{N}=V/TO_{g}. We show that there is a Lie subalegbra 𝒦0{\cal K}_{0}\subset{\cal H} of the same dimension as 𝒦{\cal K} such that 𝒦0{\cal K}_{0} stabilizes fb¯N¯\overline{f_{b}}\in\overline{N}. Furthermore 𝒦0{\cal K}_{0} is the Lie subalgebra spanned by {𝔥i(0)}\{{\mathfrak{h}}_{i}(0)\}.

It turns out that there is an ϵ\epsilon-extension Lie algebra 𝒦(ϵ){\cal K}(\epsilon) of 𝒦0{\cal K}_{0} whose structure constants are incrementally closer to 𝒦{\cal K}. The extendability of 𝒦0{\cal K}_{0} to 𝒦(ϵ){\cal K}(\epsilon) translates into certain cohomology conditions. We discuss this connection to Lie algebra cohomology in Section 6.

In Section 4 we continue our discussion of forms and consider the case when gg is the limit of a 2 block 1-parameter subgroup acting on ff (see Section  for definition). The grading we have on the underlying space by assuming so allows us to prove stronger results. In Proposition 4.10 we show that either there is an element which stabilizes f,gf,g and the tangent of approach (i.e, a triple stabilizer), or the limiting Lie algebra 𝒦0{\cal K}_{0} is nilpotent. In Proposition 4.13 we extend this to the case when gg is the limit of a general 11-ps acting on ff.

In Section 5 we illustrate the results of Section 4.1 when ff has a reductive stabilizer and the orbit of gg is of codimension 11 in the projective closure of ff. In particular we give details when ff is the determinant of a 3×33\times 3 matrix.

In Section 7 we apply the local model to Matn()Mat_{n}(\mathbb{C}), the vector space of n×nn\times n-matrices under conjugation by GLnGL_{n}. We show in Theorem 7.22 that for a general diagonalizable matrix XX, the signature of nilpotent matrices in the projective closure of its orbit are determined by the multiplicity data of the spectrum of XX.

In Section 8 we begin a study of differential geometry at the point yy in the local model. We believe that differential geometric techniques may be useful to further relate properties of the point yy, the tangent vector of exit from yy and the ultimate limit xx. This is illustrated by an example of the cyclic shift matrix in Section 8.3, which requires us to compute the Reimannian curvature tensor. We also make connections to Kempf’s theorem in Proposition  8.6.

Acknowledgements: The second author would like to thank Prof Ketan Mulmuley, Computer Science Department, University of Chicago, for many discussions in the period 1999-2002 during which the θ\theta map was discovered. The third author was supported by a MATRICS grant from the Department of Science and Technology, India and by a grant from the Infosys foundation.

2 The local model

Let GG be a complex algebraic, connected, reductive group, with Lie algebra 𝒢{\cal G}, acting on a complex vector space VV, by a rational representation ρ:GGL(V)\rho:G\rightarrow GL(V). Our interest is in a local model for the action of 𝒢{\cal G}, the Lie algebra of GG, on a neighbourhood of a point xVx\in V, with stabilizer HH, and Lie algbra {\cal H}. Note that ρ:GGL(V)\rho:G\rightarrow GL(V) induces a Lie algebra map ρ:𝒢End(V)\rho:{\cal G}\rightarrow End(V).

An elementary description of a neighborhood of xx is as follows.

  • The orbit O(x)=GxO(x)=G\cdot x is a smooth manifold of dimension dim(G)dim(H)dim(G)-dim(H).

  • Let 𝒮𝒢{\cal S}\subseteq{\cal G} be a complement to {\cal H} within 𝒢{\cal G}, i.e., 𝒢=𝒮{\cal G}={\cal S}\oplus{\cal H}. Then

    𝒮x={ρ(𝔰)x|𝔰𝒮}{\cal S}\cdot x=\{\rho(\mathfrak{s})\cdot x|\mathfrak{s}\in{\cal S}\}

    is the tangent space TxO(x)T_{x}O(x) to the orbit O(x)O(x).

  • Let NTxVN\subseteq T_{x}V be a complement to TxOTxVT_{x}O\subseteq T_{x}V, such that TxV=TxON=𝒮xNT_{x}V=T_{x}O\oplus N={\cal S}\cdot x\oplus N. In any chosen neighborhood UVU\subseteq V of xx, we identify TuVT_{u}V with VV, for all uUu\in U. We may choose a neighborhood UU of xx such that for all uUu\in U, TuV=𝒮uNT_{u}V={\cal S}\cdot u\oplus N.

We would like to describe the action of any 𝔤𝒢\mathfrak{g}\in{\cal G} in this neighborhood UU in terms of the above division of TuV=𝒮uNT_{u}V={\cal S}\cdot u\oplus N. In particular, we will choose u=x+nu=x+n, i.e., the section x+Nx+N, as the domain of interest. The specification of ρ(𝔤)\rho(\mathfrak{g}) on such a section may be suitably extended to sections at other points xO(x)x^{\prime}\in O(x) in the vicinity of xx and therefore to a neighborhood of xx. We show that the action of 𝔤\mathfrak{g} on this section is better understood by (i) using the decomposition 𝒢=𝒮{\cal G}={\cal S}\oplus{\cal H} to express 𝔤𝒢\mathfrak{g}\in{\cal G} as 𝔤=𝔰+𝔥\mathfrak{g}=\mathfrak{s}+\mathfrak{h} with 𝔰𝒮\mathfrak{s}\in{\cal S} and 𝔥\mathfrak{h}\in{\cal H}, and (ii) obtaining the descriptions of 𝔰\mathfrak{s} and 𝔥\mathfrak{h} separately, on the section x+Nx+N. This description is the local model and it relies on the following basic definition and theorem on the formation of quotients.

2.1 The quotient model

Definition 2.1

Let 𝒳{\cal X} be a smooth manifold and let Vec(𝒳)Vec({\cal X}) be the Lie algebra of smooth vector fields on 𝒳{\cal X}. Let 𝒢{\cal G} be a Lie algebra. We say that 𝒳{\cal X} is a 𝒢{\cal G}-manifold via γ\gamma if there is a Lie algebra homomorphism γ:𝒢Vec(𝒳)\gamma:{\cal G}\rightarrow Vec({\cal X}). We denote this by the data (𝒢,γ,𝒳)({\cal G},\gamma,{\cal X}).

Next, let (𝒢,γ,𝒳)({\cal G},\gamma,{\cal X}) and (𝒢,ρ,𝒴)({\cal G},\rho,{\cal Y}) be two 𝒢{\cal G}-manifolds. A map μ:𝒳𝒴\mu:{\cal X}\rightarrow{\cal Y} is called 𝒢{\cal G}-equivariant iff for all x𝒳x\in{\cal X} and 𝔤𝒢\mathfrak{g}\in{\cal G}, we have μx(γ(𝔤)(x))=ρ(𝔤)(μ(x))\mu^{*}_{x}(\gamma(\mathfrak{g})(x))=\rho(\mathfrak{g})(\mu(x)), where μx:Tx𝒳Tμ(x)𝒴\mu^{*}_{x}:T_{x}{\cal X}\rightarrow T_{\mu(x)}{\cal Y} is the tangent map corresponding to the map μ\mu.

We define 𝒟μ(x)Tx𝒳{\cal D}_{\mu}(x)\subseteq T_{x}{\cal X} as those elements ΔxTx𝒳\Delta x\in T_{x}{\cal X} such that μx(Δx)=0\mu^{*}_{x}(\Delta x)=0. Thus 𝒟μ{\cal D}_{\mu} records the kernel of the map μ\mu^{*} at every x𝒳x\in{\cal X}.

Theorem 2.2

Let (𝒢,γ,𝒳)({\cal G},\gamma,{\cal X}) and (𝒢,ρ,𝒴)({\cal G},\rho,{\cal Y}) be two 𝒢{\cal G}-manifolds of dimensions mm and nn, respectively. Let μ:𝒳𝒴\mu:{\cal X}\rightarrow{\cal Y} be 𝒢{\cal G}-equivariant. Suppose that at the point x𝒳x\in{\cal X}, we have μx(Tx𝒳)=Tμ(x)𝒴\mu^{*}_{x}(T_{x}{\cal X})=T_{\mu(x)}{\cal Y}. Moreover, suppose that {\cal M} is a locally closed submanifold of 𝒳{\cal X} of dimension mnm-n such that xx\in{\cal M} and the tangent space Tx(M)T_{x}(M) is transverse to 𝒟μ{\cal D}_{\mu} at xx. Then:

  1. 1.

    There is an open submanifold {\cal M}^{\prime}\subseteq{\cal M} containing xx such that for all mm\in{\cal M}^{\prime}, the tangent space TxT_{x}{\cal M}^{\prime} is complementary to 𝒟μ(m){\cal D}_{\mu}(m).

  2. 2.

    There is a corresponding 𝒢{\cal G}-action γ\gamma_{{\cal M}^{\prime}} on {\cal M}^{\prime} such that the map μ:𝒴\mu:{\cal M}^{\prime}\rightarrow{\cal Y} is a 𝒢{\cal G}-equivariant diffeomorphism in a neighborhood of xx.

  3. 3.

    Moreover, for mm\in{\cal M}^{\prime}, since we have Tm𝒳=Tm𝒟μ(m)T_{m}{\cal X}=T_{m}{\cal M}^{\prime}\oplus{\cal D}_{\mu}(m), let π:Tm𝒳Tm\pi_{\cal M^{\prime}}:T_{m}{\cal X}\rightarrow T_{m}{\cal M}^{\prime} be the corresponding projection. Then we have:

    γ(𝔤)(m)=π(γ(𝔤)(m)) for all 𝔤𝒢 and m\gamma_{{\cal M}^{\prime}}(\mathfrak{g})(m)=\pi_{{\cal M}^{\prime}}(\gamma(\mathfrak{g})(m))\mbox{ for all $\mathfrak{g}\in{\cal G}$ and $m\in{\cal M}^{\prime}$}

For the definition of a 𝒢{\cal G}-manifold see  [AM95],  [Mic08]. The proof of the theorem is straightforward. It allows the construction of a local quotient model {\cal M}^{\prime} within 𝒳{\cal X} of the space 𝒴{\cal Y}.

Definition 2.3

Let GG act on VV via ρ\rho as before. The GG-space G×VG\times V is the collection of tuples (g,v)(g,v) with gGg\in G and vVv\in V. The action γ\gamma of GG on G×VG\times V is given by γ(g)(g,v)=(gg,v)\gamma(g^{\prime})(g,v)=(g^{\prime}g,v). Let xVx\in V be a point with stabilizer HGH\subseteq G, as above. The map μ:G×VV\mu:G\times V\rightarrow V is given by μ(g,v)=ρ(g)(x+v)\mu(g,v)=\rho(g)(x+v) (abbreviated as g(x+v)g\cdot(x+v)).

Lemma 2.4

The following assertions hold:

  1. (i)

    The map μ\mu is GG-equivariant.

  2. (ii)

    Let vVv\in V, then the tangent space T(e,v)(G×V)T_{(e,v)}(G\times V) is (𝔤,Δv)(\mathfrak{g},\Delta v), where 𝔤𝒢\mathfrak{g}\in{\cal G} and ΔvTvVV\Delta v\in T_{v}V\cong V.

  3. (iii)

    The tangent map is given as follows: μ(e,v)(𝔤,Δv)=𝔤(x+v)+Δv\mu^{*}_{(e,v)}(\mathfrak{g},\Delta v)=\mathfrak{g}\cdot(x+v)+\Delta v. Whence, the distribution 𝒟μ(e,v)={(𝔤,Δv)|𝔤(x+v)+Δv=0}{\cal D}_{\mu}(e,v)=\{(\mathfrak{g},\Delta v)|\mathfrak{g}\cdot(x+v)+\Delta v=0\}.

  4. (iv)

    μ(e,0)=𝔤x+Δv\mu^{*}_{(e,0)}=\mathfrak{g}\cdot x+\Delta v. Thus μ\mu^{*} is surjective at (e,0)(e,0).

The proof is straightforward.

Proposition 2.5

Let NN be a vector space complement to TO(x)xTO(x)_{x} within VV. Let 𝒮{\cal S} be a complement to {\cal H} within 𝒢{\cal G}. Let MM be a locally closed submanifold of GG of dimension dim(G)dim(H)dim(G)-dim(H) such that eMe\in M and TeM=𝒮T_{e}M={\cal S}. Then there is an open submanifold MMM^{\prime}\subseteq M containing ee, such that 𝒢{\cal G} acts on M×NG×VM^{\prime}\times N\subseteq G\times V via γ\gamma^{\prime}. This action is derived from the action γ\gamma of 𝒢{\cal G} on G×VG\times V modulo the distribution 𝒟μ{\cal D}_{\mu}. Moreover, for this neighborhood of M×NM^{\prime}\times N containing (e,0)(e,0), we have:

  • (i)

    The map μ:M×NV\mu:M^{\prime}\times N\rightarrow V is a 𝒢{\cal G}-equivariant diffeomorphism from the space (𝒢,γ,M×N)({\cal G},\gamma^{\prime},M^{\prime}\times N) to (𝒢,ρ,V)({\cal G},\rho,V). It maps a neighborhood of [e,0]M×N[e,0]\in M^{\prime}\times N to a neighborhood of xVx\in V.

  • (ii)

    For any element 𝔤=𝔰+𝔥\mathfrak{g}=\mathfrak{s}+\mathfrak{h}, with 𝔰𝒮\mathfrak{s}\in{\cal S} and 𝔥\mathfrak{h}\in{\cal H}, the action of γ(𝔤)\gamma^{\prime}(\mathfrak{g}) at the point (e,n)M×N(e,n)\in M^{\prime}\times N is given by

    γ(𝔤)(e,n)=(𝔰+𝔰,n)T(e,n)(M×N)\gamma^{\prime}(\mathfrak{g})(e,n)=(\mathfrak{s}+\mathfrak{s}^{\prime},n^{\prime})\in T_{(e,n)}(M^{\prime}\times N)

    where 𝔰𝒮\mathfrak{s}^{\prime}\in{\cal S} and nNn^{\prime}\in N are unique and satisfy the equation

    𝔰(x+n)+n=𝔥n.\mathfrak{s}^{\prime}\cdot(x+n)+n^{\prime}=\mathfrak{h}\cdot n.

Proof: Let us verify that the conditions required by Theorem 2.2 hold for the map μ\mu. From lemma 2.4, (i) and (iv), μ\mu is GG-equivariant and μ\mu^{*} is surjective at (e,0)(e,0). By the choice of MM and NN, we see that μ(e,0)(M×N)=𝒮x+N=V\mu^{*}(e,0)(M\times N)={\cal S}\cdot x+N=V. Thus T(e,0)M×NT_{(e,0)}M\times N is transverse to 𝒟μ{\cal D}_{\mu}. Thus Theorem 2.2 applies at the point (e,0)(e,0) and a suitable action γ\gamma^{\prime} on M×NM^{\prime}\times N is available for which μ\mu is an equivariant diffeomorphism. Thus (i) is proved. We also have that M×NM^{\prime}\times N is transverse to the distribution DμD_{\mu}. To prove (ii), we must make an explicit computation of γ\gamma^{\prime} at the points (e,n)M×NG×V(e,n)\in M^{\prime}\times N\subseteq G\times V modulo the distribution DμD_{\mu}. For this, we must evaluate πM:T(e,n)(G×V)T(e,n)(M×N)\pi_{M^{\prime}}:T_{(e,n)}(G\times V)\rightarrow T_{(e,n)}(M\times N). We do this in two steps:

T(e,n)G×VT(e,n)M×VT(e,n)M×NT_{(e,n)}G\times V\rightarrow T_{(e,n)}M\times V\rightarrow T_{(e,n)}M\times N

From Lemma 2.4(ii), for a typical element 𝔤=𝔰+𝔥\mathfrak{g}=\mathfrak{s}+\mathfrak{h}, we have γ(𝔤)(e,n)=(𝔰+𝔥,0)T(e,n)(G×V)\gamma(\mathfrak{g})(e,n)=(\mathfrak{s}+\mathfrak{h},0)\in T_{(e,n)}(G\times V). We observe that for any 𝔥\mathfrak{h}\in{\cal H}, we have

μ(e,n)(𝔥,0)=𝔥(x+n)=𝔥n=μ(e,n)(0,𝔥n)\mu^{*}_{(e,n)}(\mathfrak{h},0)=\mathfrak{h}\cdot(x+n)=\mathfrak{h}\cdot n=\mu^{*}_{(e,n)}(0,\mathfrak{h}\cdot n)

Thus, we have:

γM(𝔰+𝔥)(e,n)=(𝔰+𝔥,0)=(𝔰,𝔥n)T(e,n)M×V (modulo 𝒟μ)\gamma_{M}(\mathfrak{s}+\mathfrak{h})(e,n)=(\mathfrak{s}+\mathfrak{h},0)=(\mathfrak{s},\mathfrak{h}\cdot n)\in T_{(e,n)}M\times V\mbox{ (modulo ${\cal D}_{\mu}$)}

Thus, we have reduced the first component from 𝒢{\cal G} to 𝒮=TeM{\cal S}=T_{e}M.

Using Lemma 2.4(iii) it follows that

μ(e,n)(𝔰,𝔥n)=𝔰(x+n)+𝔥nT(x+n)V\mu^{*}_{(e,n)}(\mathfrak{s},\mathfrak{h}\cdot n)=\mathfrak{s}\cdot(x+n)+\mathfrak{h}\cdot n\in T_{(x+n)}V

Consider the vector 𝔥nT(x+n)V\mathfrak{h}\cdot n\in T_{(x+n)}V. Since μ\mu is a local diffeomorphism between (e,0)M×N(e,0)\in M\times N and xVx\in V, we may assume that μ\mu is regular at x+nx+n. Whence, there must be a unique 𝔰𝒮\mathfrak{s^{\prime}}\in{\cal S} and nNn^{\prime}\in N such that for (𝔰,n)T(e,n)M×N(\mathfrak{s^{\prime}},n^{\prime})\in T_{(e,n)}M\times N, we have μ(e,n)(𝔰,n)=𝔥n\mu^{*}_{(e,n)}(\mathfrak{s}^{\prime},n^{\prime})=\mathfrak{h}\cdot n, or in other words:

𝔰(x+n)+n=𝔥n.\mathfrak{s}^{\prime}\cdot(x+n)+n^{\prime}=\mathfrak{h}\cdot n.

Hence

μ(e,n)(𝔤[e,n])=μ(e,n)(𝔰,𝔥n)=𝔰(x+n)+𝔥n=(𝔰+𝔰)(x+n)+n=μ(e,n)(𝔰+𝔰,n)\mu_{*(e,n)}(\mathfrak{g}\cdot[e,n])=\mu^{*}_{(e,n)}(\mathfrak{s},\mathfrak{h}\cdot n)=\mathfrak{s}\cdot(x+n)+\mathfrak{h}\cdot n=(\mathfrak{s}+\mathfrak{s}^{\prime})\cdot(x+n)+n^{\prime}=\mu^{*}_{(e,n)}(\mathfrak{s}+\mathfrak{s^{\prime}},n^{\prime})

Since (𝔰+𝔰,n)T(M×N)(e,n)(\mathfrak{s}+\mathfrak{s^{\prime}},n^{\prime})\in T(M\times N)_{(e,n)}, it must be the projection of (𝔰,𝔥n)(\mathfrak{s},\mathfrak{h}\cdot n) to T(e,n)(M×N)T_{(e,n)}(M\times N), and thus from Theorem 2.2(3) it describes the action of 𝔤\mathfrak{g} on M×NM\times N at the point (e,n)M×N(e,n)\in M\times N . This proves (ii). \Box

2.2 The θ\theta map.

It is now clear that the computation modulo the distribution DμD_{\mu} at (e,n)(e,n) essentially revolves around the solution to the following equation in VV, where ΔvVT(x+n)V{\Delta v}\in V\cong T_{(x+n)}V is a given and 𝔰𝒮{\mathfrak{s}}\in{\cal S}, 𝔫N{\mathfrak{n}}\in N are the unknowns:

𝔰(x+n)+𝔫=Δv.{\mathfrak{s}}\cdot(x+n)+{\mathfrak{n}}=\Delta v.

Since V=𝒮(x+n)NV={\cal S}\cdot(x+n)\oplus N, we may define projection λ𝒮(n):V𝒮\lambda_{\cal S}(n):V\rightarrow{\cal S} and λN(n):VN\lambda_{N}(n):V\rightarrow N so that the above equation may be written as:

Δv=(λ𝒮(n)(Δv))(x+n)+(λN(n)(Δv))\Delta v=(\lambda_{\cal S}(n)(\Delta v))(x+n)+(\lambda_{N}(n)(\Delta v))

We now solve the above equation, with nn as a parameter. We note that for n=0n=0, the equation is easily solved by our choice of NN and the isomorphism V=TxV=TxO(x)NV=T_{x}V=T_{x}O(x)\oplus N with TxO(x)=𝒮xT_{x}O(x)={\cal S}\cdot x. We may write the projections λ𝒮(0):V𝒮\lambda_{\cal S}(0):V\rightarrow{\cal S} and λN(0):VN\lambda_{N}(0):V\rightarrow N, or simply λ𝒮,λN\lambda_{\cal S},\lambda_{N}, such that

λ𝒮(Δv)x+λN(Δv)=Δv.\lambda_{\cal S}(\Delta v)\cdot x+\lambda_{N}(\Delta v)=\Delta v.

By adding an “error term” λ𝒮(Δv)n\lambda_{\cal S}(\Delta v)\cdot n on both sides, we get

λ𝒮(Δv)(x+n)+λN(Δv)=Δv+λ𝒮(Δv)n.\lambda_{\cal S}(\Delta v)\cdot(x+n)+\lambda_{N}(\Delta v)=\Delta v+\lambda_{\cal S}(\Delta v)\cdot n.

For a fixed nNn\in N, we may now define θ(n):VV\theta(n):V\rightarrow V, as follows: for ΔvV\Delta v\in V, we define θ(n)(Δv)=λ𝒮(Δv)n\theta(n)(\Delta v)=\lambda_{\cal S}(\Delta v)\cdot n. We see that θ(n):VV\theta(n):V\rightarrow V is a linear operator VV which depends linearly on nn. Thus, the above equation may be rewritten as:

λ𝒮(Δv)(x+n)+λN(Δv)=Δv+θ(n)(Δv)\lambda_{\cal S}(\Delta v)\cdot(x+n)+\lambda_{N}(\Delta v)=\Delta v+\theta(n)(\Delta v)

Substituting θi(Δv)\theta^{i}(\Delta v) for Δv\Delta v in the above equaltion gives us the ii-th approximation term as:

λ𝒮(θi(n)(Δv))(x+n)+λN(θi(n)(Δv))=θi(n)(Δv)+θi+1(Δv).\lambda_{\cal S}(\theta^{i}(n)(\Delta v))\cdot(x+n)+\lambda_{N}(\theta^{i}(n)(\Delta v))=\theta^{i}(n)(\Delta v)+\theta^{i+1}(\Delta v).

Since θ(n)\theta(n) depends linearly on nn, we may assume that nn is small enough and hence all the eigen-values of θ(n)\theta(n) may be assumed to be less than 11 in magnitude. Thus the system,

𝔰(x+n)+𝔫=Δv,{\mathfrak{s}}\cdot(x+n)+{\mathfrak{n}}=\Delta v,

has the solution:

[λ𝒮(1θ(n)+θ2(n))](Δv)(x+n)+[λN(1θ(n)+θ2(n))](Δv)=Δv.[\lambda_{\cal S}\circ(1-\theta(n)+\theta^{2}(n)-\cdots)](\Delta v)\cdot(x+n)+[\lambda_{N}\circ(1-\theta(n)+\theta^{2}(n)-\cdots)](\Delta v)=\Delta v.

Thus, we have proved the following basic lemma:

Lemma 2.6

For any element ΔvT(x+n)V\Delta v\in T_{(x+n)}V, there are unique 𝔰𝒮\mathfrak{s}\in{\cal S} and nNn^{\prime}\in N such that:

𝔰(x+n)+n=Δv\mathfrak{s}\cdot(x+n)+n^{\prime}=\Delta v

Moreover 𝔰=λ𝒮((1+θ(n))1(Δv))\mathfrak{s}=\lambda_{\cal S}\circ((1+\theta(n))^{-1}(\Delta v)) and n=λN(1+θ(n))1(Δv)n^{\prime}=\lambda_{N}\circ(1+\theta(n))^{-1}(\Delta v). Thus, if (n):VV{\cal I}(n):V\rightarrow V is the map (1+θ(n))1(1+\theta(n))^{-1}, then λ𝒮(n)=λ𝒮(n)\lambda_{\cal S}(n)=\lambda_{\cal S}\circ{\cal I}(n) and λN(n)=λN(n)\lambda_{N}(n)=\lambda_{N}\circ{\cal I}(n).

We will use the above calculations and Proposition 2.5 to prove our main theorem.

Recall that HH stabilizes the point xx. We may choose a Levi decomposition H=RQH=RQ, where QQ is the unipotent radical of HH and RR is a reductive complement. Let {\cal R} denote the Lie algebra of RR and 𝒬{\cal Q} denote the Lie algebra of QQ.

We may choose NN in the above proposition to be an RR-invariant complement to the space OxO_{x} inside VV. We may also choose 𝒮{\cal S} to be an RR-invariant complement to {\cal H} with 𝒢{\cal G}. Thus 𝒮{\cal S} now acquires an RR-module structure.

Theorem 2.7

Let ρ\rho be a representation of GG on a vector space VV. Let xVx\in V be a point with stabilizer HH. Let H=RQH=RQ be a Levi decomposition and ,𝒬{\cal H},{\cal Q} and {\cal R} be the Lie algebras of H,QH,Q and RR, respectively. Then there is a submanifold M×NG×HVM\times N\subset G\times_{H}V in the neighbourhood of (e,0)(e,0) with an action of 𝒢{\cal G}. The map μ:M×NV\mu:M\times N\rightarrow V given by μ(g,n)=g(x+n)\mu(g,n)=g\cdot(x+n) is a 𝒢{\cal G}-equivariant diffeomorphism which maps a neighborhood of (e,0)M×N(e,0)\in M\times N to a neighborhood of xVx\in V. We then have:

  1. 1.

    Let 𝔤\mathfrak{g} be a typical element of 𝒢{\cal G}. Let 𝔤=𝔰+𝔯+𝔮\mathfrak{g}=\mathfrak{s}+\mathfrak{r}+\mathfrak{q}, with 𝔰𝒮,𝔯\mathfrak{s}\in{\cal S},\mathfrak{r}\in{\cal R} and 𝔮𝒬\mathfrak{q}\in{\cal Q}. Then the action γ\gamma^{\prime} of 𝔤\mathfrak{g} on M×NM\times N is given as:

    𝔤(e,n)=(𝔰+λ𝒮(1+θ(n))1(𝔮n),𝔯n+λN(1+θ(n))1(𝔮n))\mathfrak{g}\circ(e,n)=(\mathfrak{s}+\lambda_{\cal S}\circ(1+\theta(n))^{-1}(\mathfrak{q}\cdot n),\>\mathfrak{r}\cdot n+\lambda_{N}\circ(1+\theta(n))^{-1}(\mathfrak{q}\cdot n))
  2. 2.

    If 𝒢n𝒢{\cal G}_{n}\subseteq{\cal G} is the stabilizer of x+nVx+n\in V, then there is a subspace n{\cal H}_{n}\subseteq{\cal H} of the same dimension as 𝒢n{\cal G}_{n} such that for every 𝔤𝒢n\mathfrak{g}\in{\cal G}_{n}, there is a unique 𝔥n\mathfrak{h}\in{\cal H}_{n} with 𝔥=𝔮+𝔯\mathfrak{h}=\mathfrak{q}+\mathfrak{r} such that:

    𝔯n+λN(1+θ(n))1(𝔮n)=0\mathfrak{r}\cdot n+\lambda_{N}\circ(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=0

    Moreover, 𝔤=𝔥+𝔰\mathfrak{g}=\mathfrak{h}+\mathfrak{s}, where 𝔰\mathfrak{s} is as given below:

    𝔰=λ𝒮(1+θ(n))1(𝔮n)\mathfrak{s}=-\lambda_{\cal S}\circ(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)

    The element 𝔤\mathfrak{g} is called the 𝒮{\cal S}-completion of 𝔥\mathfrak{h} at the point x+nx+n.

Proof: Let us express 𝔤=𝔰+𝔥\mathfrak{g}=\mathfrak{s}+\mathfrak{h} and apply Prop. 2.5, part (ii). We then have that γ(𝔤)(e,n)=(𝔰+𝔰,n)\gamma^{\prime}(\mathfrak{g})(e,n)=(\mathfrak{s}+\mathfrak{s^{\prime}},n^{\prime}), where 𝔰\mathfrak{s}^{\prime} and nn^{\prime} solve the equation:

𝔥n=𝔰(x+n)+n\mathfrak{h}\cdot n=\mathfrak{s}^{\prime}(x+n)+n^{\prime}

By lemma 2.6, these are precisely λ𝒮(1+θ(n)1(𝔥n)\lambda_{\cal S}(1+\theta(n)^{-1}(\mathfrak{h}\cdot n) and n=λN(1+θ(n))1(𝔥n)n^{\prime}=\lambda_{N}(1+\theta(n))^{-1}(\mathfrak{h}\cdot n). Now, breaking 𝔥\mathfrak{h} as 𝔥=𝔮+𝔯\mathfrak{h}=\mathfrak{q}+\mathfrak{r}, and recalling that 𝔯NN\mathfrak{r}\cdot N\subseteq N, we see λN(𝔯n)=𝔯n\lambda_{N}(\mathfrak{r}\cdot n)=\mathfrak{r}\cdot n, and λ𝒮(𝔯n)=0\lambda_{\cal S}(\mathfrak{r}\cdot n)=0. Thus θ(n)(𝔯n)=0\theta(n)(\mathfrak{r}\cdot n)=0 and thus λN(1+θ(n))1(𝔯n+𝔮n)=λN(𝔯n)+λN(1+θ(n))1(𝔮n)=𝔯n+λN(1+θ(n))1(𝔮n)\lambda_{N}(1+\theta(n))^{-1}(\mathfrak{r}\cdot n+\mathfrak{q}\cdot n)=\lambda_{N}(\mathfrak{r}\cdot n)+\lambda_{N}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=\mathfrak{r}\cdot n+\lambda_{N}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n). Similarly, λ𝒮(1+θ(n))1(𝔯n+𝔮n)=λ𝒮(𝔯n)+λ𝒮(1+θ(n))1(𝔮n)=λ𝒮(1+θ(n))1(𝔮n)\lambda_{\cal S}(1+\theta(n))^{-1}(\mathfrak{r}\cdot n+\mathfrak{q}\cdot n)=\lambda_{\cal S}(\mathfrak{r}\cdot n)+\lambda_{\cal S}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=\lambda_{\cal S}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n).

This proves (1).

The hypothesis that 𝔤𝒢n\mathfrak{g}\in{\cal G}_{n} is equivalent to the claim that γ(𝔤)(e,n)=0\gamma^{\prime}(\mathfrak{g})(e,n)=0. This requires both the components of the action of 𝔤(e,n)\mathfrak{g}\cdot(e,n) to vanish. This implies that 𝔯n+λN(1+θ(n))1(𝔮n)=0\mathfrak{r}\cdot n+\lambda_{N}\circ(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=0. This is the first assertion.

Now, given that 𝔥=𝔮+𝔯\mathfrak{h}=\mathfrak{q}+\mathfrak{r} satisfies the above, 𝔰\mathfrak{s} is uniquely determined by the requirement that the first component vanishes, i.e., 𝔰+λ𝒮(1+θ(n))1(𝔮n)\mathfrak{s}+\lambda_{\cal S}\circ(1+\theta(n))^{-1}(\mathfrak{q}\cdot n) must vanish. This determines 𝔰\mathfrak{s} as required.

Finally, if there were 𝔤=𝔥+𝔰\mathfrak{g}^{\prime}=\mathfrak{h}+\mathfrak{s} and 𝔤=𝔥+𝔰\mathfrak{g}^{\prime}=\mathfrak{h}+\mathfrak{s}^{\prime} both elements of 𝒢n{\cal G}_{n}, then we would have 𝔰𝔰𝒢n\mathfrak{s}-\mathfrak{s^{\prime}}\in{\cal G}_{n}. Whence, (𝔰𝔰)(e,n)=(𝔰𝔰,0)0(\mathfrak{s}-\mathfrak{s}^{\prime})\cdot(e,n)=(\mathfrak{s}-\mathfrak{s}^{\prime},0)\neq 0 which violates the requirement that the first component vanishes. \Box

Remark 2.8

Consider the case when the stabilizer HH is reductive, whence R=HR=H and Q=0Q=0. We may select an NN which is HH-invariant as above. Since 𝒬=0{\cal Q}=0, we see that 𝔮n=0\mathfrak{q}\cdot n=0, and thus, the 𝒮{\cal S}-completion of any element 𝔥H\mathfrak{h}\in H is 𝔥\mathfrak{h} itself. Thus the stabilizer 𝒢n{\cal G}_{n} of x+nx+n must be a subalgebra of {\cal H}. This is the Lie algebraic version of Luna’s slice theorem.

We now prove certain equivariant properties of maps closely related to θ\theta. We continue with the Levi decomposition of H=RQH=RQ and the choice of NN and 𝒮{\cal S}.

The Lie algebra action defined as a map ρ:𝒢End(V)\rho:{\cal G}\rightarrow End(V) may well be defined as ρ^:𝒢VV\hat{\rho}:{\cal G}\otimes V\rightarrow V. Note that 𝒢{\cal G} is an RR-module under the adjoint action, and thus so is 𝒢V{\cal G}\otimes V. Then we have the following lemma:

Lemma 2.9

The map ρ^\hat{\rho} is RR-equivariant. Next, let TxOT_{x}O be the tangent space of the orbit of xx under the action of GG and 𝒮{\cal S} be an RR-invariant complement to 𝒢{\cal H}\subseteq{\cal G}. Then the map α:𝒮TxO\alpha:{\cal S}\rightarrow T_{x}O given by α(𝔰)=ρ(𝔰)x\alpha(\mathfrak{s})=\rho(\mathfrak{s})\cdot x is an isomorphism of RR-modules.

Proof. For the first assertion, note that for rR,vVr\in R,v\in V and 𝔤𝒢\mathfrak{g}\in{\cal G}, we have:

rρ^(𝔤v)=rρ(𝔤)r1rv=r(ρ(𝔤)v)r\cdot\hat{\rho}(\mathfrak{g}\otimes v)=r\cdot\rho(\mathfrak{g})\cdot r^{-1}\otimes r\cdot v=r\cdot(\rho(\mathfrak{g})\cdot v)

Thus ρ^\hat{\rho} is indeed RR-equivariant. Next, the map α:𝒮TxO\alpha:{\cal S}\rightarrow T_{x}O is merely the restriction to ρ^\hat{\rho} to the tensor product of two RR-submodules, viz., 𝒮𝒢{\cal S}\subset{\cal G} and xV\mathbb{C}\cdot x\subseteq V. This proves the second assertion. \Box

Recall that θ(n):VV\theta(n):V\rightarrow V is defined as the composite:

θ:Vλ𝒮𝒮nV\theta:V\stackrel{{\scriptstyle\lambda_{\cal S}}}{{\rightarrow}}{\cal S}\stackrel{{\scriptstyle\cdot n}}{{\rightarrow}}V

We define two associated maps: Θ:VNV\Theta:V\otimes N\rightarrow V and Φ:𝒮×N𝒮\Phi:{\cal S}\times N\rightarrow{\cal S} as:

Θ(vn)=θ(n)(v)=λ𝒮(v)n\Theta(v\otimes n)=\theta(n)(v)=\lambda_{\cal S}(v)\cdot n

and:

Φ(𝔰n)=λ𝒮(𝔰n)\Phi(\mathfrak{s}\otimes n)=\lambda_{\cal S}(\mathfrak{s}\cdot n)

Then, we have the following:

Proposition 2.10

The maps Θ\Theta and Φ\Phi are RR-equivariant.

Proof: Θ\Theta is merely the composite:

VNλ𝒮𝚒𝚍𝒮Nρ^VV\otimes N\stackrel{{\scriptstyle\lambda_{\cal S}\otimes{\tt id}}}{{\longrightarrow}}{\cal S}\otimes N\stackrel{{\scriptstyle\hat{\rho}}}{{\longrightarrow}}V

As for Φ\Phi, it is:

𝒮Nρ^Vλ𝒮𝒮{\cal S}\otimes N\stackrel{{\scriptstyle\hat{\rho}}}{{\longrightarrow}}V\stackrel{{\scriptstyle\lambda_{\cal S}}}{{\longrightarrow}}{\cal S}

This proves the proposition. \Box

Remark 2.11

The map Φ\Phi above is the algebraic analogue of the Second Fundamental Tensor at the point xx for the orbit O(x)O(x) regarded as a submanifold of VV. Indeed, suppose that GG is the orthogonal group and VV is equipped with a GG-invariant inner product. Let NN be chosen orthogonal to TxOT_{x}O, the tangent space of the orbit O(x)O(x) at xx. Let v1,,vKv_{1},\ldots,v_{K} be an orthonormal basis for NN. Let 𝔰1,,𝔰L\mathfrak{s}_{1},\ldots,\mathfrak{s}_{L} be chosen such that the elements {vK+i=(𝔰ix)|i=1,,L}\{v_{K+i}=(\mathfrak{s}_{i}\cdot x)|i=1,\ldots,L\} is an orthonormal basis for TxOT_{x}O. The data for Φ\Phi is the tuple {(αijr)|i,j=1,,L,r=1,,K}\{(\alpha_{ij}^{r})|i,j=1,\ldots,L,r=1,\ldots,K\}, where we have:

λ𝒮(𝔰ivr)=jαijr(𝔰jx)\lambda_{\cal S}(\mathfrak{s}_{i}\cdot v_{r})=\sum_{j}\alpha_{ij}^{r}(\mathfrak{s}_{j}\cdot x)

Now, one must recall that ρ(𝔰i)\rho(\mathfrak{s}_{i}) may be given as a linear vector field as follows:

ρ(𝔰i)=j=1K+Lk=1K+LSi(j,k)vkvj\rho(\mathfrak{s}_{i})=\sum_{j=1}^{K+L}\sum_{k=1}^{K+L}S_{i}(j,k)v_{k}\frac{\partial}{\partial v_{j}}

where Si(K+L)×(K+L)S_{i}\in\mathbb{C}^{(K+L)\times(K+L)}. We abbreviate this as ρ(𝔰i)=Siv¯\rho(\mathfrak{s}_{i})=\nabla S_{i}\overline{v}, where \nabla is the row vector [(/v1),,(/vK+L)][(\partial/\partial v_{1}),\ldots,(\partial/\partial v_{K+L})] and v¯\overline{v} is the column vector [v1,,vK+L]T[v_{1},\ldots,v_{K+L}]^{T}.

Note that ρ\rho preserves inner products implies that SiT=SiS_{i}^{T}=-S_{i}. The data αijr\alpha_{ij}^{r} may then be obtained as:

αijr=(Sjx)T(Sivr)=xTSjSivr\alpha_{ij}^{r}=(S_{j}x)^{T}(S_{i}v_{r})=-x^{T}S_{j}S_{i}v_{r}

We also have the Second Fundamental Form of the orbit O(x)O(x) as a map Π:TxO×TxON\Pi:T_{x}O\times T_{x}O\rightarrow N defined as follows. For vi,vjTxOv_{i},v_{j}\in T_{x}O, let Xi,XjX_{i},X_{j} be vector fields on O(x)O(x) such Xi(x)=viX_{i}(x)=v_{i} and Xj(x)=vjX_{j}(x)=v_{j}. Then Π(vi,vj)=λN(DXi(Xj))=r=1Kβijrvr\Pi(v_{i},v_{j})=\lambda_{N}(D_{X_{i}}(X_{j}))=\sum_{r=1}^{K}\beta_{ij}^{r}v_{r}. Now, we have ρ(𝔰i)\rho(\mathfrak{s}_{i}) available to us as the XiX_{i} above. Then, it is easily checked that DXi(Xj)D_{X_{i}}(X_{j}) is given SjSiv¯\nabla S_{j}S_{i}\overline{v}. Whence the evaluation of Π(vi,vj)\Pi(v_{i},v_{j}) at the point xx and in the direction vrv_{r} is given by βijr=vrTSjSix\beta_{ij}^{r}=v_{r}^{T}S_{j}S_{i}x. However, this is equal to vrTSiSjx+vrT[Sj,Si]xv_{r}^{T}S_{i}S_{j}x+v_{r}^{T}[S_{j},S_{i}]x. Since there is a 𝔤=[𝔰i,𝔰j]𝒢\mathfrak{g}=[\mathfrak{s}_{i},\mathfrak{s}_{j}]\in{\cal G} such that ρ(𝔤)=[Sj,Si]v¯\rho(\mathfrak{g})=\nabla[S_{j},S_{i}]\overline{v}, this evaluated at xx must be in the tangent space TxOT_{x}O, and thus vrT[Sj,Si]x=0v_{r}^{T}[S_{j},S_{i}]x=0. Whence

βijr=vrTSiSjx=xTSjTSiTvr=xTSjSivr=αijr\beta_{ij}^{r}=v_{r}^{T}S_{i}S_{j}x=x^{T}S_{j}^{T}S_{i}^{T}v_{r}=x^{T}S_{j}S_{i}v_{r}=-\alpha_{ij}^{r}

This proves the claim. \Box

We explore further uses of the Reimannian curvature tensor to study projective orbit closures in Section 8.2.

2.3 Applying the local model

We close this section with a general application of the local model to projective limits of a particular type, and a simple concrete example.

Proposition 2.12

Let VV be a GG-module, where GGL(X)G\subseteq GL(X), for a suitable XX and let p,qVp,q\in V. Let II be the identity matrix and tItI act on pp as tIp=tdptI\cdot p=t^{d}p. Let A(t)GA(t)\subseteq G be an algebraic family of elements in GG such that:

A(t)q=tkp+ higher degree terms A(t)\cdot q=t^{k}\cdot p+\mbox{ higher degree terms }

Suppose that the stabilizer of pp is HH. Then the local model at pp contains a scalar multiple of a GG-conjugate of qq which is of the form p+np+n, i.e., appears on the normal slice NN at pp.

Proof. Consider the family A(t)=tk/dA(t)A^{\prime}(t)=t^{-k/d}\cdot A(t), i.e., whose elements are scalar multiples of A(t)A(t) and thus are elements of GG. Applying this family to qq, we see that:

A(t)q=p+ positive powers of tA^{\prime}(t)\cdot q=p+\mbox{ positive powers of $t$}

Thus, limt0A(t)q=p\lim_{t\rightarrow 0}A^{\prime}(t)q=p, and for every ϵ>0\epsilon>0, there is a q0=A(t0)qq_{0}=A^{\prime}(t_{0})\cdot q, i.e., a scalar multiple of a GG-conjugate of qq such q0p<ϵ\|q_{0}-p\|<\epsilon. Thus, in the local model μ:M×NV\mu:M\times N\rightarrow V, we have a (g,n)(g,n) such that g(p+n)=q0g\cdot(p+n)=q_{0}. Applying g1g^{-1} on both sides, we see that g1q0=p+ng^{-1}q_{0}=p+n. This proves the claim. \Box

Let us consider a small example to illustrate the computational ease of the local model.

Example 2.13

Consider sl(2)sl(2) given as the matrix below acting of X\mathbb{C}\cdot X, where X={x,y}X=\{x,y\}. Let V=Sym2(X)V=Sym^{2}(X) with the ordered basis B={x2,xy,y2}B=\{x^{2},xy,y^{2}\}. The action of the generic group element is also shown below:

sl(2)={ga,b,c=[abca] with a,b,c}sl(2)=\left\{g_{a,b,c}=\left[\begin{array}[]{cr}a&b\\ c&-a\end{array}\right]\mbox{ with }a,b,c\in\mathbb{C}\right\}

On Sym2(X)Sym^{2}(X), it may be useful to interpret this as the operation:

a(xxyy)+byx+cxya(x\frac{\partial}{\partial x}-y\frac{\partial}{\partial y})+by\frac{\partial}{\partial x}+cx\frac{\partial}{\partial y}

Whence:

ga,b,c[x2xyy2]=[2a2b0c0b02c2a][x2xyy2]g_{a,b,c}\cdot\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]=\left[\begin{array}[]{ccc}2a&2b&0\\ c&0&b\\ 0&2c&-2a\end{array}\right]\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]

Let us consider three forms p0=0,p1=x2p_{0}=0,p_{1}=x^{2} and p2=x2+y2p_{2}=x^{2}+y^{2} and apply the local model to each.

  1. 1.

    For p0=0p_{0}=0, we see that 0{\cal H}_{0}, the stabilizer of p0p_{0} is the complete group sl(2)sl(2). Thus 𝒮=0{\cal S}=0, the orbit is 0-dimensional, and the maps θ\theta is 0. The only choice for NN is VV, and for any nn, the stabilizer 𝒢n0=sl(2){\cal G}_{n}\subseteq{\cal H}_{0}=sl(2).

  2. 2.

    Consider next p2=x2+y2p_{2}=x^{2}+y^{2}. We see that the stabilizer 2{\cal H}_{2} is the group Osl(2)O\subset sl(2), given below. A complementary subspace is also shown.

    O={[0bb0]|b}𝒮={[acca]|a,c}O=\left\{\left[\begin{array}[]{cr}0&b\\ -b&0\end{array}\right]|b\in\mathbb{C}\right\}\>\>\>{\cal S}=\left\{\left[\begin{array}[]{cr}a&c\\ c&-a\end{array}\right]|a,c\in\mathbb{C}\right\}

    Since 𝒮{\cal S} is two dimensional, so is the orbit. Applying 𝒮{\cal S} to p2p_{2} gives Tp2O=(x2y2)+xyT_{p_{2}}O=\mathbb{C}\cdot(x^{2}-y^{2})+\mathbb{C}\cdot xy. Thus a normal nn to the orbit is x2+y2x^{2}+y^{2}, i.e., p2p_{2} itself. Next, we compute θ(n):VV\theta(n):V\rightarrow V. For this, we must first compute λS\lambda_{S}, the 𝒮{\cal S}-parametrization of the projection of VTp2OV\rightarrow T_{p_{2}}O, and then apply this to nn. But since n=p2n=p_{2}, this makes θ(n)\theta(n) a projection onto Tp2OT_{p_{2}}O. On the other hand, since 2{\cal H}_{2} is reductive, 𝒬=0{\cal Q}=0, and the stabilizer of p2+np_{2}+n, is O2O_{2}.

  3. 3.

    Let us consider next p1=x2p_{1}=x^{2}. The stabilizer 1{\cal H}_{1} and 𝒮{\cal S} are given below:

    1={[00c0]|c}𝒮={[ab0a]|a,c}{\cal H}_{1}=\left\{\left[\begin{array}[]{cr}0&0\\ c&0\end{array}\right]|c\in\mathbb{C}\right\}\>\>\>{\cal S}=\left\{\left[\begin{array}[]{cr}a&b\\ 0&-a\end{array}\right]|a,c\in\mathbb{C}\right\}

    We see that 𝒮p1=2ax2+2bxy{\cal S}\cdot p_{1}=2ax^{2}+2bxy, and thus, we may choose n=y2n=y^{2}. Note that =0{\cal R}=0 and 𝒬={\cal Q}={\cal H}. Let us choose 𝔮\mathfrak{q} as the element g0,0,1g_{0,0,1}, and note that 𝔮n=2xy\mathfrak{q}\cdot n=2xy. Let us compute θ(n)\theta(n) next. Since λ𝒮(y2)=0\lambda_{\cal S}(y^{2})=0, we see:

    λ𝒮[x2xyy2]=[g1/2,0,0g0,1/2,00]and thus θ(n)[x2xyy2]=[001000000][x2xyy2]\lambda_{\cal S}\cdot\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]=\left[\begin{array}[]{c}g_{1/2,0,0}\\ g_{0,1/2,0}\\ 0\end{array}\right]\>\>\mbox{and thus }\theta(n)\cdot\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]=\left[\begin{array}[]{ccc}0&0&-1\\ 0&0&0\\ 0&0&0\end{array}\right]\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]

    This gives us:

    (1+θ(n))1[x2xyy2]=[101010001][x2xyy2](1+\theta(n))^{-1}\cdot\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]=\left[\begin{array}[]{ccc}1&0&1\\ 0&1&0\\ 0&0&1\end{array}\right]\left[\begin{array}[]{c}x^{2}\\ xy\\ y^{2}\end{array}\right]

    Whence λN(1+θ(n))1(𝔮n)=0\lambda_{N}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=0 and thus n=1{\cal H}_{n}={\cal H}_{1}. The 𝒮{\cal S}-completion is given by the condition that 𝔰+λ𝒮(1+θ(n))1(𝔮n)=0\mathfrak{s}+\lambda_{\cal S}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=0. Since 𝔮n=2xy\mathfrak{q}\cdot n=2xy, we get λ𝒮(1+θ(n))1(𝔮n)=g0,1,0\lambda_{\cal S}(1+\theta(n))^{-1}(\mathfrak{q}\cdot n)=g_{0,1,0}. So 𝔰=g0,1,0\mathfrak{s}=-g_{0,-1,0}, and thus, the 𝒮{\cal S}-completion of the element g0,0,1g_{0,0,1} is g0,1,1g_{0,-1,1}, the generator of OO. Thus, the stabilizer of p2=p1+y2p_{2}=p_{1}+y^{2} matches the one computed earlier.

3 Forms and their limits

We will now specialize the above local model to the space of forms, and examine projective limits when they arise as leading terms under a 1-parameter family. This question is central to the framing of several questions in computational complexity theory. In Section 3.1 we start the investigation with a general 1-parameter family. Let gg be the limit of a form ff under the action of such a 1-parameter family A(t)A(t). So gg is the leading term of A(t)fA(t)\cdot f. We show a connection (Theorem 3.13) between the stabilizer of ff at a generic value of the parameter tt, the stabilizer of gg and the stabilizer of second leading term of A(t)fA(t)\cdot f, what we call the tangent of approach. In Section  4 we consider the case when the 1-parameter family has more structure, and is in fact a 1-parameter subgroup (1-PS). In this case one can exploit the fact that the stabilizers come with a natural grading to get a tighter relation between the stabizers of f,gf,g and the stabilizers of the tangent of approach and the tangent of exit.

3.1 Stabilizers of Limits of Forms under a 1-parameter family

Let XX be a set of indeterminates and let VV be the GL(X)GL(X)-module Symd(X)Sym^{d}(X) and f,gVf,g\in V be non-zero forms. Let 𝒢{\cal G} be the algebra gl(X)gl(X), 𝒦𝒢{\cal K}\subseteq{\cal G} be the stabilizer of ff, and its dimension be kk. Let {\cal H} be the Lie algebra stabilizing gg and its dimension be rr.

Suppose that we have an algebraic family A(t)GL(X)A(t)\subseteq GL(X), parametrized by tt such that A(1)=eA(1)=e, the identity element. Let

f(t)=A(t).f=tag+tbfb+tb+1fb+1++tDfD,f(t)=A(t).f=t^{a}g+t^{b}f_{b}+t^{b+1}f_{b+1}+\ldots+t^{D}f_{D},

By normalizing the family A(t)A(t), as in the proof of Prop. 2.12, and using suitable powers of tt, we may assume that:

f(t)=A(t).f=g+tbfb+tb+1fb+1++tDfD,f(t)=A(t).f=g+t^{b}f_{b}+t^{b+1}f_{b+1}+\ldots+t^{D}f_{D},

with fb0f_{b}\neq 0. We will write f+(t)f^{+}(t) for the sum i=bDtifi\sum_{i=b}^{D}t^{i}f_{i}. Thus limt0f(t)=g\lim_{t\rightarrow 0}f(t)=g and limt0tbf+(t)=fb\lim_{t\rightarrow 0}t^{-b}f^{+}(t)=f_{b}. We call fbf_{b} as the tangent of approach.

Assumption 3.1

Transversality Assumption. We assume that the vector space spanned by fb,,fDf_{b},\ldots,f_{D} intersects TgO(g)T_{g}O(g) trivially, where TgO(g)T_{g}O(g) is the tangent space of the GL(X)GL(X)-orbit O(g)O(g) at point gg.

Let 𝒮{\cal S} be a complement to {\cal H} in 𝒢{\cal G}. Note that the tangent space TgO(g)T_{g}O(g) of the GL(X)GL(X)-orbit of gg is precisely 𝒮g{\cal S}\cdot g and is of the same dimension as 𝒮{\cal S}. Let NN be a complement to TgO(g)T_{g}O(g) in TVgVTV_{g}\cong V. By the transversality assumption, we may assume that f+(t)Nf^{+}(t)\in N for all tt. This completes the data required to compute the local model.

For a Lie algebra element 𝔤=𝔰+𝔥\mathfrak{g}=\mathfrak{s}+\mathfrak{h}, with 𝔰𝒮\mathfrak{s}\in{\cal S} and 𝔥\mathfrak{h}\in{\cal H}, let us compute the action of 𝔤\mathfrak{g} on f(t)=g+f+(t)f(t)=g+f^{+}(t). Identifying the tangent space TVgTV_{g} with 𝒮N{\cal S}\oplus N as in Section 2.2, we have:

(𝔰+𝔥)f(t)=(𝔰+λ𝒮((1+θ(f+(t)))1(𝔥f+(t)),λN((1+θ(f+(t)))1(𝔥f+(t)).(\mathfrak{s}+\mathfrak{h})\cdot f(t)=(\mathfrak{s}+\lambda_{\cal S}((1+\theta(f^{+}(t)))^{-1}(\mathfrak{h}\cdot f^{+}(t)),\lambda_{N}((1+\theta(f^{+}(t)))^{-1}(\mathfrak{h}\cdot f^{+}(t)). (1)

Thus, for a generic t0t_{0}\in\mathbb{C}, every element 𝔨𝒦(t0)\mathfrak{k}\in{\cal K}(t_{0}), the Lie algebra of the stabilizer of f(t0)f(t_{0}), may be expressed as 𝔨=𝔥+𝔰\mathfrak{k}=\mathfrak{h}+\mathfrak{s} with 𝔥\mathfrak{h} satisfying the condition below:

λN((1+θ(f+(t0)))1((𝔥f+(t0))=0.\lambda_{N}((1+\theta(f^{+}(t_{0})))^{-1}((\mathfrak{h}\cdot f^{+}(t_{0}))=0. (2)

and 𝔰\mathfrak{s}, the 𝒮{\cal S}-completion of 𝔥\mathfrak{h} at the point f+(t0)f^{+}(t_{0}). We then have the following:

Lemma 3.2

For a generic t0t_{0}\in\mathbb{C} the stabilizer 𝒦(t0){\cal K}(t_{0}) of f(t0)f(t_{0}) is conjugate to the stabilizer 𝒦{\cal K} of ff. The dimension of 𝒦(t0){\cal K}(t_{0}) is also kk.

Proof: If t0t_{0} is so chosen so that A(t0)A(t_{0}) is invertible, then f(t0)f(t_{0}) is in the GL(X)GL(X)-orbit of ff and the assertion follows. \Box

Our objective is to construct a uniform basis {𝔨1(t),,𝔨k(t)}\{\mathfrak{k}_{1}(t),\ldots,\mathfrak{k}_{k}(t)\} so that for a generic t0t_{0}, the elements 𝔨1(t0),,𝔨k(t0)\mathfrak{k}_{1}(t_{0}),\ldots,\mathfrak{k}_{k}(t_{0}) is a \mathbb{C}-basis for 𝒦(t0){\cal K}(t_{0}).

Let u1,,umu_{1},\ldots,u_{m} be a basis of NN and let us extend it by the elements um+1=𝔰1g,,um+p=𝔰pgu_{m+1}=\mathfrak{s}_{1}\cdot g,\ldots,u_{m+p}=\mathfrak{s}_{p}\cdot g to a basis of TVg=VTV_{g}=V, where 𝔰1,,𝔰p\mathfrak{s}_{1},\ldots,\mathfrak{s}_{p} is a basis of 𝒮{\cal S}. This basis implements the isomorphism TVg=NTOgTV_{g}=N\oplus TO_{g}. Let us denote these bases by 𝔲N\mathfrak{u}_{N} and 𝔲𝒮\mathfrak{u}_{\cal S}.

Note that (1+θ(f+(t)))1(1+\theta(f^{+}(t)))^{-1} is a map VVV\rightarrow V and we may represent it as the (m+p)×(m+p)(m+p)\times(m+p) matrix (θ,t){\cal I}(\theta,t) with entries in (t)\mathbb{C}(t) such that (1+θ(f+(t)))1[𝔲N,𝔲𝒮]=[𝔲N,𝔲𝒮](θ,t)(1+\theta(f^{+}(t)))^{-1}[\mathfrak{u}_{N},\mathfrak{u}_{\cal S}]=[\mathfrak{u}_{N},\mathfrak{u}_{\cal S}]\cdot{\cal I}(\theta,t).

Lemma 3.3

Let Δ(t)\Delta(t) be the determinant of the matrix I+θ(f+(t))I+\theta(f^{+}(t)) . Then Δ=1+c.t2b+ higher degree terms\Delta=1+c.t^{2b}+\mbox{ higher degree terms} (where cc\in\mathbb{C}). The matrix Δ(θ,t)\Delta\cdot{\cal I}(\theta,t) has entries in [t]\mathbb{C}[t] and is of the following form:

Δ(θ,t)=Itbθ(fb)+ higher degree terms.\Delta\cdot{\cal I}(\theta,t)=I-t^{b}\theta(f_{b})+\text{ higher degree terms}.

Proof: By linearity of θ\theta, we have:

θ(f+(t))=tbθ(fb)+tb+1θ(fb+1)+ higher degree terms.\theta(f^{+}(t))=t^{b}\theta(f_{b})+t^{b+1}\theta(f_{b+1})+\text{ higher degree terms}.

Let J=I+θ(f+(t))J=I+\theta(f^{+}(t)) in the above basis, then Jij[t]J_{ij}\in\mathbb{C}[t] with the following properties: (i) for iji\neq j, we have tbt^{b} divides JijJ_{ij}, and (ii) tbt^{b} divides Jii1J_{ii}-1. The first assertion on Δ(t)\Delta(t), the determinant of JJ, follows. For the second assertion, again for iji\neq j each minor J~ji(t)\tilde{J}_{ji}(t) of J is tbθ(fb)ij+ higher terms-t^{b}\theta(f_{b})_{ij}+\mbox{ higher terms}, while for i=ji=j, we have J~ii=1tbθ(fb)ii\tilde{J}_{ii}=1-t^{b}\theta(f_{b})_{ii} +higher terms. This proves the second part. \Box

Thus, if we express (θ,t){\cal I}(\theta,t) as a matrix with entries in [[t]]\mathbb{C}[[t]], then:

(θ,t)=Itbθ(fb)+higher terms.{\cal I}(\theta,t)=I-t^{b}\theta(f_{b})+\mbox{higher terms}.

Let N(t):N{\cal M}_{N}(t):{\cal H}\rightarrow N be the map 𝔥λN(1+θ(f+(t)))1(𝔥f+(t))\mathfrak{h}\rightarrow\lambda_{N}\circ(1+\theta(f^{+}(t)))^{-1}(\mathfrak{h}\cdot f^{+}(t)) and 𝒮(t):𝒮{\cal M}_{\cal S}(t):{\cal H}\rightarrow{\cal S} be the map 𝔥λ𝒮(1+θ(f+(t)))1(𝔥f+(t))\mathfrak{h}\rightarrow\lambda_{\cal S}\circ(1+\theta(f^{+}(t)))^{-1}(\mathfrak{h}\cdot f^{+}(t)). Note that N{\cal M}_{N} and 𝒮{\cal M}_{\cal S} are linear maps which depend on the parameter tt. The kernel ker(N(t))ker({\cal M}_{N}(t)) determines the {\cal H}-part of every element of 𝒦(t){\cal K}(t), while 𝒮{\cal M}_{\cal S} is the 𝒮{\cal S}-completion at the point f+(t)f^{+}(t).

Let us now compute matrices for MNM_{N} and M𝒮M_{\cal S} for N{\cal M}_{N} and 𝒮{\cal M}_{\cal S}, respectively.

Let 𝔥1,,𝔥r\mathfrak{h}_{1},\ldots,\mathfrak{h}_{r} be a basis of {\cal H} and let us denote this sequence by 𝔟¯\overline{\mathfrak{b}}. Let us form the matrix MNM_{N}, an m×rm\times r matrix with entries in (t)\mathbb{C}(t) as follows. Let N(𝔥i)=k=1m(MN)kiuk{\cal M}_{N}(\mathfrak{h}_{i})=\sum_{k=1}^{m}(M_{N})_{ki}u_{k}. Similarly, let the p×rp\times r-matrix M𝒮M_{\cal S} be defined as follows: Let 𝒮(𝔥i)=k=1p(M𝒮)ki𝔰k{\cal M}_{\cal S}(\mathfrak{h}_{i})=\sum_{k=1}^{p}(M_{\cal S})_{ki}\mathfrak{s}_{k}. We collect this information in the lemma below:

Lemma 3.4

In the above notation, we have:

  1. 1.

    The matrices MN(t)M_{N}(t) and M𝒮(t)M_{\cal S}(t) satisfy the following equations:

    𝔲NMN(t)=λN(θ,t)(𝔟¯f+(t))=[N(𝔟¯)] and 𝔲𝒮M𝒮(t)=λ𝒮(θ,t)(𝔟¯f+(t))\mathfrak{u}_{N}M_{N}(t)=\lambda_{N}\cdot{\cal I}(\theta,t)\cdot(\overline{\mathfrak{b}}\cdot f^{+}(t))=[{\cal M}_{N}(\overline{\mathfrak{b}})]\>\>\mbox{ and }\>\mathfrak{u}_{\cal S}M_{\cal S}(t)=\lambda_{\cal S}\cdot{\cal I}(\theta,t)\cdot(\overline{\mathfrak{b}}\cdot f^{+}(t))
  2. 2.

    Let 𝒦(t0){\cal K}(t_{0}) be the stabilizer of f(t0)f(t_{0}) for some generic t0t_{0}\in\mathbb{C} and let 𝔨𝒦(t0)\mathfrak{k}\in{\cal K}(t_{0}). Express 𝔨\mathfrak{k} as 𝔨=𝔥+𝔰\mathfrak{k}=\mathfrak{h}+\mathfrak{s} and let 𝔥=𝔟¯α¯\mathfrak{h}=\overline{\mathfrak{b}}\cdot\overline{\alpha}, i.e., a linear combination α¯\overline{\alpha} of the basis elements 𝔟¯\overline{\mathfrak{b}}. Then

    𝔲NMN(t0)α¯=0 and 𝔰g=𝔲𝒮M𝒮(t0)α¯.\mathfrak{u}_{N}M_{N}(t_{0})\overline{\alpha}=0\mbox{ and }\mathfrak{s}\cdot g=-\mathfrak{u}_{\cal S}M_{\cal S}(t_{0})\overline{\alpha}.
  3. 3.

    Conversely, any element 𝔨=𝔥+𝔰\mathfrak{k}=\mathfrak{h}+\mathfrak{s} with 𝔥=𝔟¯α¯\mathfrak{h}=\overline{\mathfrak{b}}\cdot\overline{\alpha}, satisfying the above conditions, is an element of 𝒦(t0){\cal K}(t_{0}).

Proof: All the assertions follow from the construction of MNM_{N} and M𝒮M_{\cal S} and the requirements of Eqn. 1. \Box

Lemma 3.5
  1. (i)

    Let (t)\mathfrak{R}\subseteq\mathbb{C}(t) be the localization of [t]\mathbb{C}[t] at Δ\Delta. In other words, ={p/Δd|p[t] and d0}\mathfrak{R}=\{p/\Delta^{d}|p\in\mathbb{C}[t]\mbox{ and }d\geq 0\}. Then, the entries of MN(t)M_{N}(t) and M𝒮(t)M_{\cal S}(t) are elements of \mathfrak{R}.

  2. (ii)

    tbt^{b} divides the matrix M𝒮(t)M_{\cal S}(t) and MN(t)M_{N}(t) in \mathfrak{R}. In particular, M𝒮(0)=0M_{\cal S}(0)=0.

  3. (iii)

    For any generic t0t_{0}\in\mathbb{C}, the rank of MN(t0)M_{N}(t_{0}) is a constant and equals the (t)\mathbb{C}(t)-rank of MN(t)M_{N}(t). This equals dim()dim(𝒦)dim({\cal H})-dim({\cal K}), i.e., rkr-k.

  4. (iv)

    The (t)\mathbb{C}(t)-rank of M𝒮(t)M_{\cal S}(t) is zero iff f(t)N{\cal H}\cdot f(t)\subseteq N.

Proof: The matrix MN(t)M_{N}(t) is the product of operations expressed in appropriate bases, viz., (a) λN:VN\lambda_{N}:V\rightarrow N, (b), (1+θ(f+(t))1:VV(1+\theta(f^{+}(t))^{-1}:V\rightarrow V and (c) (𝔟¯f+(t)):V(\overline{\mathfrak{b}}\cdot f^{+}(t)):{\cal H}\rightarrow V. Let us look at each one in turn. The first is a projection and is a matrix over \mathbb{C} of rank m+pm+p. For (b), this is precisely the matrix (θ,t){\cal I}(\theta,t) which we have shown has entries in \mathfrak{R}. Moreover, we know that (θ,0){\cal I}(\theta,0) is the identity matrix. Finally, since {\cal H} acts linearly on VV and tbt^{b} divides f+(t)f^{+}(t), the matrix (𝔟¯f+(t))(\overline{\mathfrak{b}}\cdot f^{+}(t)) has entries in tb[t]t^{b}\cdot\mathbb{C}[t]. Thus the product, viz., MN(t)M_{N}(t) has entries in \mathfrak{R} and tbt^{b} divides it (in \mathfrak{R}). The same reasoning holds for M𝒮M_{\cal S} as well. This proves (i) and (ii).

For (iii), by lemma 3.4, the dimension of 𝒦(t0){\cal K}(t_{0}) for a generic t0t_{0} equals kk and this must be the nullity of MN(t0)M_{N}(t_{0}). But this is the same as nullity of the matrix MN(t)M_{N}(t) as a matrix over (t)\mathbb{C}(t). Thus k=rrank(t)(MN(t))=rrank(MN(t0))k=r-rank_{\mathbb{C}(t)}(M_{N}(t))=r-rank_{\mathbb{C}}(M_{N}(t_{0})). This proves (iii).

Finally, for (iv), the condition that M𝒮M_{\cal S} is identically zero is equivalent to 𝒮{\cal M}_{\cal S} being zero, which is equivalent to f(t)N{\cal H}\cdot f(t)\subset N. \Box

Lemma 3.5 now allows us to construct 𝒦(t){\cal K}(t) as the column annihilator space of MN(t)M_{N}(t). Since rank(MN(t))=rkrank(M_{N}(t))=r-k, we may choose column vectors α¯1(t),,α¯k(t)(t)r\overline{\alpha}_{1}(t),\ldots,\overline{\alpha}_{k}(t)\in\mathbb{C}(t)^{r} such that MNα¯i=0M_{N}\cdot\overline{\alpha}_{i}=0. Since the entries of MNM_{N} are \mathfrak{R}, we may assume that these column annihilators may be chosen over [t]\mathbb{C}[t] and in a standard form as given by the following lemma:

Lemma 3.6

Let A=[α¯1,,α¯k]A=[\overline{\alpha}_{1},\ldots,\overline{\alpha}_{k}] be the r×kr\times k-matrix with entries in (t)\mathbb{C}(t) and the columns α¯1,,α¯k\overline{\alpha}_{1},\ldots,\overline{\alpha}_{k} as above. Then we may assume that there is another basis (over (t)\mathbb{C}(t)) α¯1,,α¯k\overline{\alpha}^{\prime}_{1},\ldots,\overline{\alpha}^{\prime}_{k}, for the column space of AA and a matrix A=[α¯1,,α¯k]A^{\prime}=[\overline{\alpha}^{\prime}_{1},\ldots,\overline{\alpha}^{\prime}_{k}] such that all entries of AA^{\prime} are in [t]\mathbb{C}[t] and A(0)A^{\prime}(0) is of rank kk. Moreover, if cc is any column in the column space of AA, such that c(0)c(0) is defined, then c(0)c(0) is in the column space of A(0)A^{\prime}(0).

Proof: We first note that elementary column operations on AA do not change the corresponding 𝒦(t){\cal K}(t). Next, by multiplying by a suitable polynomial p(t)p(t) and pulling out an appropriate power of tt from each column, we may assume that AA has entries in [t]\mathbb{C}[t] and that A(0)A(0) exists and each of its column is non-zero. Finally, if a column, say, α¯k\overline{\alpha}_{k} of AA is such that α¯k(0)=i=1k1μiα¯i(0)\overline{\alpha}_{k}(0)=\sum_{i=1}^{k-1}\mu_{i}\overline{\alpha}_{i}(0), with μi\mu_{i}\in\mathbb{C}, then we may replace α¯k\overline{\alpha}_{k} by ta(α¯ki=1k1μiα¯i)t^{-a}(\overline{\alpha}_{k}-\sum_{i=1}^{k-1}\mu_{i}\overline{\alpha}_{i}) for some suitable a>0a>0, to get a new column of a lower degree. This process may be repeated till the assertions in the lemma are satisfied. Let us come to the second part. Let R={r1,,rk}R=\{r_{1},\ldots,r_{k}\} be the row indices such that det(A(0)[R])0det(A^{\prime}(0)[R])\neq 0 and let Δ(t)=det(A(t)[R])\Delta(t)=det(A^{\prime}(t)[R]). Since cc is in the column space of AA^{\prime}, there are polynomials p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{C}[t] such that Δ(t)c=i=1kpiα¯i\Delta(t)c=\sum_{i=1}^{k}p_{i}\overline{\alpha}^{\prime}_{i}. The result follows. \Box

Henceforth, we assume that the chosen α1¯,,αk¯\overline{\alpha_{1}},\ldots,\overline{\alpha_{k}} have the properties assured to us by lemma 3.6. This has an immediate corollary.

Proposition 3.7
  1. 1.

    There is a (t)\mathbb{C}(t)-basis {𝔨i(t)}i=1k\{\mathfrak{k}_{i}(t)\}_{i=1}^{k} of 𝒦(t){\cal K}(t), the stabilizer Lie algebra of f(t)f(t) and a large number DD such that

    𝔨i(t)=j=0D(𝔰ij+𝔥ij)tj\mathfrak{k}_{i}(t)=\sum_{j=0}^{D}(\mathfrak{s}_{ij}+\mathfrak{h}_{ij})t^{j}

    for suitable elements 𝔰ij𝒮,𝔥ij\mathfrak{s}_{ij}\in{\cal S},\mathfrak{h}_{ij}\in{\cal H}.

  2. 2.

    The elements 𝔰ij\mathfrak{s}_{ij} are zero for all ii, and for all j=0,,b1j=0,\ldots,b-1. As a result, we also have:

    𝔨i(t)=𝔥i(t)+tb𝔰i(t)\mathfrak{k}_{i}(t)=\mathfrak{h}_{i}(t)+t^{b}\mathfrak{s}_{i}(t)

    where 𝔥i(t)[t]\mathfrak{h}_{i}(t)\in\mathbb{C}[t]\otimes{\cal H} and 𝔰i(t)[t]𝒮\mathfrak{s}_{i}(t)\in\mathbb{C}[t]\otimes{\cal S}.

  3. 3.

    The space 𝒦0={𝔨1(0),,𝔨k(0)}{\cal K}_{0}=\mathbb{C}\cdot\{\mathfrak{k}_{1}(0),\ldots,\mathfrak{k}_{k}(0)\}, the subspace formed by \mathbb{C}-linear combinations of the leading terms 𝔨i(0)=𝔨i0=𝔥i0\mathfrak{k}_{i}(0)=\mathfrak{k}_{i0}=\mathfrak{h}_{i0} is a Lie subalgebra of {\cal H} and of dimension kk. Moreover, if 𝔨(t)𝒦(t)\mathfrak{k}(t)\in{\cal K}(t) is any element such that 𝔨(0)\mathfrak{k}(0) is defined, then 𝔨(0)𝒦0\mathfrak{k}(0)\in{\cal K}_{0}.

  4. 4.

    For any element 𝔥𝒦0\mathfrak{h}\in{\cal K}_{0}, we have λN(𝔥fb)=0\lambda_{N}(\mathfrak{h}\cdot f_{b})=0 and thus, there is an 𝔰𝒮\mathfrak{s}\in{\cal S} such that 𝔰g+λ𝒮(𝔥fb)=0\mathfrak{s}\cdot g+\lambda_{\cal S}(\mathfrak{h}\cdot f_{b})=0

Proof: For (1), let us assume that α¯1(t),,α¯k(t)\overline{\alpha}_{1}(t),\ldots,\overline{\alpha}_{k}(t) are as in Lemma 3.6. We may then compute the element 𝔥i(t)=𝔟α¯i(t)\mathfrak{h}_{i}(t)=\mathfrak{b}\cdot\overline{\alpha}_{i}(t). Next, we define 𝔰i(t)\mathfrak{s}_{i}(t) such that 𝔰i(t)g=𝔲𝒮M𝒮α¯i(t)\mathfrak{s}_{i}(t)\cdot g=-\mathfrak{u}_{\cal S}\cdot M_{\cal S}\overline{\alpha}_{i}(t). Then 𝔨i(t)=𝔰i(t)+𝔥i(t)\mathfrak{k}_{i}(t)=\mathfrak{s}_{i}(t)+\mathfrak{h}_{i}(t) stabilizes f(t)f(t). Note that these are elements of 𝒢[t]{\cal G}\otimes\mathbb{C}[t], i.e., linear combinations of Lie algebra elements with coefficients in [t]\mathbb{C}[t]. Assertion (1) is an expansion of these as a polynomial in tt and DD is a bound on the degree of these polynomials. Assertion (2) follows from the fact that the element 𝔰i(t)=j=0D𝔰ijtj\mathfrak{s}_{i}(t)=\sum_{j=0}^{D}\mathfrak{s}_{ij}t^{j} arises from the matrix equation 𝔰i(t)g=𝔲𝒮M𝒮αi¯(t)\mathfrak{s}_{i}(t)\cdot g=-\mathfrak{u}_{\cal S}M_{\cal S}\cdot\overline{\alpha_{i}}(t), and that tbt^{b} divides M𝒮M_{\cal S} as in lemma 3.5 (ii).

Let us now come to (3). Since 𝒦(t){\cal K}(t) is a Lie algebra, for any two elements 𝔨i(t),𝔨j(t)\mathfrak{k}_{i}(t),\mathfrak{k}_{j}(t), we have:

[𝔨i(t),𝔨j(t)]=[𝔨i0,𝔨j0]+([𝔨i0,𝔨j1][𝔨i1,𝔨j0])t1+\begin{array}[]{rcl}[\mathfrak{k}_{i}(t),\mathfrak{k}_{j}(t)]&=&[\mathfrak{k}_{i0},\mathfrak{k}_{j0}]+([\mathfrak{k}_{i0},\mathfrak{k}_{j1}][\mathfrak{k}_{i1},\mathfrak{k}_{j0}])t^{1}+\ldots\\ \end{array}

This implies that limt0[𝔨i(t),𝔨j(t)]\lim_{t\rightarrow 0}[\mathfrak{k}_{i}(t),\mathfrak{k}_{j}(t)] exists and equals [𝔨i0,𝔨j0][\mathfrak{k}_{i0},\mathfrak{k}_{j0}]. Next, note that lemma 3.6 implies that 𝒦(t){\cal K}(t) is a kk-dimensional subspace of gl(X)(t)gl(X)\otimes\mathbb{C}(t) with the basis {𝔨1(t),,𝔨k(t)}\{\mathfrak{k}_{1}(t),\ldots,\mathfrak{k}_{k}(t)\} such that {𝔨1(0),,𝔨k(0)}\{\mathfrak{k}_{1}(0),\ldots,\mathfrak{k}_{k}(0)\} continue to be linearly independent over \mathbb{C} in gl(X)gl(X). Thus we may express [𝔨i(t),𝔨j(t)][\mathfrak{k}_{i}(t),\mathfrak{k}_{j}(t)] in this basis as

[𝔨i(t),𝔨j(t)]=u=1kβuij(t)𝔨u(t),\begin{array}[]{rcl}[\mathfrak{k}_{i}(t),\mathfrak{k}_{j}(t)]&=&\sum_{u=1}^{k}\beta^{ij}_{u}(t)\mathfrak{k}_{u}(t),\\ \end{array}

where βuij(t)(t)\beta^{ij}_{u}(t)\in\mathbb{C}(t). The condition on Δ(t)\Delta(t) in the proof of lemma 3.6 give us that βij(0)\beta_{ij}(0) exist. Taking limits on both sides gives us:

[𝔨i(0),𝔨j(0)]=u=1kβuij(0)𝔨u(0).\begin{array}[]{rcl}[\mathfrak{k}_{i}(0),\mathfrak{k}_{j}(0)]&=&\sum_{u=1}^{k}\beta^{ij}_{u}(0)\mathfrak{k}_{u}(0).\\ \end{array}

This proves that 𝒦0{\cal K}_{0} is a Lie algebra. Note that since 𝔰i0=0\mathfrak{s}_{i0}=0 for all ii, the elements 𝔨i0\mathfrak{k}_{i0} are elements of {\cal H} and thus 𝒦0{\cal K}_{0}\subseteq{\cal H}. The assertion that {hi0}\{h_{i0}\} are linearly independent, and the second assertion, follow from Lemma 3.6.

Finally, let us come to the (4). Recall that 𝔟\mathfrak{b} is the ordered basis {𝔥i,,𝔥r}\{\mathfrak{h}_{i},\ldots,\mathfrak{h}_{r}\}. We see that the condition that 𝔥i(t)=𝔟α¯i(t)\mathfrak{h}_{i}(t)=\mathfrak{b}\cdot\overline{\alpha}_{i}(t) where MNα¯i(t)=0M_{N}\cdot\overline{\alpha}_{i}(t)=0 is tantamount to asserting that

𝔲NMNα¯i(t)=λN(1+θ(f+))1(𝔥i(t)f+(t))=0,\mathfrak{u}_{N}M_{N}\overline{\alpha}_{i}(t)=\lambda_{N}\cdot(1+\theta(f^{+}))^{-1}\cdot(\mathfrak{h}_{i}(t)\cdot f^{+}(t))=0,

and, setting 𝔰ig=𝔲𝒮M𝒮α¯i\mathfrak{s}_{i}\cdot g=-\mathfrak{u}_{\cal S}M_{\cal S}\overline{\alpha}_{i} that,

𝔰i(t)g+λ𝒮(1+θ(f+))1(𝔥i(t)f+(t))=0.\mathfrak{s}_{i}(t)\cdot g+\lambda_{\cal S}\cdot(1+\theta(f^{+}))^{-1}\cdot(\mathfrak{h}_{i}(t)\cdot f^{+}(t))=0.

Both together imply that 𝔨i(t)\mathfrak{k}_{i}(t) stabilizes f(t)f(t). Now comparing coefficient of the lowest degree, i.e., tbt^{b}, on both sides of the two equations, gives us:

λN𝔥i0fb=0 and 𝔰i,bg+λ𝒮𝔥i0fb=0.\lambda_{N}\cdot\mathfrak{h}_{i0}\cdot f_{b}=0\>\mbox{ and }\>\mathfrak{s}_{i,b}\cdot g+\lambda_{\cal S}\cdot\mathfrak{h}_{i0}\cdot f_{b}=0.

This finishes the proof of the proposition. \Box

Remark 3.8

Surprisingly, the fact that the next lowest degree term in (1+θ(f+))1(1+\theta(f^{+}))^{-1} is tbθ(fb)-t^{b}\theta(f_{b}) is neither used nor required.

We now have the following proposition.

Proposition 3.9

Let 𝒦0{\cal K}_{0} be as above. Then:

  1. 1.

    Let NVN\subseteq V be as above. Define the \star action of {\cal H} on NN as below:

    𝔥n=λN(𝔥n)\mathfrak{h}\star n=\lambda_{N}(\mathfrak{h}\cdot n)

    Then :×NN\star:{\cal H}\times N\rightarrow N is a Lie algebra action on NN and matches the action of the quotient module N¯=V/TOg\overline{N}=V/TO_{g}. The isomorphism from NN to N¯\overline{N} is given by nn¯=n+TOgn\rightarrow\overline{n}=n+TO_{g}.

  2. 2.

    The subalgebra 𝒦0{\cal K}_{0}\subseteq{\cal H} stabilizes the element fb¯N¯\overline{f_{b}}\in\overline{N}. In other words, 𝒦0b{\cal K}_{0}\subseteq{\cal H}_{b}\subseteq{\cal H}, the stabilizer of fb¯N¯\overline{f_{b}}\in\overline{N}.

Proof: Part (1) is straightforward.

For part (2), note that Prop. 3.7 (4) tells us that λN(𝔥fb)=0\lambda_{N}(\mathfrak{h}\cdot f_{b})=0, for any 𝔥𝒦0\mathfrak{h}\in{\cal K}_{0}. Thus, 𝔥\mathfrak{h} stabilizes fb¯\overline{f_{b}}. \Box

Definition 3.10

Let 𝒦(t){\cal K}(t) be the stabilizer of f(t)f(t) and let 𝔨i(t)=𝔥i(t)+𝔰i(t)\mathfrak{k}_{i}(t)=\mathfrak{h}_{i}(t)+\mathfrak{s}_{i}(t), for i=1,,ki=1,\ldots,k, be a basis for 𝒦(t){\cal K}(t) as above. We define the space (t){\cal H}(t) as the [t]\mathbb{C}[t]-space spanned by {𝔥i(t)|i=1,,k}\{\mathfrak{h}_{i}(t)|i=1,\ldots,k\}. For a t0t_{0}\in\mathbb{C}, let (t0){\cal H}(t_{0}) be the \mathbb{C}-space spanned by the vectors {𝔥i(t0)|i=1,,k}\{\mathfrak{h}_{i}(t_{0})|i=1,\ldots,k\}. Note that, by definition, (0)=𝒦0{\cal H}(0)={\cal K}_{0}.

Lemma 3.11

Let t0t_{0}\in\mathbb{C} be generic and 𝔨=𝔰+𝔥\mathfrak{k}=\mathfrak{s}+\mathfrak{h} be an element of 𝒦(t0){\cal K}(t_{0}), then 𝔥(t0)\mathfrak{h}\in{\cal H}(t_{0}).

Proof: The assertion is straightforward and follows from the fact that for generic t0t_{0}, the rank of the matrix A(t0)A(t_{0}) of lemma 3.6 is kk. \Box

We now come to the main theorem of this section which summarizes the above results. We will recall the definition of 𝒮{\cal S}-completion.

Definition 3.12

For any element 𝔥\mathfrak{h}\in{\cal H}, and nNn\in N, we define its 𝒮{\cal S} completion as the element 𝔥+𝔰\mathfrak{h}+\mathfrak{s}, with 𝔰𝒮\mathfrak{s}\in{\cal S} such that λ𝒮(𝔥n)=𝔰g\lambda_{\cal S}(\mathfrak{h}\cdot n)=\mathfrak{s}\cdot g. For a set 𝒥{\cal J}\subseteq{\cal H}, we define the 𝒮{\cal S}-completion of 𝒥{\cal J} as the collection of all 𝒮{\cal S}-completions of every element 𝔧𝒥\mathfrak{j}\in{\cal J}.

Note that since the map 𝒮𝒮g{\cal S}\rightarrow{\cal S}\cdot g is a bijection, the 𝒮{\cal S}-completion of an element 𝔥\mathfrak{h}\in{\cal H} always exists and is unique.

Theorem 3.13

Let A(t)f=f(t)=g+tbfb+ higher terms,A(t)\cdot f=f(t)=g+t^{b}f_{b}+\mbox{ higher terms}, as above, with 𝒦(t){\cal K}(t) of generic dimension kk as its stabilizer, and fbf_{b}, the tangent of approach. Then

  1. (i)

    There is a subalgebra 𝒦0{\cal K}_{0} of b{\cal H}_{b}\subseteq{\cal H} the stabilizer of fb¯\overline{f_{b}} for the \star-action of {\cal H} whose dimension is the same as the dimension of 𝒦(t0){\cal K}(t_{0}) for a generic t0t_{0}\in\mathbb{C}.

  2. (ii)

    There is a basis {𝔨i(t)|i=1,,k}\{\mathfrak{k}_{i}(t)|i=1,\ldots,k\} of 𝒦(t){\cal K}(t) such that {𝔨i(0)|i=1,,k}\{\mathfrak{k}_{i}(0)|i=1,\ldots,k\} is a basis for 𝒦0{\cal K}_{0}. Moreover, 𝒦0=(0){\cal K}_{0}={\cal H}(0).

  3. (iii)

    Let (t){\cal H}(t) be the [t]\mathbb{C}[t]-span of the vectors 𝔥1(t),,𝔥k(t)\mathfrak{h}_{1}(t),\ldots,\mathfrak{h}_{k}(t). Then, for any generic t0t_{0}\in\mathbb{C}, the subspace (t0){\cal H}(t_{0})\subseteq{\cal H} has dimension kk and 𝒦(t0){\cal K}(t_{0}) is the 𝒮{\cal S}-completion of (t0){\cal H}(t_{0}) for the point f+(t0)f^{+}(t_{0}).

Proof: Part (i) and (ii) are shown in Prop. 3.7. Part (iii) follows from lemma 3.11. This proves the theorem. \Box

4 Projective limits under the action of 1-ps

We specialize the results of the previous section to the case when the 1-parameter family is a 1-parameter subgroup λ(t)\lambda(t) (1-PS). Each λ\lambda effectively partitions the variable set XX into several parts X=X1XkX=X_{1}\cup\ldots\cup X_{k} with λ\lambda operating on XiX_{i} simply as λ(t)(Xi)=tdiXi\lambda(t)(X_{i})=t^{d_{i}}X_{i}. In Section 4.1 we tackle the two block case, when the variable set XX is partitioned into two parts YY and ZZ, with the 1-ps acting trivially on YY and scaling by tt the variables in ZZ. Proposition 4.10 shows that in this case either there are elements which stabilizes f,gf,g and the tangent of approach and exit, i.e., a triple stabilizer, or the limiting Lie algebra 𝒦0{\cal K}_{0} is nilpotent. Next, we study the general 1-ps case in Section 4.2. We prove Proposition 4.15, which is similar in spirit to that proved in the two block 1-ps.

4.1 Two block λ\lambda

Let us partition XX as X=YZX=Y\cup Z. Let λ:GL(X)\lambda:\mathbb{C}^{*}\rightarrow GL(X) be a 1-PS such that λ(t)y=y\lambda(t)y=y for all yYy\in Y but λ(t)z=tz\lambda(t)z=tz. Let us denote by ViV_{i} the space Symdi(Y)Symi(Z)Sym^{d-i}(Y)\otimes Sym^{i}(Z) and note that Symd(X)=i=0dViSym^{d}(X)=\oplus_{i=0}^{d}V_{i}. Moreover, for any viViv_{i}\in V_{i}, we have λ(t)vi=tivi\lambda(t)v_{i}=t^{i}v_{i}. Let ff and gg be forms so that:

λ(t)f=f(t)=tag+tbfb++tdfd\lambda(t)f=f(t)=t^{a}g+t^{b}f_{b}+\ldots+t^{d}f_{d}

with fb0f_{b}\neq 0. In other words, gVag\in V_{a} is the leading term for the action of λ\lambda on ff and fbf_{b} is the tangent of approach. The family A(t)=ta/dλ(t)A(t)=t^{-a/d}\lambda(t) is a suitable family and satisfies the conditions specified in the previous section. Whence, if the transversality assumption holds, then all the results, including Theorem 3.13 are available. The special structure of A(t)A(t) as above allows some more observations. We continue to use the same notation 𝒦(t){\cal K}(t), (t){\cal H}(t) from the previous section.

First note that λ(t)\lambda(t) acts on 𝒢{\cal G} by conjugation whence, we have the weight space decomposition 𝒢1𝒢0𝒢1{\cal G}_{-1}\oplus{\cal G}_{0}\oplus{\cal G}_{1}, where 𝒢1=Hom(Z,Y){\cal G}_{-1}=Hom(Z,Y), 𝒢0=Hom(Y,Y)Hom(Z,Z){\cal G}_{0}=Hom(Y,Y)\oplus Hom(Z,Z) and 𝒢1=Hom(Y,Z){\cal G}_{1}=Hom(Y,Z). Note that λ(t)𝔤iλ(t)1=ti𝔤i\lambda(t)\mathfrak{g}_{i}\lambda(t)^{-1}=t^{i}\mathfrak{g}_{i} for any 𝔤i𝒢i\mathfrak{g}_{i}\in{\cal G}_{i}, and that 𝒢iVj=Vi+j{\cal G}_{i}\cdot V_{j}=V_{i+j}.

Since gVag\in V_{a} and is homogeneous in the degree in YY, we have λ(t)g=tag\lambda(t)g=t^{a}g is the leading term of λ(t)f\lambda(t)f. Since the Lie algebra stabilizers of tagt^{a}g and gg are the same, we have:

Lemma 4.1

In the above notation, the stabilizer {\cal H} of gg is graded, i.e., =i=1,0,1i{\cal H}=\oplus_{i=-1,0,1}{\cal H}_{i} such that λ(t)iλ(t)1=tii\lambda(t){\cal H}_{i}\lambda(t)^{-1}=t^{i}{\cal H}_{i}. Moreover, we may choose 𝒮i{\cal S}_{i} as the complement of i𝒢i{\cal H}_{i}\subseteq{\cal G}_{i}, then 𝒮=i=1,0,1𝒮i{\cal S}=\oplus_{i=-1,0,1}{\cal S}_{i} is a complement of {\cal H} in 𝒢{\cal G}, and λ(t)𝒮iλ(t)1ti𝒮i\lambda(t){\cal S}_{i}\lambda(t)^{-1}\subseteq t^{i}{\cal S}_{i}. Moreover, let π𝒮\pi_{\cal S} and π\pi_{\cal H} be the projections from 𝒢=𝒮{\cal G}={\cal S}\oplus{\cal H}, to the first and second components, then λ(t)\lambda(t) intertwines with these projections. In other words, for any 𝔤𝒢\mathfrak{g}\in{\cal G}, we have π𝒮((λ(t)𝔤λ(t)1)=λ(t)π𝒮(𝔤)λ(t)1\pi_{\cal S}((\lambda(t)\mathfrak{g}\lambda(t)^{-1})=\lambda(t)\pi_{\cal S}(\mathfrak{g})\lambda(t)^{-1} and π((λ(t)𝔤λ(t)1)=λ(t)π(𝔤)λ(t)1\pi_{\cal H}((\lambda(t)\mathfrak{g}\lambda(t)^{-1})=\lambda(t)\pi_{\cal H}(\mathfrak{g})\lambda(t)^{-1}

Thus, TgO(g)=𝒮gVa1VaVa+1T_{g}O(g)={\cal S}\cdot g\subseteq V_{a-1}\oplus V_{a}\oplus V_{a+1}. The transversality condition translates into the requirement that fbTgO(g)f_{b}\not\in T_{g}O(g).

In view of Lemma 4.1 and the resulting discussion, the stabilization conditions of Theorem 3.13 can be further studied. We present this analysis in the Appendix B.

Lemma 4.2

Let u,vu,v\in\mathbb{C} be generic. Then λ(u)(v)λ(u)1(uv)\lambda(u){\cal H}(v)\lambda(u)^{-1}\subseteq{\cal H}(uv).

Proof: Let 𝒦{\cal K} be the stabilizer of f=f(1)f=f(1). Now 𝒦(u)=λ(u)𝒦λ(u)1{\cal K}(u)=\lambda(u){\cal K}\lambda(u)^{-1} and that λ(uv)=λ(u)λ(v)\lambda(uv)=\lambda(u)\lambda(v), tells us that 𝒦(uv)=λ(u)𝒦(v)λ(u)1{\cal K}(uv)=\lambda(u){\cal K}(v)\lambda(u)^{-1}. Moreover, note that π(𝒦(t0))=(t0)\pi_{\cal H}({\cal K}(t_{0}))={\cal H}(t_{0}), for any generic t0t_{0}\in\mathbb{C}, and thus may be applied to 𝒦(uv){\cal K}(uv) above. This proves the lemma. \Box

We also have the important:

Proposition 4.3

The algebra 𝒦0{\cal K}_{0} has a weight space decomposition 𝒦0=𝒦0,1𝒦0,0𝒦0,1{\cal K}_{0}={\cal K}_{0,-1}\oplus{\cal K}_{0,0}\oplus{\cal K}_{0,1}, where 𝒦0,ii{\cal K}_{0,i}\subseteq{\cal H}_{i}.

Proof: Let {𝔥i|i=1,,k}\{\mathfrak{h}_{i}|i=1,\ldots,k\} be a basis of (1){\cal H}(1) and let 𝔥i=j𝔥ij\mathfrak{h}_{i}=\oplus_{j}\mathfrak{h}_{ij} be its the weight space decomposition. We may then write:

𝔟i(t)=jtj𝔥ij\mathfrak{b}_{i}(t)=\sum_{j}t^{j}\mathfrak{h}_{ij}

The previous lemma implies that {𝔟i(t)|i=1,,k}\{\mathfrak{b}_{i}(t)|i=1,\ldots,k\} is another (t)\mathbb{C}(t)-basis for (t){\cal H}(t). The leading terms of this basis would decompose by weights and generate (0)=𝒦0{\cal H}(0)={\cal K}_{0}. \Box

Lemma 4.4
  1. 1.

    Let =logt(λ(t))gl(X)\ell=\log_{t}(\lambda(t))\in gl(X), i.e., =diag([0,,0,1,,1])\ell=diag([0,\ldots,0,1,\ldots,1]) in a suitable basis.Then for any 𝔤𝒢i\mathfrak{g}\in{\cal G}_{i}, [,𝔤]=i𝔤[\ell,\mathfrak{g}]=i\mathfrak{g}. For any vViv\in V_{i}, v=iv\ell\cdot v=i\cdot v.

  2. 2.

    The algebra b{\cal H}_{b} is graded. Let =logt(ta/dλ(t))=a/dI\ell^{\prime}=\log_{t}(t^{-a/d}\lambda(t))=\ell-a/d\cdot I. Then \ell^{\prime}\in{\cal H}, \ell^{\prime} normalizes both b{\cal H}_{b} and 𝒦0{\cal K}_{0}, but b\ell^{\prime}\not\in{\cal H}_{b}.

Proof: The first part of (1) is straightforward. For the action of \ell on VV, note that \ell is also the differential operator zZzz\sum_{z\in Z}z\frac{\partial}{\partial z}. Thus v=iv\ell v=i\cdot v for vViv\in V_{i}. Coming to (2), note that fb¯\overline{f_{b}} is homogeneous and {\cal H} is graded, hence b{\cal H}_{b} is graded. [,b,i]b,i[\ell^{\prime},{\cal H}_{b,i}]\subseteq{\cal H}_{b,i}. By the same token [,𝒦0,i]𝒦0,i[\ell^{\prime},{\cal K}_{0,i}]\subseteq{\cal K}_{0,i}. By construction, fb¯=(ba)fb¯0\ell^{\prime}\cdot\overline{f_{b}}=(b-a)\overline{f_{b}}\neq 0. \Box

The graded case also allows us to analyse the tangent of exit.

Definition 4.5

Let λ(t)\lambda(t) be a 1-PS acting on the form ff and let:

f(t)=λ(t)f=tag+tbfb++tDfDf(t)=\lambda(t)\cdot f=t^{a}g+t^{b}f_{b}+\ldots+t^{D}f_{D}

Then the tangent of exit is the form limt1f(t)f(1)t1\lim_{t\rightarrow 1}\frac{f(t)-f(1)}{t-1}.

Since

f=afa+bfb++DfD\ell f=af_{a}+bf_{b}+\ldots+Df_{D}

the tangent of exit is ff\ell f-f. Note that both ff and f\ell f are elements of TOfTO_{f}, the tangent space of the orbit O(f)O(f) at the point ff. Thus, the tangent of exit is given (upto addition of a suitable multiple of the form ff) by the action f\ell f, where 𝒦\ell\not\in{\cal K}.

An important question is the stabilizer of f\ell f within TOfTO_{f} under the action of 𝒦{\cal K}, the stabilizer of ff. This is answered by the following lemma:

Lemma 4.6

Πi:𝒢𝒢i\Pi_{i}:{\cal G}\rightarrow{\cal G}_{i} be the weight projections as per λ\lambda. Then the stabilizer 𝒦f𝒦{\cal K}_{\ell f}\subseteq{\cal K} within the stabilizer of ff consists of precisely those elements 𝔨𝒦\mathfrak{k}\in{\cal K} such that [𝔨,]𝒦[\mathfrak{k},\ell]\in{\cal K}. In particular, it contains all elements of 𝒦{\cal K} which are pure in weight, i.e., elements 𝔨\mathfrak{k} such that Πi(𝔨)=i𝔨\Pi_{i}(\mathfrak{k})=i\cdot\mathfrak{k} for some ii. Thus, these elements 𝔨𝒦\mathfrak{k}\in{\cal K} of pure weight are triple stabilizers, i.e., members of 𝒦𝒦f{\cal K}\cap{\cal H}\cap{\cal K}_{\ell f}.

Proof: Let 𝔨𝒦\mathfrak{k}\in{\cal K} stabilize f\ell f. Then, since 𝔨f=0\mathfrak{k}f=0, we have 𝔨f=[𝔨,]f+𝔨f=[𝔨,]f\mathfrak{k}\ell f=[\mathfrak{k},\ell]f+\ell\mathfrak{k}f=[\mathfrak{k},\ell]f. This, in turn, is equivalent to [𝔨,]𝒦[\mathfrak{k},\ell]\in{\cal K}. This proves the first assertion. For the second, note that if 𝔨𝒦\mathfrak{k}\in{\cal K} is of pure weight, then [𝔨,]=i𝔨[\mathfrak{k},\ell]=i\cdot\mathfrak{k} is certainly an element of 𝒦{\cal K}. The last assertion comes from the fact that pure weight elements are also elements of 𝒦0{\cal K}_{0}\subseteq{\cal H}. Note \Box

Lemma 4.7

Pure elements 𝔨𝒦\mathfrak{k}\in{\cal K} stabilize all components in the decomposition:

f=fa+fb++fDf=f_{a}+f_{b}+\ldots+f_{D}

Proof: Let [𝔨,]=r𝔨[\mathfrak{k},\ell]=r\mathfrak{k}. We claim that 𝔨kf=0\mathfrak{k}\ell^{k}f=0 for all k0k\geq 0, which we prove by induction on kk. The assertion for k=0k=0 is obvious. Then, for general kk, we have:

0=𝔨kf=𝔨k1f+r𝔨k1f=0+0\begin{array}[]{rcl}0=\mathfrak{k}\ell^{k}f&=&\ell\mathfrak{k}\ell^{k-1}f+r\mathfrak{k}\ell^{k-1}f\\ &=&0+0\end{array}

This proves the assertion. Next, we apply 𝔨k\mathfrak{k}\ell^{k} to the above equation to get:

0=𝔨(kf)=𝔨(kfa)+𝔨(kfb)++𝔨(kfD)=𝔨(akfa+bkfb++DkfD)=ak(𝔨fa)+bk(𝔨fb)++Dk(𝔨fD)\begin{array}[]{rcl}0=\mathfrak{k}(\ell^{k}f)&=&\mathfrak{k}(\ell^{k}f_{a})+\mathfrak{k}(\ell^{k}f_{b})+\ldots+\mathfrak{k}(\ell^{k}f_{D})\\ &=&\mathfrak{k}(a^{k}f_{a}+b^{k}f_{b}+\ldots+D^{k}f_{D})\\ &=&a^{k}(\mathfrak{k}f_{a})+b^{k}(\mathfrak{k}f_{b})+\ldots+D^{k}(\mathfrak{k}f_{D})\end{array}

Since all of these are zero, we must have 𝔨fc=0\mathfrak{k}f_{c}=0 for all cc. \Box

Thus, pure elements stabilize limits, tangents of approach and exit and all components in between. That there is an element 𝔨𝒦\mathfrak{k}\in{\cal K} which is pure with respect \ell is a significant alignment between 𝒦{\cal K} and \ell.

Remark 4.8

The above condition also connects with the weight structure of 𝒦0{\cal K}_{0}. Let 𝔨𝒦\mathfrak{k}\in{\cal K} be such that its weight space decomposition is 𝔨=𝔨1+𝔨0+𝔨1\mathfrak{k}=\mathfrak{k}^{-1}+\mathfrak{k}^{0}+\mathfrak{k}^{1}. We define

𝒦0={𝔨𝒦|𝔨1=0} and 𝒦1={𝔨𝒦|𝔨0=0,𝔨1=0}{\cal K}^{\geq 0}=\{\mathfrak{k}\in{\cal K}|\mathfrak{k}^{-1}=0\}\mbox{ and }{\cal K}^{\geq 1}=\{\mathfrak{k}\in{\cal K}|\mathfrak{k}^{0}=0,\mathfrak{k}^{-1}=0\}

Note that 𝒦1𝒦0𝒦(1)=𝒦{\cal K}^{\geq 1}\subseteq{\cal K}^{\geq 0}\subseteq{\cal K}^{\geq(-1)}={\cal K} is a nested sequence of subalgebras such that 𝒦i/𝒦(i+1)(𝒦0)i{\cal K}^{\geq i}/{\cal K}^{\geq(i+1)}\cong({\cal K}_{0})_{i}. Since (𝒦0)ii({\cal K}_{0})_{i}\subseteq{\cal H}_{i}, we have dim((𝒦0)i)dim(i)dim(({\cal K}_{0})_{i})\leq dim({\cal H}_{i}).

In Section 5.2 we work out this decomposition for various 1-PS acting on the form det3det_{3}, the 3×33\times 3-determinant.

Proposition 4.9

Let f,g,λf,g,\lambda and \ell be as above. Let P(λ)P(\lambda) be defined as below:

P(λ)={gG|limt0λ(t)gλ(t)1 exists}P(\lambda)=\{g\in G|\lim_{t\rightarrow 0}\lambda(t)g\lambda(t)^{-1}\mbox{ exists}\}

and let U(λ)U(\lambda) be its unipotent radical. Suppose that 𝔨𝒫(λ)𝒦\mathfrak{k}\in{\cal P}(\lambda)\cap{\cal K} is a semi-simple element, then there is a unipotent element uU(λ)u\in U(\lambda) such that:

  1. 1.

    If fu=uff^{u}=u\cdot f, then:

    λ(t)fu=tag+ higher terms \lambda(t)f^{u}=t^{a}g+\mbox{ higher terms } (3)

    Thus, gg is the limit of fuf^{u} under λ\lambda.

  2. 2.

    The element 𝔨u=u𝔨u1\mathfrak{k}^{u}=u\mathfrak{k}u^{-1} stabilizes fuf^{u}, gg and fu\ell f^{u}. Thus, 𝔨u\mathfrak{k}^{u} is a triple stabilizer of the limit gg, fuf^{u} and the tangent of exit fu\ell f^{u}.

Proof: Let L(λ)L(\lambda) be the elements of GL(X)GL(X) which commute with λ\lambda. Then P(λ)=L(λ)U(λ)=U(λ)L(λ)P(\lambda)=L(\lambda)U(\lambda)=U(\lambda)L(\lambda) is a Levi factorization, with L(λ)L(\lambda) as a reductive complement. Applying this to the corresponding Lie algebras, for any semisimple 𝔨\mathfrak{k} as above, there is a uU(λ)u\in U(\lambda) so that 𝔨u=u𝔨u1(λ)\mathfrak{k}^{u}=u\mathfrak{k}u^{-1}\in{\cal L}(\lambda) and thus is an element of pure weight. Moreover 𝔨u\mathfrak{k}^{u} stabilizes fu=uff^{u}=u\cdot f. Since uU(λ)u\in U(\lambda) is unipotent, for any hcVch_{c}\in V_{c}, the element uhu\cdot h has the following weight decomposition with respect to λ(t)\lambda(t):

uhc=hc+hc+1+higher termsuh_{c}=h_{c}+h^{\prime}_{c+1}+\mbox{higher terms}

Now since ff has the weight decomposition:

f=ga+fb+ higher terms f=g_{a}+f_{b}+\mbox{ higher terms } (4)

on applying uu to Eq. 4 we have:

uf=ug+ufb+higher order terms=g+fa+1+ higher terms\begin{array}[]{rcl}uf&=&u\cdot g+u\cdot f_{b}+\mbox{higher order terms}\\ &=&g+f^{\prime}_{a+1}+\mbox{ higher terms}\\ \end{array}

Thus, gg continues to be the leading term of fuf^{u}. Whence, we have:

λ(t)(uf)=tag+ta+1fa+1+higher terms\lambda(t)(uf)=t^{a}g+t^{a+1}f^{\prime}_{a+1}+\mbox{higher terms}

Thus for the data fu,λ,f^{u},\lambda,\ell and gg, we have that gg is the limit of fuf^{u} under the action of λ\lambda. Now note that 𝔨u𝒦fu\mathfrak{k}^{u}\in{\cal K}_{f^{u}} and is of pure weight. Thus, by lemma 4.6, 𝔨u\mathfrak{k}^{u} stabilizes the tangent of exit fu\ell f^{u} and gg. \Box

Proposition 4.10

Let f,g,λf,g,\lambda and \ell be as above. Then at least one of the following hold:

  1. (A)

    𝒦0{\cal K}_{0} is a nilpotent algebra, or

  2. (B)

    there is a unipotent element uU(λ)u\in U(\lambda) and an element 𝔨𝒦\mathfrak{k}\in{\cal K} such that gg is a limit of fuf^{u} under λ\lambda and 𝔨u\mathfrak{k}^{u} is a triple stabilizer for the data (fu,fu,g)(f^{u},\ell f^{u},g).

Proof: Note that 𝒢0𝒢1=𝒫(λ){\cal G}_{0}\oplus{\cal G}_{1}={\cal P}(\lambda) and 𝒢1=𝒰(λ){\cal G}_{1}={\cal U}(\lambda), the Lie algebras of P(λ)P(\lambda) and U(λ)U(\lambda), and that the above decomposition is a Levi decomposition of 𝒫(λ){\cal P}(\lambda) with (λ)=𝒢0{\cal L}(\lambda)={\cal G}_{0} as a semisimple complement.

Case 1: Suppose that dim(Π1(𝒦))=dim(𝒦)dim(\Pi_{-1}({\cal K}))=dim({\cal K}), i.e., for every 𝔨𝒦\mathfrak{k}\in{\cal K}, if 𝔨0\mathfrak{k}\neq 0 then so is Π1(𝔨)\Pi_{-1}(\mathfrak{k}). If this happens, then 𝒦0𝒢1{\cal K}_{0}\subseteq{\cal G}_{-1} and 𝒦0{\cal K}_{0} is nilpotent. Thus (A) holds.

Case 2: On the other hand, if dim(Π1(𝒦))<dim(𝒦)dim(\Pi_{-1}({\cal K}))<dim({\cal K}), then either:

  1. 2a

    There is an element 𝔨𝒰(λ)𝒦\mathfrak{k}\in{\cal U}(\lambda)\cap{\cal K}. In this case, 𝔨\mathfrak{k} is a pure element and by Lemma 4.6, 𝔨\mathfrak{k} stabilizes f,gf,g and f\ell f. Thus (B) holds with u=1u=1.

  2. 2b

    All elements 𝔨𝒫(λ)\mathfrak{k}\in{\cal P}(\lambda) may be written as 𝔨=𝔨0+𝔨1\mathfrak{k}=\mathfrak{k}_{0}+\mathfrak{k}_{1} with 𝔨00\mathfrak{k}_{0}\neq 0. Let us use Jordan’s decomposition to the element 𝔨\mathfrak{k} to express 𝔨=𝔨ss+𝔨n\mathfrak{k}=\mathfrak{k}_{ss}+\mathfrak{k}_{n}, where 𝔨ss\mathfrak{k}_{ss} is semisimple and 𝔨n\mathfrak{k}_{n} is nilpotent. Both are elements of 𝒫(λ)𝒦{\cal P}(\lambda)\cap{\cal K}. If there is a 𝔨ss0\mathfrak{k}_{ss}\neq 0 as above, then by Prop. 4.9, (B) holds.

  3. 2c

    We come to the final case where for all 𝔨𝒫(λ)𝒦\mathfrak{k}\in{\cal P}(\lambda)\cap{\cal K}, we have 𝔨=𝔨n=𝔨0+𝔨1\mathfrak{k}=\mathfrak{k}_{n}=\mathfrak{k}_{0}+\mathfrak{k}_{1} with 𝔨00\mathfrak{k}_{0}\neq 0. Since 𝔨nk=0\mathfrak{k}_{n}^{k}=0 for some kk, we have 𝔨0k=0\mathfrak{k}_{0}^{k}=0 as well and thus 𝔨0\mathfrak{k}_{0} is nilpotent. Now, note that 𝔨0\mathfrak{k}_{0} is the leading term of 𝔨\mathfrak{k} and thus 𝔨0𝒦0\mathfrak{k}_{0}\in{\cal K}_{0}. Whence we have 𝒦0=(𝒦0)0+(𝒦0)1{\cal K}_{0}=({\cal K}_{0})_{0}+({\cal K}_{0})_{-1} where the general element 𝔨𝒦0\mathfrak{k}^{\prime}\in{\cal K}_{0} may be written as 𝔨0+𝔨1\mathfrak{k}^{\prime}_{0}+\mathfrak{k}^{\prime}_{-1}, where both are individually nilpotent. The grading ensures that 𝔨\mathfrak{k}^{\prime} is nilpotent too. Thus 𝒦0{\cal K}_{0} is nilpotent, and (A) holds.

\Box

Remark 4.11

The condition that 𝒦0𝒢1{\cal K}_{0}\not\subseteq{\cal G}_{1} is tantamount to saying that dim(Π1(𝒦))<dim(𝒦)dim(\Pi_{-1}({\cal K}))<dim({\cal K}), i.e., 𝒦{\cal K} is not generically placed with respect to \ell. If indeed that is so, then all leading terms of 𝒦{\cal K} have weight 1-1 and 𝒦0{\cal K}_{0} is nilpotent.

4.2 General 1-PS

We extend the results of the previous section to a general 1-PS λ\lambda, beyond the two-component case considered earlier. Thus, let X=X1XkX=X_{1}\cup\ldots\cup X_{k}, with |Xi|=ni|X_{i}|=n_{i}, and let λ(t)GL(X)=G\lambda(t)\in GL(X)=G be defined so that λ(t)(xi)=tdixi\lambda(t)(x_{i})=t^{d_{i}}x_{i}. We call d¯=(d1n1,,dknk)\overline{d}=(d_{1}^{n_{1}},\ldots,d_{k}^{n_{k}}) as the type of λ\lambda. Thus, in the chosen basis of (xi)=X(x_{i})=X, λ(t)\lambda(t) is the diagonal matrix diag(td¯)diag(t^{\overline{d}}). As before, VV is a GG-module via the map ρ:GL(X)GL(V)\rho:GL(X)\rightarrow GL(V). We also have the Lie algebra map (also) ρ:gl(X)gl(V)\rho:gl(X)\rightarrow gl(V). We now have V=χVχV=\oplus_{\chi}V^{\chi}, the weight space decomposition of VV. Note that χ\chi\in\mathbb{Z}. Let =logt(λ(t))\ell=\log_{t}(\lambda(t)) as before. Then for vχVχv^{\chi}\in V^{\chi}, we have ρ()vχ=χvχ\rho(\ell)\cdot v^{\chi}=\chi\cdot v^{\chi}. We also have 𝒢=χ𝒢χ{\cal G}=\oplus_{\chi}{\cal G}^{\chi}.

Definition 4.12

For any 𝒢{\cal G}-module MM, we say that mMm\in M is pure if m=χm\ell\cdot m=\chi\cdot m for some χ\chi\in\mathbb{Z}. We say that MM is graded if M=χMχM=\oplus_{\chi}M^{\chi}, the sum of its pure elements.

Proposition 4.13

For the above situation, we have the following:

  1. 1.

    The stabilizer 𝒢{\cal H}\subseteq{\cal G} of xx is graded.

  2. 2.

    Let 𝒦{\cal K} be the stabilizer of yy, and 𝒦0{\cal K}_{0} be the limit of 𝒦{\cal K} under the action of λ\lambda, then 𝒦0{\cal K}_{0} is graded.

  3. 3.

    The tangent space TyO(y)T_{y}O(y) of yy as an element of its GG-orbit O(y)O(y) is a 𝒦{\cal K}-module. Let 𝒦y𝒦{\cal K}_{\ell y}\subseteq{\cal K} be the stabilizer of the tangent vector y\ell y. Then 𝔨𝒦\mathfrak{k}\in{\cal K} stabilizes y\ell y if and only if [,𝔨]𝒦[\ell,\mathfrak{k}]\in{\cal K}. Consequently, if 𝔨𝒦\mathfrak{k}\in{\cal K} is of pure weight, then 𝔨𝒦y\mathfrak{k}\in{\cal K}_{\ell y} as well.

Proof: The arguments of Section 4.1 are easily modified for the general case. \Box

Prop. 4.9 holds in the general case as well and Prop. 4.14 below is a restatement in the notation of this section. However, Prop. 4.10, on the structure of 𝒦0{\cal K}_{0} does not; its modification is Prop. 4.15.

Proposition 4.14

Let y,x,λy,x,\lambda and \ell be as above. Let P(λ)P(\lambda) be defined as below:

P(λ)={gG|limt0λ(t)gλ(t)1 exists}P(\lambda)=\{g\in G|\lim_{t\rightarrow 0}\lambda(t)g\lambda(t)^{-1}\mbox{ exists}\}

and let U(λ)U(\lambda) be its unipotent radical. Suppose that 𝔨𝒫(λ)𝒦\mathfrak{k}\in{\cal P}(\lambda)\cap{\cal K} is a semi-simple element, then there is a unipotent element uU(λ)u\in U(\lambda) such that:

  1. 1.

    If yu=uyy^{u}=u\cdot y, then:

    λ(t)yu=tax+ higher terms \lambda(t)y^{u}=t^{a}x+\mbox{ higher terms } (5)

    Thus, there is a yuy^{u} in the orbit of yy such that xx is the limit of yuy^{u} under λ\lambda.

  2. 2.

    The element 𝔨u=u𝔨u1\mathfrak{k}^{u}=u\mathfrak{k}u^{-1} stabilizes yuy^{u}, xx and yu\ell y^{u}. Thus, 𝔨u\mathfrak{k}^{u} is a triple stabilizer of the limit xx, yuy^{u} and the tangent of exit yu\ell y^{u}.

The proof of this is a straightforward adaptation of the proof of Prop. 4.9.

Proposition 4.15

Let y,x,λy,x,\lambda and \ell be as above. Then at least one of the following holds:

  1. (A)

    Let 𝒦=𝒦0{\cal K}^{\prime}={\cal K}_{0}\oplus\mathbb{C}\ell, then 𝒦{\cal K}^{\prime} is a Lie algebra of rank 1, i.e., the dimension of any maximal torus in 𝒦{\cal K}^{\prime} is 11, [,𝒦0]𝒦0[\ell,{\cal K}_{0}]\subseteq{\cal K}_{0}, and every graded component (𝒦0)i({\cal K}_{0})_{i} is composed of nilpotent elements or

  2. (B)

    there is a unipotent element uU(λ)u\in U(\lambda) and an element 𝔨𝒦\mathfrak{k}\in{\cal K} such that xx is a limit of yuy^{u} under λ\lambda and 𝔨u\mathfrak{k}^{u} is a triple stabilizer for the data (yu,yu,x)(y^{u},\ell y^{u},x).

Proof: Note that 𝒫(λ)=χ0𝒢χ{\cal P}(\lambda)=\oplus_{\chi\geq 0}\>{\cal G}_{\chi} and 𝒰(λ)=χ>0𝒢χ{\cal U}(\lambda)=\oplus_{\chi>0}\>{\cal G}_{\chi}, the Lie algebras of P(λ)P(\lambda) and U(λ)U(\lambda), and that the above decomposition is a Levi decomposition of 𝒫(λ){\cal P}(\lambda) with 𝒢0{\cal G}_{0} as a semisimple complement. Next note that 𝒦=𝒦0{\cal K}^{\prime}={\cal K}_{0}\oplus\mathbb{C}\ell is indeed a Lie algebra since [,𝒦0]𝒦0[\ell,{\cal K}_{0}]\subseteq{\cal K}_{0}. Now we do a case analysis.

Case 1: Let Π:𝒢χ<0𝒢χ\Pi_{-}:{\cal G}\rightarrow\oplus_{\chi<0}\>{\cal G}_{\chi} be the projection. Suppose that dim(Π(𝒦))=dim(𝒦)dim(\Pi_{-}({\cal K}))=dim({\cal K}). If this happens, then 𝒦0χ<0𝒢χ{\cal K}_{0}\subseteq\oplus_{\chi<0}{\cal G}_{\chi} and 𝒦0{\cal K}_{0} is nilpotent. Even more, 𝒦=𝒦0{\cal K}^{\prime}={\cal K}_{0}\oplus\mathbb{C}\ell is a Levi decomposition of 𝒦{\cal K}^{\prime}. Thus (A) holds.

Case 2: On the other hand, if dim(Π1(𝒦))<dim(𝒦)dim(\Pi_{-1}({\cal K}))<dim({\cal K}), then 𝒦𝒫(λ)0{\cal K}\cap{\cal P}(\lambda)\neq 0.

  1. 2a

    If there is an 𝒫(λ)𝒦\ell^{\prime}\in{\cal P}(\lambda)\cap{\cal K} such that [,]=0[\ell^{\prime},\ell]=0, then it must be that 𝒢0\ell^{\prime}\in{\cal G}_{0}, i.e., it must be of pure weight and therefore a triple stabilizer, and therefore (B) holds.

  2. 2b

    If there is a semi-simple element 𝔨𝒫(λ)𝒦\mathfrak{k}\in{\cal P}(\lambda)\cap{\cal K} then, by Prop. 4.15, there is a uu and a triple stabilizer as required. Thus (B) holds.

  3. 2d

    We come to the final case where all 𝔨𝒫(λ)𝒦\mathfrak{k}\in{\cal P}(\lambda)\cap{\cal K} are nilpotent. Let 𝔨χ0\mathfrak{k}_{\chi_{0}} be the leading term of 𝔨=χ0𝔨χ\mathfrak{k}=\sum_{\chi\leq 0}\mathfrak{k}_{\chi}, i.e., the smallest degree dd such that 𝔨d0\mathfrak{k}_{d}\neq 0. If χ0>0\chi_{0}>0 then 𝔨χ0𝒦0\mathfrak{k}_{\chi_{0}}\in{\cal K}_{0} is nilpotent. Otherwise, we may write 𝔨=𝔨0+χ>0𝔨χ\mathfrak{k}=\mathfrak{k}_{0}+\sum_{\chi>0}\mathfrak{k}_{\chi}. Since 𝔨k=0\mathfrak{k}^{k}=0 for some kk, we have 𝔨0k=0\mathfrak{k}_{0}^{k}=0 as well and thus 𝔨0\mathfrak{k}_{0} is nilpotent. Thus, for every χ\chi, 𝒦0𝒢χ{\cal K}_{0}\cap{\cal G}_{\chi} consists of nilpotent elements. From Case 2a above, we already know that the rank of 𝒦=𝒦{\cal K}^{\prime}={\cal K}\oplus\mathbb{C}\ell is 11

This proves the proposition. \Box

5 Closures of affine orbits

We apply the results of the last two sections to the closure of forms whose SL(X)SL(X)-orbit is affine. As mentioned in the introduction (see also Appendix A), the motivation for studying comes from geometric complexity theory. Examples of such forms are det(X)det(X), the determinant of the matrix XX, and perm(X)perm(X), its permanent. In the next Section, we cover some general results. In Section 5.2 we apply these and earlier results to the 3×33\times 3 determinant form.

5.1 General Results

Let f(X)Symd(X)f(X)\in Sym^{d}(X) be such that the stabilizer KGL(X)K\subseteq GL(X) is reductive (of dimension kk). This implies that the SL(X)SL(X)-orbit of ff is an affine variety. Let O(f)O(f) be the GL(X)GL(X)-orbit of ff and O(f)¯\overline{O(f)} be its projective closure. It is known that the closed set O(f)¯O(f)\overline{O(f)}-O(f) is of co-dimension one. Moreover, there are forms g1,,gmg_{1},\ldots,g_{m} such that it is contained in the union of the projective closures {O(gi)¯}i=1m\{\overline{O(g_{i})}\}_{i=1}^{m}.

Separately, it is known that for any form ff, forms gg which appear in the closure of the orbit O(f)O(f) appear as leading terms under a 1-parameter substitutions A(t)XA(t)\cdot X, where A(t)GL(X)A(t)\in GL(X) has entries which are polynomials in tt. Thus, for the above gig_{i}, there are linear transformations Ai(t)A_{i}(t), with polynomial entries, such that f(Ai(t)X)=tdgi+f(A_{i}(t)\cdot X)=t^{d}g_{i}+ higher degree terms. It is also clear from dimension counting that for all ii, the stabilizer HiH_{i} of gig_{i} is of dimension k+1k+1.

Our focus will be on one such those gi=gg_{i}=g. We begin with a lemma:

Lemma 5.1

Given a transformation A(t)gl(X)A(t)\in gl(X) such that gg is the leading term of f(t)=f(A(t)X)f(t)=f(A(t)\cdot X), we may assume that A(t)GL(X)A(t)\in GL(X), i.e., A(t)A(t) is invertible for most tt\in\mathbb{C}.

Proof: Let us consider A(t)=A(t)+tDIA^{\prime}(t)=A(t)+t^{D}I for a large power D>0D>0. Then gg continues to be the leading term of f(t)=f(A(t)X)f^{\prime}(t)=f(A^{\prime}(t)\cdot X). Moreover, A(t)A^{\prime}(t) is invertible for most tt\in\mathbb{C}. \Box

By the above lemma, we may assume that A(t)A(t) is generically invertible and that:

f(t)=f(A(t)X)=tag+tbfb++tdfdf(t)=f(A(t)\cdot X)=t^{a}g+t^{b}f_{b}+\ldots+t^{d}f_{d}

Let {\cal H} be the Lie algebra stabilizer of gg as before and 𝒦(t)=A(t)𝒦A(t)1{\cal K}(t)=A(t){\cal K}A(t)^{-1} that of f(t)f(t), with dim(𝒦(t))=kdim({\cal K}(t))=k. Let 𝒦0{\cal K}_{0} be the limit of 𝒦(t){\cal K}(t) and note that dim(𝒦0)=kdim({\cal K}_{0})=k and that 𝒦0b{\cal K}_{0}\subseteq{\cal H}_{b}, the stabilizer for \star-action of {\cal H} on N¯\overline{N}. We have the following lemma:

Lemma 5.2

The algebra 𝒦0{\cal K}_{0} is precisely b{\cal H}_{b}, the stabilizer of fb¯N¯\overline{f_{b}}\in\overline{N} under the \star-action of {\cal H}.

Proof: We have the containment 𝒦0b{\cal K}_{0}\subseteq{\cal H}_{b}\subset{\cal H}. Next, the orbit of gg is of co-dimension 1 within the orbit of f(t)f(t). This, gives us that dim()=dim(𝒦(t))1=k+1dim({\cal H})=dim({\cal K}(t))-1=k+1. Since b{\cal H}_{b}\neq{\cal H}, we must have b=𝒦0{\cal H}_{b}={\cal K}_{0}. \Box

Lemma 5.3

Suppose further that A(t)A(t) is a one-parameter subgroup λ(t)\lambda(t). So we have

λ(t)f=f(t)=tag+tbfb+higher terms\lambda(t)\cdot f=f(t)=t^{a}\cdot g+t^{b}\cdot f_{b}+\mbox{higher terms}

Let =log(ta/dλ(t))gl(X)\ell=\log(t^{-a/d}\lambda(t))\in gl(X) be the Lie algebra element generating the 1-PS taλ(t)t^{-a}\lambda(t). Suppose that fb¯0\overline{f_{b}}\neq 0. Then (i) \ell\in{\cal H} but does not stabilize fb¯\overline{f_{b}}, (ii) 𝒦0=b{\cal K}_{0}={\cal H}_{b}, and (iii) =𝒦0{\cal H}={\cal K}_{0}\oplus\mathbb{C}\cdot\ell.

Proof: Since taλ(t)g=gt^{-a}\lambda(t)\cdot g=g but taλ(t)fb=tbafbt^{-a}\lambda(t)f_{b}=t^{b-a}f_{b}, (i) is clear. Thus 𝒦0\ell\in{\cal H}-{\cal K}_{0}. Since the dimension of 𝒦0{\cal K}_{0} is one less than the dimension of {\cal H}, we must have =𝒦0{\cal H}={\cal K}_{0}\oplus\ell. This proves the lemma. \Box

When gg is obtained as the limit of λ(t)\lambda(t) where \ell has only two distinct entries (i.e., the 2-block case), we have the following result:

Proposition 5.4

Let gg be a projective limit of codimension 11 of ff under λ\lambda as above. Then at least one of the following conditions hold:

  1. 1.

    𝒦0{\cal K}_{0} is nilpotent and =𝒦0{\cal H}={\cal K}_{0}\oplus\mathbb{C}\ell is a Levi factorization of {\cal H}.

  2. 2.

    There is a unipotent element uU(λ)u\in U(\lambda) (see 4.10) and an element 𝔨𝒦\mathfrak{k}\in{\cal K} such that u𝔨u1u\mathfrak{k}u^{-1} is a stabilizer of g,fug,f^{u} and fu\ell f^{u}, and gg is the limit of fuf^{u} under λ\lambda.

Proof: This is a consequence of Prop 4.10 and the fact that =𝒦0{\cal H}={\cal K}_{0}\oplus\mathbb{C}\ell. \Box

5.2 The analysis of det3det_{3}, the 3×33\times 3 - determinant.

We consider here the set X={x1,,x9}X=\{x_{1},\ldots,x_{9}\} and the 165165-dimensional space Sym3(X)Sym^{3}(X) of homogeneous forms of degree 33, acted upon by GL(X)GL(X). The 3×33\times 3-determinant, det3(X)det_{3}(X) is a special element and its definition is given below:

det3(X)=det([x1x2x3x4x5x6x7x8x9])det_{3}(X)=det\left(\left[\begin{array}[]{ccc}x_{1}&x_{2}&x_{3}\\ x_{4}&x_{5}&x_{6}\\ x_{7}&x_{8}&x_{9}\\ \end{array}\right]\right)

The obvious question are the codimension one forms in the orbit closure of the orbit of det3(X)det_{3}(X).

These have been identified by Hüttenhain[Hüt17] and are given by the two forms Q1Q_{1} and Q2Q_{2} below:

Q1(X)=det([x1x2x3x4x5x6x7x8x5x1])Q2(X)=x4x12+x5x22+x6x32+x7x1x2+x8x2x3+x9x1x3\begin{array}[]{rcl}Q_{1}(X)&=&det\left(\left[\begin{array}[]{ccc}x_{1}&x_{2}&x_{3}\\ x_{4}&x_{5}&x_{6}\\ x_{7}&x_{8}&-x_{5}-x_{1}\\ \end{array}\right]\right)\\ Q_{2}(X)&=&x_{4}x_{1}^{2}+x_{5}x_{2}^{2}+x_{6}x_{3}^{2}+x_{7}x_{1}x_{2}+x_{8}x_{2}x_{3}+x_{9}x_{1}x_{3}\end{array}

We will study each in turn. Before we begin, let us look at the stabilizer 𝒦{\cal K} of det3(X)det_{3}(X). This will serve as our form ff, in the notation of this section. Within Sym3(X)Sym^{3}(X) the dimension of the GL(X)GL(X)-orbit O(det3(X))O(det_{3}(X)) of det3(X)det_{3}(X) is given by dim(GL(X))dim(𝒦))dim(GL(X))-dim({\cal K})). It is known that this is composed of the transformations XAXBX\rightarrow AXB with det(AB)=1det(AB)=1 and the XXTX\rightarrow X^{T}. The condition det(AB)=1det(AB)=1 and that (cA)X(c1B)(cA)X(c^{-1}B) lead to the same linear transformation on XX, give us two independent constraints on GL3×GL3𝒦GL_{3}\times GL_{3}\rightarrow{\cal K}. This gives us that dim(𝒦(1))=182=16dim({\cal K}(1))=18-2=16 and dim(O(det3(X))=65dim(O(det_{3}(X))=65.

Coming to Q1Q_{1}, we see that:

det3(X)=Q1(X)+(x1+x5+x9)(x1x5x2x4)det_{3}(X)=Q_{1}(X)+(x_{1}+x_{5}+x_{9})(x_{1}x_{5}-x_{2}x_{4})

We denote by Q1Q_{1}^{\prime} the form (x1+x5+x9)(x1x5x2x4)(x_{1}+x_{5}+x_{9})(x_{1}x_{5}-x_{2}x_{4}). This motivates us to define Y={x1,,x8}Y=\{x_{1},\ldots,x_{8}\} and Z={x1+x5+x9}Z=\{x_{1}+x_{5}+x_{9}\}. We define λ1(t)GL(X)\lambda_{1}(t)\in GL(X) as λ1(t)xi=xi\lambda_{1}(t)x_{i}=x_{i} for i=1,,8i=1,\ldots,8 and λ(t)(z)=tz\lambda(t)(z)=tz, where z=(x1+x5+x9)z=(x_{1}+x_{5}+x_{9}). We thus see that:

λ1(t)det3(X)=Q1+tQ1\lambda_{1}(t)\cdot det_{3}(X)=Q_{1}+t\cdot Q_{1}^{\prime}

Thus in the notation of this section, f=det3(X)f=det_{3}(X), g=Q1g=Q_{1} and fb=Q1f_{b}=Q_{1}^{\prime}, with a=0a=0 and b=1b=1.

Lemma 5.5

The stabilizer of 1{\cal H}_{1} of Q1Q_{1} within gl(X)gl(X) has dimension 17. Moreover, the stabilizer 1{\cal H}_{1} has a direct sum decomposition into degree 0 and degree 1 parts, as in Section 4. The tangent of exit is the form Q1Q^{\prime}_{1} and the dimension of its stabilizer within 𝒦{\cal K}, the stabilizer of det3(X)det^{3}(X), is 44.

Proof: Let us identify YY as the space of all trace zero matrices yy labelled with their 9 entries minus the entry y(3,3)y(3,3). Then, Q1Q_{1} is precisely the determinant of yYy\in Y. For any element AGL3A\in GL_{3}, the action of AA on yy given by yAyA1y\rightarrow AyA^{-1} preserves the space YY of trace zero matrices as well as Q1Q_{1}. Whence, we get a map GL31GL_{3}\rightarrow{\cal H}_{1}, the stabilizer of Q1Q_{1}. Since this map has a 1-dimensional kernel (viz., the matrices cIcI, with c0c\neq 0), we have an 88-dimensional image Y1{\cal R}_{Y}\subseteq{\cal H}_{1}. Next, let Hom(Z,YZ)Hom(Z,Y\oplus Z) be the collection of all linear substitutions for zz in terms of the variables of YYand zz. Since Q1Q_{1} does not involve zz, these substitutions leave Q1Q_{1} invariant. This gives us a map Hom(Z,Y)Hom(Z,Z)1Hom(Z,Y)\oplus Hom(Z,Z)\rightarrow{\cal H}_{1}. We call this image as the spaces 𝒬1Z{\cal Q}_{1}\oplus{\cal R}_{Z}. It can be shown that 𝒬1,Y{\cal Q}_{1},{\cal R}_{Y} and Z{\cal R}_{Z} intersect trivially and thus YZ𝒬11{\cal R}_{Y}\oplus{\cal R}_{Z}\oplus{\cal Q}_{1}\subseteq{\cal H}_{1} to obtain a 1717-dimensional subalgebra. The result of [Hüt17] implies that the dimension can not be more, since we know that the dimension of the orbit of Q1Q_{1} is only one less than that of det3(X)det_{3}(X), and hence its stabilizer dimension must be 1717.

We then see that 1=YZ𝒬1=𝒬1{\cal H}_{1}={\cal R}_{Y}\oplus{\cal R}_{Z}\oplus{\cal Q}_{1}={\cal R}\oplus{\cal Q}_{1} and splits by degree, i.e., (1)0=({\cal H}_{1})_{0}={\cal R} and (1)1=𝒬1({\cal H}_{1})_{-1}={\cal Q}_{1}.

The stabilizer dimension of Q1Q^{\prime}_{1} is through a direct computation. \Box

Lemma 5.6

We have Hb=Y𝒬1H_{b}={\cal R}_{Y}\oplus{\cal Q}_{1} and 𝒦0=b{\cal K}_{0}={\cal H}_{b}. Let 1=log(λ(t))\ell_{1}=\log(\lambda(t)), then 11\ell_{1}\in{\cal H}_{1} so that 1=b1{\cal H}_{1}={\cal H}_{b}\oplus\ell_{1} and [1,b]b[\ell_{1},{\cal H}_{b}]\subseteq{\cal H}_{b}. Moreover, 1det3(X)=Q1\ell_{1}\cdot det^{3}(X)=Q^{\prime}_{1}, the tangent of exit, and b{\cal H}_{b} is an ideal of 1{\cal H}_{1}. Finally, as in Case (i), Section B with ba=1b-a=1, (i) for every element 𝔯Y\mathfrak{r}\in{\cal R}_{Y}, we have an element 𝔰\mathfrak{s} in 𝒮1{\cal S}_{1} such that 𝔯fb=𝔰Q1\mathfrak{r}\cdot f_{b}=\mathfrak{s}\cdot Q_{1}, and (ii) for any 𝔮𝒬1\mathfrak{q}\in{\cal Q}_{1}, there is an 𝔰𝒮0\mathfrak{s}\in{\cal S}_{0} such that 𝔮fb=𝔰Q1\mathfrak{q}\cdot f_{b}=\mathfrak{s}\cdot Q_{1}.

Proof: The equality of 𝒦0{\cal K}_{0} and b{\cal H}_{b} comes because Q1Q_{1} is one of the co-dimension 1 forms in the closure of det3(X)det_{3}(X). The second assertion follows from Lemma 5.3. The last set of assertions follow from the fact that bfb𝒮Q1{\cal H}_{b}\cdot f_{b}\subseteq{\cal S}\cdot Q_{1}. \Box

Remark 5.7

The statement that in this case b{\cal H}_{b} is an ideal in {\cal H} follows from a characterization of codimension 1 Lie subalgebras of a Lie algebra due to Hoffman [Hof65]. In Section 6 when we discuss connections to Lie algebra cohomology, Hoffman’s theorem is stated as Theorem 6.20.

Example 5.8

We do an example of an 𝔯\mathfrak{r} and a corresponding 𝔰\mathfrak{s} such that 𝔯fb=𝔰Q1\mathfrak{r}\cdot f_{b}=\mathfrak{s}\cdot Q_{1}. However, constructing elements 𝔯Y\mathfrak{r}\in{\cal R}_{Y} needs some care. Consider the matrix EijE_{ij} such that Eij(i,j)=1E_{ij}(i,j)=1 and Eij(k,l)=0E_{ij}(k,l)=0 for all other tuples (k,l)(k,l). Let YY be the matrix such that det(Y)=Q1det(Y)=Q_{1}, as given below:

Y=[x1x2x3x4x5x6x7x8x1x5]Y=\left[\begin{array}[]{ccc}x_{1}&x_{2}&x_{3}\\ x_{4}&x_{5}&x_{6}\\ x_{7}&x_{8}&-x_{1}-x_{5}\end{array}\right]

We construct [E23,Y]=E23YYE23[E_{23},Y]=E_{23}Y-YE_{23} as below:

[E23,Y]=R23=[0x30x7x6x8x1+x50x1x50][E_{23},Y]=R_{23}=\left[\begin{array}[]{ccc}0&x_{3}&0\\ -x_{7}&x_{6}-x_{8}&x_{1}+x_{5}\\ 0&-x_{1}-x_{5}&0\end{array}\right]

The corresponding element 𝔯\mathfrak{r} is to be constructed as the linear operator below:

𝔯=x3x2x7x4+(x6x8)x5+(x1+x5)x6(x1+x5)x8\mathfrak{r}=x_{3}\frac{\partial}{\partial x_{2}}-x_{7}\frac{\partial}{\partial x_{4}}+(x_{6}-x_{8})\frac{\partial}{\partial x_{5}}+(x_{1}+x_{5})\frac{\partial}{\partial x_{6}}-(x_{1}+x_{5})\frac{\partial}{\partial x_{8}}

We may check that 𝔯Q1=0\mathfrak{r}\cdot Q_{1}=0. On the other hand, we have:

𝔯fb=z((x2x7x1x8)(x3x4x1x6))=(zx6+zx8)Q1=𝔰Q1\mathfrak{r}\cdot f_{b}=z((x_{2}x_{7}-x_{1}x_{8})-(x_{3}x_{4}-x_{1}x_{6}))=(-z\frac{\partial}{\partial x_{6}}+z\frac{\partial}{\partial x_{8}})\cdot Q_{1}=\mathfrak{s}\cdot Q_{1}

The second operator 𝔰\mathfrak{s} is clearly an element of 𝒮1{\cal S}_{1}.

Let us now come to Q2Q_{2}.

That Q2Q_{2} is in the orbit closure of det3det_{3} is shown by the following lemma.

Lemma 5.9

[Hüt17] Let Y,ZY,Z be the generic matrices below and let X=YZX=Y\oplus Z.

Y=[0x1x2x10x3x2x30]Z=[2x6x8x9x82x5x7x9x72x4]Y=\left[\begin{array}[]{ccc}0&x_{1}&-x_{2}\\ -x_{1}&0&x_{3}\\ x_{2}&-x_{3}&0\\ \end{array}\right]\>\>\>\>Z=\left[\begin{array}[]{ccc}2x_{6}&x_{8}&x_{9}\\ x_{8}&2x_{5}&x_{7}\\ x_{9}&x_{7}&2x_{4}\\ \end{array}\right]

Let λ2(t)\lambda_{2}(t) be such that λ2(t)Y=Y\lambda_{2}(t)\cdot Y=Y and λ2(t)Z=tZ\lambda_{2}(t)\cdot Z=tZ. Let us define det3(X)det^{3}(X) as the determinant of the matrix Y+ZY+Z. Then:

det3(λ2(t)X))=det(Y+tZ)=tQ2+t3Q3det^{3}(\lambda_{2}(t)\cdot X))=det(Y+tZ)=tQ_{2}+t^{3}Q_{3}

where:

Q2(X)=x4x12+x5x22+x6x32+x7x1x2+x8x2x3+x9x1x3Q3(X)=8x4x5x62x6x722x4x822x5x92+2x7x8x9\begin{array}[]{rcl}Q_{2}(X)&=&x_{4}x_{1}^{2}+x_{5}x_{2}^{2}+x_{6}x_{3}^{2}+x_{7}x_{1}x_{2}+x_{8}x_{2}x_{3}+x_{9}x_{1}x_{3}\\ Q_{3}(X)&=&8x_{4}x_{5}x_{6}-2x_{6}x_{7}^{2}-2x_{4}x_{8}^{2}-2x_{5}x_{9}^{2}+2x_{7}x_{8}x_{9}\end{array}

The proof is a verification. Note that this parametrization of the generic matrix XX, instead of the one in the case of Q1Q_{1}, is merely a reordering of the variables and hence the expression det3(X)det^{3}(X) of the determinant is in the orbit of det3(X)det_{3}(X), and hence the orbit closures of det(X)det(X) and det3(X)det^{3}(X) match. The expansion of det3(λ2(t)X)det^{3}(\lambda_{2}(t)\cdot X) puts us into the familiar situation where Q2Q_{2} is the limit, fb=Q3f_{b}=Q_{3} with and a=1a=1 and b=3b=3.

Lemma 5.10

The stabilizer of 2{\cal H}_{2} of Q2Q_{2} within gl(X)gl(X) has dimension 17. The tangent of exit is the form Q3Q_{3} and the dimension of its stabilizer with 𝒦{\cal K}, the stabilizer of det3(X)det^{3}(X), is 88.

We may write Q2Q_{2} as the inner product:

Q2(X)=RC=[x12x22x32x1x2x2x3x1x3][x4x5x6x7x8x9]Q_{2}(X)=R\cdot C=\left[\begin{array}[]{cccccc}x_{1}^{2}&x_{2}^{2}&x_{3}^{2}&x_{1}x_{2}&x_{2}x_{3}&x_{1}x_{3}\\ \end{array}\right]\left[\begin{array}[]{c}x_{4}\\ x_{5}\\ x_{6}\\ x_{7}\\ x_{8}\\ x_{9}\\ \end{array}\right]

Let Y={x1,x2,x3}Y=\{x_{1},x_{2},x_{3}\} and Z={x4,,x9}Z=\{x_{4},\ldots,x_{9}\} and let ρ:GL(Y)GL(Sym2(Y))\rho:GL(Y)\rightarrow GL(Sym^{2}(Y)) be the representation of GL(Y)GL(Y). We see that for any AGL(Y)A\in GL(Y), we have:

Q2(X)=Rρ(A)T(ρ(A)T)1CQ_{2}(X)=R\cdot\rho(A)^{T}(\rho(A)^{T})^{-1}\cdot C

Thus, we have an embedding α:GL(Y)2GL(Y)×GL(Z)\alpha:GL(Y)\rightarrow{\cal H}_{2}\subseteq GL(Y)\times GL(Z) of GL(Y)GL(Y) into 2{\cal H}_{2}, given by AA×(ρ(A)T)1A\rightarrow A\times(\rho(A)^{T})^{-1}. The image =α(GL(Y)){\cal R}=\alpha(GL(Y)) is the reductive part of H2H_{2} and has dimension equal to dim(GL(Y))=9dim(GL(Y))=9. We may further factor {\cal R} as follows. Let Dgl(Y)D\subset gl(Y) be multiples of the identity matrix. We may then write gl(X)=Dsl(X)gl(X)=D\oplus sl(X) giving us a decomposition of =𝒟{\cal R}={\cal D}\oplus{\cal R}^{\prime}, where α(D)=𝒟\alpha(D)={\cal D} is a 11-dimensional subspace of {\cal R}, and =α(sl(X)){\cal R}^{\prime}=\alpha(sl(X)).

Let us now come to the nilpotent part which is a subspace of Hom(Z,Y)Hom(Z,Y). Let AHom(Z,Y)A\in Hom(Z,Y) be a 6×36\times 3-matrix and let us write the action of AA on Q2Q_{2} as:

AQ2(X)=RC=[x12x22x32x1x2x2x3x1x3]([x4x5x6x7x8x9]+[a11a12a13a61a62a63][x1x2x3])A\cdot Q_{2}(X)=R\cdot C=\left[\begin{array}[]{cccccc}x_{1}^{2}&x_{2}^{2}&x_{3}^{2}&x_{1}x_{2}&x_{2}x_{3}&x_{1}x_{3}\\ \end{array}\right]\left(\left[\begin{array}[]{c}x_{4}\\ x_{5}\\ x_{6}\\ x_{7}\\ x_{8}\\ x_{9}\\ \end{array}\right]+\left[\begin{array}[]{ccc}a_{11}&a_{12}&a_{13}\\ &&\\ &\vdots&\\ &&\\ &\vdots&\\ a_{61}&a_{62}&a_{63}\\ \end{array}\right]\left[\begin{array}[]{c}x_{1}\\ x_{2}\\ x_{3}\\ \end{array}\right]\right)

The requirement that AA stabilizes Q2Q_{2} is tantamount to the condition that:

p(A,Y)=RA[x1x2x3]=0p(A,Y)=R\cdot A\cdot\left[\begin{array}[]{c}x_{1}\\ x_{2}\\ x_{3}\\ \end{array}\right]=0

Since p(A,Y)Sym3(Y)p(A,Y)\in Sym^{3}(Y) and dim(Sym3(Y))=10dim(Sym^{3}(Y))=10, the condition p(A,Y)=0p(A,Y)=0 translates to 1010 equations. Thus, we define 𝒜{\cal A} as follows:

𝒜={AHom(Z,Y)|p(A,Y)=0}{\cal A}=\{A\in Hom(Z,Y)|p(A,Y)=0\}

Then we have 𝒜2{\cal A}\rightarrow{\cal H}_{2} and that dim(𝒜)=dim(Hom(Z,Y))10=8dim({\cal A})=dim(Hom(Z,Y))-10=8. Thus 2=𝒜{\cal H}_{2}={\cal R}\oplus{\cal A} and dim(2)=8+9=17dim({\cal H}_{2})=8+9=17. The stabilizer dimension of Q3Q_{3} is through a direct computation. \Box

Lemma 5.11

We have Hb=𝒜H_{b}={\cal R}^{\prime}\oplus{\cal A} and 𝒦0=b{\cal K}_{0}={\cal H}_{b}. Let 2=log(t1/3λ2(t))\ell_{2}=\log(t^{-1/3}\lambda_{2}(t)), then 22\ell_{2}\in{\cal H}_{2} so that 2=b2{\cal H}_{2}={\cal H}_{b}\oplus\ell_{2} and [2,b]b[\ell_{2},{\cal H}_{b}]\subseteq{\cal H}_{b}. Moreover, 2det3(X)=Q3\ell_{2}\cdot det^{3}(X)=Q_{3}, the tangent of exit. b{\cal H}_{b} is an ideal of 2{\cal H}_{2} and condition (3) of Theorem  6.20 holds. Finally, as in Case (ii) Section  B, with ba=2b-a=2, we have for every element 𝔞𝒜Y\mathfrak{a}\in{\cal A}_{Y}, there is an element 𝔰\mathfrak{s} in 𝒮1{\cal S}_{1} such that 𝔞fb=𝔰Q1\mathfrak{a}\cdot f_{b}=\mathfrak{s}\cdot Q_{1}.

Remark 5.12

The algebras 2{\cal H}_{2} and 𝒦0{\cal K}_{0} provide an extremely important insight into the possible structure of the Levi decomposition of 𝒦0{\cal K}_{0}. The coupling of the End(Y)End(Y) and End(Z)End(Z) components through different representations of the same algebra gl(3)gl(3), and the representation on End(Z,Y)End(Z,Y) as one which interleaves between the two, provides an important template for the Lie algebra structure of 𝒦0{\cal K}_{0}.

Finally, we cover the case of a limit of det3(X)det^{3}(X) which is not one of the components of the orbit closure. Let Y={x1,,x4}Y=\{x_{1},\ldots,x_{4}\}, Z={x5,,x9}Z=\{x_{5},\ldots,x_{9}\} so that X=YZX=Y\cup Z. Let λ3(t):GL(X)\lambda_{3}(t):\mathbb{C}^{*}\rightarrow GL(X) be such that λ3(t)x=x\lambda_{3}(t)\cdot x=x for all xYx\in Y, while λ3(t)x=tx\lambda_{3}(t)\cdot x=tx for all xZx\in Z. We may write:

λ4(t)det3(X)=t[x1(x5x9x6x8)+x7(x2x6x3x5)]+t2[x4(x2x9x3x8)]\lambda_{4}(t)\cdot det^{3}(X)=t[x_{1}(x_{5}x_{9}-x_{6}x_{8})+x_{7}(x_{2}x_{6}-x_{3}x_{5})]+t^{2}[-x_{4}(x_{2}x_{9}-x_{3}x_{8})]

We may write this as tQ4+t2Q4tQ_{4}+t^{2}Q^{\prime}_{4}, where Q4=x1(x5x9x6x8)+x7(x2x6x3x5)Q_{4}=x_{1}(x_{5}x_{9}-x_{6}x_{8})+x_{7}(x_{2}x_{6}-x_{3}x_{5}) and Q4=x4(x2x9x3x8)Q^{\prime}_{4}=-x_{4}(x_{2}x_{9}-x_{3}x_{8}). We see that dim()=21dim({\cal H})=21 and exceeds dim(𝒦)+1dim({\cal K})+1. Moreover, 𝒦{\cal K} has pure elements of degree 1-1 and 11 with respect to 4\ell_{4} and 𝒦4f{\cal K}_{\ell_{4}f} is of dimension 88.

Remark 5.13

The stabilizer 4{\cal H}_{4} of Q4Q_{4} is of dimension 2121. The tangent of exit is Q4Q^{\prime}_{4} and its stabilizer within 𝒦{\cal K} is of dimension 88. Let 4=log(t1/3λ4(t))\ell_{4}=\log(t^{-1/3}\lambda_{4}(t)), then 44\ell_{4}\in{\cal H}_{4} and 4det3(X)=Q4\ell_{4}\cdot det^{3}(X)=Q^{\prime}_{4}, the tangent of exit.

We end this section with the following table of filtered dimensions for various 1-PS operating on the form det3det_{3}, the limit gg and the dimensions of the graded components of the stabilizers of the limit gg and the tangent of exit f\ell f. 𝒦f{\cal K}_{\ell f} is the subalgebra of triple stabilizers. These calculations illustrate Lemma  4.6 and Proposition  4.3.

1PSformsdim((𝒦0)i)fdim(𝒦f)dim(()i)(f,g)1011det3088Q1088Q10+4+02det3088Q2088Q30+8+04det31105Q41137Q41+6+1\begin{array}[]{|c|c|c|c|c|c|c|}\hline\cr 1-PS&forms&\vrule\lx@intercol\hfil dim(({\cal K}_{0})_{i})\hfil\lx@intercol\vrule\lx@intercol&\ell f&dim({\cal K}_{\ell f})\\ &&\vrule\lx@intercol\hfil dim(({\cal H})_{i})\hfil\lx@intercol\vrule\lx@intercol&&\\ \hline\cr&(f,g)&1&0&-1&&\\ \hline\cr\hline\cr\ell_{1}&det_{3}&0&8&8&&-\\ &Q_{1}&0&8&8&Q^{\prime}_{1}&0+4+0\\ \hline\cr\ell_{2}&det_{3}&0&8&8&&-\\ &Q_{2}&0&8&8&Q_{3}&0+8+0\\ \hline\cr\ell_{4}&det_{3}&1&10&5&&-\\ &Q_{4}&1&13&7&Q^{\prime}_{4}&1+6+1\\ \hline\cr\end{array}

6 Local Stabilizers and Lie algebra Cohomology obstructions

We begin this section with two examples. Both concern the action of a 1-PS on a form ff with leading term gg and direction of approach fbf_{b}. We continue to use the notation 𝒦(t){\cal K}(t), 𝒦0{\cal K}_{0} from the previous sections. We give an explicit description of the limiting algebra in both these examples. This motivates the question of recovering 𝒦(t){\cal K}(t) from the 𝒦0{\cal K}_{0}.

Example 6.1

Let X={z,y}X=\{z,y\} and let f=(y2+z2)2f=(y^{2}+z^{2})^{2}. Let g=y4g=y^{4} and note that with λ(t)y=y\lambda(t)\cdot y=y and λ(t)z=tz\lambda(t)\cdot z=t\cdot z, gives us:

f(t)=λ(t)f=y4+2t2y2z2+t4z4f(t)=\lambda(t)\cdot f=y^{4}+2t^{2}\cdot y^{2}z^{2}+t^{4}z^{4}

Thus, limt0λ(t)f(t)=g\lim_{t\rightarrow 0}\lambda(t)f(t)=g, and that a=0a=0 and b=2b=2 in the notation of this section.

Let us order XX as x1=zx_{1}=z and x2=yx_{2}=y, and let e11,e12,e21,e22e_{11},e_{12},e_{21},e_{22} be basis vectors of gl(X)=gl2gl(X)=gl_{2}. In other words, we have:

e11=[1000]e12=[0100]e21=[0010]e22=[0001]e_{11}=\left[\begin{array}[]{cc}1&0\\ 0&0\\ \end{array}\right]\>\>e_{12}=\left[\begin{array}[]{cc}0&1\\ 0&0\\ \end{array}\right]\>\>e_{21}=\left[\begin{array}[]{cc}0&0\\ 1&0\\ \end{array}\right]\>\>e_{22}=\left[\begin{array}[]{cc}0&0\\ 0&1\\ \end{array}\right]\>\>

The stabilizer {\cal H} of gg consists of matrices e11+e12\mathbb{C}\cdot e_{11}+\mathbb{C}\cdot e_{12}, as show below:

=[00]{\cal H}=\left[\begin{array}[]{cc}*&*\\ 0&0\\ \end{array}\right]

We may choose 𝒮=e21+e22{\cal S}=\mathbb{C}\cdot e_{21}+\mathbb{C}\cdot e_{22} and note that e21g=4y3ze_{21}\cdot g=4y^{3}z and e22g=4ge_{22}\cdot g=4g. Thus TOgTO_{g} consists of the vector y4+y3z\mathbb{C}\cdot y^{4}+\mathbb{C}\cdot y^{3}z. We may decompose V=TVgV=TV_{g} as:

V=(y4+y3z)(y2z2+yz3+z4)=TOgN.V=(\mathbb{C}\cdot y^{4}+\mathbb{C}\cdot y^{3}z)\oplus(\mathbb{C}\cdot y^{2}z^{2}+\mathbb{C}\cdot yz^{3}+\mathbb{C}\cdot z^{4})=TO_{g}\oplus N.

We may solve the equation from Proposition 2.5(ii) as follows. As per the recipe of that equation, we to find 𝔥\mathfrak{h}, 𝔰\mathfrak{s} which when applied to the left and right of the expressions below within brackets, gives n=0n=0. We compute the terms of

𝔥(2t2y2z2+t4z4)=𝔰(y4+2t2y2z2+t4z4)+n,\mathfrak{h}(2t^{2}y^{2}z^{2}+t^{4}z^{4})=\mathfrak{s}\cdot(y^{4}+2t^{2}y^{2}z^{2}+t^{4}z^{4})+n,

in the following table:

𝔥λ𝒮λNe1104t2y2z2+4t4z4e12t2e210\begin{array}[]{|c|c|c|}\hline\cr\mathfrak{h}&\lambda_{\cal S}&\lambda_{N}\\ \hline\cr\hline\cr e_{11}&0&4t^{2}y^{2}z^{2}+4t^{4}z^{4}\\ \hline\cr e_{12}&t^{2}e_{21}&0\\ \hline\cr\end{array}

The stabilizer condition that λN=0\lambda_{N}=0 gives us the stabilizer of f(t)f(t) as the element 𝔨(t)=e12t2e21\mathfrak{k}(t)=e_{12}-t^{2}e_{21} as shown below:

𝔨(t)=[01t20]\mathfrak{k}(t)=\left[\begin{array}[]{cc}0&1\\ -t^{2}&0\\ \end{array}\right]

One may check, that 𝒦(t)=𝔨(t){\cal K}(t)=\mathbb{C}\cdot\mathfrak{k}(t), and that 𝒦0=e12{\cal K}_{0}=\mathbb{C}\cdot e_{12}. We see that fb=4y2z2f_{b}=4y^{2}z^{2} and verify that e12fb=4y3zTOge_{12}\cdot f_{b}=4y^{3}z\in TO_{g}, and thus 𝒦0b{\cal K}_{0}\subseteq{\cal H}_{b}, the stabilizer of fb¯\overline{f_{b}} under the \star-action of {\cal H}. \Box

Example 6.2

Let X={z,y1,y2}X=\{z,y_{1},y_{2}\} in that order and let f=(y12+y22+z2)2f=(y_{1}^{2}+y_{2}^{2}+z^{2})^{2} and g=(y12+y22)2g=(y_{1}^{2}+y_{2}^{2})^{2}. Let λ\lambda be such that λ(t)yi=yi\lambda(t)\cdot y_{i}=y_{i} while λ(t)z=tz\lambda(t)\cdot z=tz. Thus:

f(t)=(y12+y22)2+2t2z2(y12+y22)+t4z4f(t)=(y_{1}^{2}+y_{2}^{2})^{2}+2t^{2}z^{2}(y_{1}^{2}+y_{2}^{2})+t^{4}z^{4}

Let us have the following elements:

𝒦(t)={[0abat20cbt2c0]|a,b,c}={=[dab00c0c0]|a,b,c,d}{\cal K}(t)=\left\{\left[\begin{array}[]{ccc}0&a&b\\ -at^{2}&0&c\\ -bt^{2}&-c&0\\ \end{array}\right]|a,b,c\in\mathbb{C}\right\}\>\>{\cal H}=\left\{=\left[\begin{array}[]{ccc}d&a&b\\ 0&0&c\\ 0&-c&0\\ \end{array}\right]|a,b,c,d\in\mathbb{C}\right\}

Whence

𝒦0={[0ab00c0c0]|a,b,c}{\cal K}_{0}=\{\left[\begin{array}[]{ccc}0&a&b\\ 0&0&c\\ 0&-c&0\\ \end{array}\right]|\>a,b,c\in\mathbb{C}\}

We note that 𝒦0{\cal K}_{0}\subseteq{\cal H} and that 𝒦0{\cal K}_{0} stabilizes fb¯=(y12+y22)z2¯\overline{f_{b}}=\overline{(y_{1}^{2}+y_{2}^{2})z^{2}} for the \star-action of {\cal H}. In other words, 𝒦0fbTOg{\cal K}_{0}\cdot f_{b}\subseteq TO_{g}.

It is instructive to form the multiplication table for 𝒦(t){\cal K}(t). Let us put:

𝔨1(t)=[010t200000]𝔨2(t)=[001000t200]𝔨3(t)=[000001010]\mathfrak{k}_{1}(t)=\left[\begin{array}[]{ccc}0&1&0\\ -t^{2}&0&0\\ 0&0&0\\ \end{array}\right]\mathfrak{k}_{2}(t)=\left[\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ -t^{2}&0&0\\ \end{array}\right]\mathfrak{k}_{3}(t)=\left[\begin{array}[]{ccc}0&0&0\\ 0&0&1\\ 0&-1&0\\ \end{array}\right]

We then have the following structure constants of the Lie bracket:

[𝔨1(t),𝔨2(t)𝔨1(t),𝔨3(t)𝔨2(t),𝔨3(t)]=[00t2010100][𝔨1(t)𝔨2(t)𝔨3(t)]\left[\begin{array}[]{c}\mathfrak{k}_{1}(t),\mathfrak{k}_{2}(t)\\ \mathfrak{k}_{1}(t),\mathfrak{k}_{3}(t)\\ \mathfrak{k}_{2}(t),\mathfrak{k}_{3}(t)\end{array}\right]=\left[\begin{array}[]{rrr}0&0&-t^{2}\\ 0&1&0\\ -1&0&0\\ \end{array}\right]\left[\begin{array}[]{c}\mathfrak{k}_{1}(t)\\ \mathfrak{k}_{2}(t)\\ \mathfrak{k}_{3}(t)\end{array}\right]

Here 𝔨1(t),𝔨2(t)\mathfrak{k}_{1}(t),\mathfrak{k}_{2}(t) on the left denotes [𝔨1(t),𝔨2(t)][\mathfrak{k}_{1}(t),\mathfrak{k}_{2}(t)] and likewise for the other rows on the left. We note that {𝔨i(0)|i=1,2,3}\{\mathfrak{k}_{i}(0)|i=1,2,3\} is a basis for 𝒦0{\cal K}_{0} and the structure constants are precisely those of 𝒦(t){\cal K}(t) specialized to t=0t=0.

In the above example the structure constants (αijk(t))(\alpha_{ij}^{k}(t)) of 𝒦(t){\cal K}(t) were elements of [t]\mathbb{C}[t] and where putting t=0t=0 gave us 𝒦0{\cal K}_{0}. Moreover we see that the \mathbb{C}-algebra 𝒦(t0){\cal K}(t_{0}), i.e., 𝒦(t){\cal K}(t) instantiated at a t0t_{0}\in\mathbb{C}, may be very different from 𝒦0{\cal K}_{0}, and perhaps “more semisimple” than both {\cal H} or 𝒦0{\cal K}_{0}. We look at the process of constructing 𝒦(t){\cal K}(t) from 𝒦0{\cal K}_{0}. This naturally leads us to a Lie algebra cohomology formulation. We continue with the notation and transversality assumption of Section  3.1.

For any system (αijk(t))(\alpha_{ij}^{k}(t)) of structure constants of 𝒦(t){\cal K}(t), the verification that these do satisfy the Jacobi conditions is in the polynomial ring [t]\mathbb{C}[t]. Let us consider the ring [ϵD][t]/(tD)\mathbb{C}[\epsilon_{D}]\cong\mathbb{C}[t]/(t^{D}) and Lie algebras over [ϵD]\mathbb{C}[\epsilon_{D}]. If (αijk(t))(\alpha_{ij}^{k}(t)) satisfy the Jacobi identity, then so do (αijk(ϵD))(\alpha_{ij}^{k}(\epsilon_{D})). Let us call the Lie algebra so generated as 𝒦D1{\cal K}_{D-1}. Note that 𝒦D1{\cal K}_{D-1} is a Lie algebra, both over \mathbb{C} and over [ϵD]\mathbb{C}[\epsilon_{D}] and that 𝒦0{\cal K}_{0} matches with our earlier definition. Note that if dim[t](𝒦(t))=kdim_{\mathbb{C}[t]}({\cal K}(t))=k, then each 𝒦D1{\cal K}_{D-1} is a free [ϵD]\mathbb{C}[\epsilon_{D}]-module of the same dimension. Moreover, for general a>0a>0, we also have surjective Lie algebra homomorphisms (over \mathbb{C}) 𝒦a𝒦a1{\cal K}_{a}\rightarrow{\cal K}_{a-1}. Finally, for any 𝒦(t){\cal K}(t) over [t]\mathbb{C}[t], there is a large DD, that the structure constants (αijk(t))(\alpha_{ij}^{k}(t)) may be verified in [ϵD]\mathbb{C}[\epsilon_{D}]. However, the corresponding 𝒦D1{\cal K}_{D-1} will not match 𝒦(t){\cal K}(t) in structure. With this remark, let us study the structure of 𝒦1{\cal K}_{1}.

Let [ϵ]\mathbb{C}[\epsilon] denote the ring formed by adding the variable ϵ\epsilon with ϵ2=0\epsilon^{2}=0. For any \mathbb{C}-vector space or algebra XX, let X[ϵ]X[\epsilon] denote the extension of coefficients, i.e., X[ϵ]=[ϵ]XX[\epsilon]=\mathbb{C}[\epsilon]\otimes_{\mathbb{C}}X. Note that X[ϵ]=XϵXX[\epsilon]=X\oplus\epsilon X as vector spaces. We also define the “evaluation at zero” map e0:X[ϵ]Xe_{0}:X[\epsilon]\rightarrow X as follows. For an element x=x0+ϵx1X[ϵ]x=x_{0}+\epsilon x_{1}\in X[\epsilon], we set e0(x)=x0e_{0}(x)=x_{0}. This will also be denoted as x(0)x(0). Let V[ϵ]V[\epsilon] and 𝒢[ϵ]{\cal G}[\epsilon] be obtained from VV and 𝒢{\cal G} as above. Note that 𝒢[ϵ]{\cal G}[\epsilon] is a Lie algebra and V[ϵ]V[\epsilon], a 𝒢[ϵ]{\cal G}[\epsilon] module. By the same token V[ϵ]V[\epsilon] is also an [ϵ]{\cal H}[\epsilon] module under the \star-action.

Lemma 6.3

𝒢[ϵ]{\cal G}[\epsilon] is a Lie algebgra over \mathbb{C} and the map e0:𝒢[ϵ]𝒢e_{0}:{\cal G}[\epsilon]\rightarrow{\cal G} defined as e0(𝔤+ϵ𝔤)=𝔤e_{0}(\mathfrak{g}+\epsilon\mathfrak{g}^{\prime})=\mathfrak{g} is a Lie \mathbb{C}-algebra homomorphism.

Let us consider the element p=g+ϵfbV[ϵ]p=g+\epsilon f_{b}\in V[\epsilon] and construct the stabilizer of the point pp in 𝒢[ϵ]{\cal G}[\epsilon]. For a typical element 𝔤0+ϵ𝔤1\mathfrak{g}_{0}+\epsilon\mathfrak{g}_{1}, we see that the stabilizer condition implies:

𝔤0g=0 and 𝔤0fb+𝔤1g=0\mathfrak{g}_{0}\cdot g=0\>\>\mbox{ and }\mathfrak{g}_{0}\cdot f_{b}+\mathfrak{g}_{1}\cdot g=0 (6)

The first condition states that 𝔤0\mathfrak{g}_{0}\in{\cal H}. The second condition further implies that 𝔤0b\mathfrak{g}_{0}\in{\cal H}_{b}\subseteq{\cal H}, the stabilizer of fb¯\overline{f_{b}} under the \star-action of {\cal H} on VV. This condition may also be interpreted as follows. Let us identify TOgTO_{g} with the quotient 𝒢/{\cal G}/{\cal H}, given simply by dg:𝔤𝔤gTOgd_{g}:\mathfrak{g}\rightarrow\mathfrak{g}\cdot g\in TO_{g}.

Lemma 6.4

The map dg:𝒢/TOgVd_{g}:{\cal G}/{\cal H}\rightarrow TO_{g}\subseteq V is an isomorphism of {\cal H}-modules.

Proof: Note that {\cal H} acts on VV as maps in End(V)End(V) under the Lie bracket. Let 𝔰𝒢\mathfrak{s}\in{\cal G} and let {𝔰}\{\mathfrak{s}\} denote the coset 𝔰+\mathfrak{s}+{\cal H}. The operation of 𝔥\mathfrak{h}\in{\cal H} on 𝒢/{\cal G}/{\cal H} is given by 𝔥{𝔰}={[𝔥,𝔰]}\mathfrak{h}\cdot\{\mathfrak{s}\}=\{[\mathfrak{h},\mathfrak{s}]\}. Whence, we then have:

dg({[𝔥,𝔰]})=[𝔥,𝔰]g=(𝔥𝔰𝔰𝔥)g=𝔥𝔰g (since 𝔥g=0)=𝔥dg(𝔰)\begin{array}[]{rcl}d_{g}(\{[\mathfrak{h},\mathfrak{s}]\})&=&[\mathfrak{h},\mathfrak{s}]\cdot g\\ &=&(\mathfrak{h}\cdot\mathfrak{s}-\mathfrak{s}\cdot\mathfrak{h})\cdot g\\ &=&\mathfrak{h}\cdot\mathfrak{s}\cdot g\mbox{ (since $\mathfrak{h}\cdot g=0$)}\\ &=&\mathfrak{h}\cdot d_{g}(\mathfrak{s})\end{array}

This completes the proof. \Box

Lemma 6.5

The stabilizer p𝒢]ϵ]{\cal H}_{p}\subseteq{\cal G}]\epsilon] is given by the set 𝔥0+ϵ𝔤\mathfrak{h}_{0}+\epsilon\mathfrak{g} such that (i) 𝔥0b\mathfrak{h}_{0}\in{\cal H}_{b}, and (ii) 𝔤dg1(𝔥0fb)\mathfrak{g}\in d_{g}^{-1}(-\mathfrak{h}_{0}f_{b}).

Proof: Let us look at the two parts of Eq. 6. The first implies that 𝔥0\mathfrak{h}_{0}\in{\cal H}. The second implies that, in fact, 𝔥0\mathfrak{h}_{0} stabilize fb¯\overline{f_{b}}, and thus 𝔥0b\mathfrak{h}_{0}\in{\cal H}_{b}. This proves (i). For (ii), note that the second condition of Eq. 6, gives us that 𝔥0fb+dg(𝔤)=0\mathfrak{h}_{0}\cdot f_{b}+d_{g}(\mathfrak{g})=0. This implies (ii). \Box

Let us now define the map db:b𝒢/d_{b}:{\cal H}_{b}\rightarrow{\cal G}/{\cal H} as follows. For any element 𝔰𝒢\mathfrak{s}\in{\cal G}, let {𝔰}\{\mathfrak{s}\} denote the coset 𝔰+\mathfrak{s}+{\cal H} in 𝒢/{\cal G}/{\cal H}. For an 𝔥b\mathfrak{h}\in{\cal H}_{b}, let db(𝔥)d_{b}(\mathfrak{h}) be that element {𝔰}𝒢/\{\mathfrak{s}\}\in{\cal G}/{\cal H} such that 𝔰g=𝔥fb\mathfrak{s}\cdot g=\mathfrak{h}\cdot f_{b}, i.e., dg({𝔰})=𝔥fbd_{g}(\{\mathfrak{s}\})=\mathfrak{h}\cdot f_{b}.

Definition 6.6

A derivation (see, for example, [Wag10]dd from 𝒢{\cal G} to a 𝒢{\cal G}-module W is a {\mathbb{C}}-linear map from 𝒢{\cal G} to WW such that for all 𝔥1,𝔥2𝒢\mathfrak{h}_{1},\mathfrak{h}_{2}\in{\cal G} we have

d([𝔥1,𝔥2])=𝔥1d(𝔥2)𝔥2d(𝔥1).d([\mathfrak{h}_{1},\mathfrak{h}_{2}])=\mathfrak{h}_{1}\circ d(\mathfrak{h}_{2})-\mathfrak{h}_{2}\circ d(\mathfrak{h}_{1}).
Lemma 6.7

(1) 𝒢/{\cal G}/{\cal H} is an b{\cal H}_{b}-module. The map db:b𝒢/d_{b}:{\cal H}_{b}\rightarrow{\cal G}/{\cal H} is a derivation. (2) Moreover, the stabilizer p𝒢]ϵ]{\cal H}_{p}\subseteq{\cal G}]\epsilon] is also given by the set 𝔥+ϵ𝔤\mathfrak{h}+\epsilon\mathfrak{g} such that (i) 𝔥0b\mathfrak{h}_{0}\in{\cal H}_{b}, and (ii) 𝔤db(𝔥0)\mathfrak{g}\in d_{b}(-\mathfrak{h}_{0}).

Proof: Since [b,][{\cal H}_{b},{\cal H}]\subseteq{\cal H}, the quotient 𝒢/{\cal G}/{\cal H} is an b{\cal H}_{b}-module under the adjoint action. Let 𝔥1,𝔥2b\mathfrak{h}_{1},\mathfrak{h}_{2}\in{\cal H}_{b}. Suppose 𝔥ifb=𝔰ig\mathfrak{h}_{i}f_{b}=\mathfrak{s}_{i}g for some 𝔰i𝒢\mathfrak{s}_{i}\in{\cal G}, whence db(𝔥i)={𝔰i}d_{b}(\mathfrak{h}_{i})=\{\mathfrak{s}_{i}\}, the coset of 𝔰i\mathfrak{s}_{i}. We must show that db([𝔥1,𝔥2])=𝔥1db(𝔥2)𝔥2db(𝔥1)d_{b}([\mathfrak{h}_{1},\mathfrak{h}_{2}])=\mathfrak{h}_{1}\circ d_{b}(\mathfrak{h}_{2})-\mathfrak{h}_{2}\circ d_{b}(\mathfrak{h}_{1}) as elements of 𝒢/{\cal G}/{\cal H}. We check this through the isomorphism dgd_{g}. Applying dgd_{g} to the RHS, we get:

(𝔥1db(𝔥2)𝔥2db(𝔥1))g=({[𝔥1,𝔰2]}{[𝔥2,𝔰1]})g=𝔥1𝔰2g𝔥2𝔰1g (since 𝔥ig=0)\begin{array}[]{rcl}(\mathfrak{h}_{1}\circ d_{b}(\mathfrak{h}_{2})-\mathfrak{h}_{2}\circ d_{b}(\mathfrak{h}_{1}))\cdot g&=&(\{[\mathfrak{h}_{1},\mathfrak{s}_{2}]\}-\{[\mathfrak{h}_{2},\mathfrak{s}_{1}]\})\cdot g\\ &=&\mathfrak{h}_{1}\mathfrak{s}_{2}g-\mathfrak{h}_{2}\mathfrak{s}_{1}g\mbox{ (since $\mathfrak{h}_{i}\cdot g=0$)}\end{array}

We check this with the LHS

dg([𝔥1,𝔥2])=[𝔥1,𝔥2]fb=𝔥1𝔥2fb𝔥2𝔥1fb=(𝔥1𝔰2𝔥2𝔰1)gd_{g}([\mathfrak{h}_{1},\mathfrak{h}_{2}])=[\mathfrak{h}_{1},\mathfrak{h}_{2}]\cdot f_{b}=\mathfrak{h}_{1}\mathfrak{h}_{2}f_{b}-\mathfrak{h}_{2}\mathfrak{h}_{1}f_{b}=(\mathfrak{h}_{1}\mathfrak{s}_{2}-\mathfrak{h}_{2}\mathfrak{s}_{1})\cdot g

This proves (1). Concerning (2), the only change from lemma 6.5 is in (ii). It is easily seen that 𝔤dg1(𝔥0fb)\mathfrak{g}\in d_{g}^{-1}(-\mathfrak{h}_{0}f_{b}) is equivalent to the condition that 𝔤db(𝔥0)\mathfrak{g}\in d_{b}(-\mathfrak{h}_{0}). \Box

Remark 6.8

The above claim is equivalent to the statement that p{\cal H}_{p} is a Lie subalgebra of 𝒢[ϵ]{\cal G}[\epsilon].

We collect some properties of p{\cal H}_{p}.

Lemma 6.9

The image e0(p)e_{0}({\cal H}_{p}) is b{\cal H}_{b}. Moreover, the nilpotent subalgebra ϵ𝒢[ϵ]\epsilon{\cal H}\subseteq{\cal G}[\epsilon] is a subset of p{\cal H}_{p}.

Definition 6.10

Let 𝒦b{\cal K}\subseteq{\cal H}_{b} be a subalgebra and let dim(𝒦)=Kdim_{\mathbb{C}}({\cal K})=K. Let subalgebra 𝒦¯p\overline{{\cal K}}\subseteq{\cal H}_{p} be such that (i) 𝒦¯\overline{\cal K} is generated as a [ϵ]\mathbb{C}[\epsilon]-module by KK elements 𝔨1,,𝔨K\mathfrak{k}_{1},\ldots,\mathfrak{k}_{K}, and (ii) e0(𝒦¯)=𝒦e_{0}(\overline{\cal K})={\cal K}. We call 𝒦¯\overline{\cal K} an ϵ\epsilon-deformation of 𝒦{\cal K}. Note that dim(𝒦¯)=2Kdim_{\mathbb{C}}(\overline{\cal K})=2K.

The condition that 𝒦¯\overline{\cal K} is generated by the KK elements as a [ϵ]\mathbb{C}[\epsilon]-module implies that we can write the Lie bracket using the structure coefficients (αijk)(\alpha_{ij}^{k}), as:

[𝔨i,𝔨j]=kαijk𝔨k for some (αijk)[ϵ][\mathfrak{k}_{i},\mathfrak{k}_{j}]=\sum_{k}\alpha_{ij}^{k}\mathfrak{k}_{k}\mbox{ for some $(\alpha_{ij}^{k})\in\mathbb{C}[\epsilon]$}
Lemma 6.11

If 𝒦¯\overline{\cal K} is an ϵ\epsilon-deformation of 𝒦{\cal K} then (𝔟i)=(e0(𝔨i))(\mathfrak{b}_{i})=(e_{0}(\mathfrak{k}_{i})) is a \mathbb{C}-basis for 𝒦{\cal K} and its structure coefficients are precisely (αijk(0))(\alpha_{ij}^{k}(0)), the evaluations of (αijk)(\alpha_{ij}^{k}) at 0. Thus, we have:

[𝔟i,𝔟j]=kαijk(0)𝔨k[\mathfrak{b}_{i},\mathfrak{b}_{j}]=\sum_{k}\alpha_{ij}^{k}(0)\mathfrak{k}_{k}
Proposition 6.12

The existence of 𝒦¯\overline{\cal K} is equivalent to the existence of a derivation db¯:𝒦𝒢/𝒦\overline{d_{b}}:{\cal K}\rightarrow{\cal G}/{\cal K} which extends the derivation db:𝒦𝒢/d_{b}:{\cal K}\rightarrow{\cal G}/{\cal H} as shown in the diagram below:

𝒢/𝒦{{{\cal G}/{\cal K}}}𝒢/{{{\cal G}/{\cal H}}}𝒦{{\cal K}}b{{\cal H}_{b}}db¯\scriptstyle{\overline{d_{b}}}db\scriptstyle{d_{b}}

Proof: Suppose that db¯\overline{d_{b}} is such an extension. Let us define 𝒦¯\overline{\cal K} as follows:

𝒦¯={𝔥+ϵ𝔰|𝔥𝒦,𝔰𝒢 such that 𝔰db¯(𝔥)}ϵ𝒦\overline{\cal K}=\{\mathfrak{h}+\epsilon\mathfrak{s}|\mathfrak{h}\in{\cal K},\mathfrak{s}\in{\cal G}\mbox{ such that }\mathfrak{s}\in\overline{d_{b}}(-\mathfrak{h})\}\oplus\epsilon{\cal K}

Clearly, since db¯\overline{d_{b}} extends dbd_{b}, we have db¯(𝔥)db(𝔥)\overline{d_{b}}(-\mathfrak{h})\subseteq d_{b}(-\mathfrak{h}) and hence every element of 𝒦¯\overline{\cal K} stabilizes pp. Thus 𝒦¯p\overline{\cal K}\subseteq{\cal H}_{p}. Other properties of Defn. 6.10 are easily verified. Thus, all that remains to show is that 𝒦¯\overline{\cal K} is closed under the Lie bracket, i.e., it is a subalgebra of p{\cal H}_{p}.

Let us now verify the Lie bracket condition. For any elements 𝔨1=𝔥1+ϵ𝔰1\mathfrak{k}_{1}=\mathfrak{h}_{1}+\epsilon\mathfrak{s}_{1} and 𝔨2=𝔥2+ϵ𝔰2\mathfrak{k}_{2}=\mathfrak{h}_{2}+\epsilon\mathfrak{s}_{2} we must check that [𝔨1,𝔨2]𝒦¯[\mathfrak{k}_{1},\mathfrak{k}_{2}]\in\overline{\cal K}. But

[𝔨1,𝔨2]=[𝔥1,𝔥2]+ϵ([𝔥1,𝔰2]+[𝔰1,𝔥2])[\mathfrak{k}_{1},\mathfrak{k}_{2}]=[\mathfrak{h}_{1},\mathfrak{h}_{2}]+\epsilon([\mathfrak{h}_{1},\mathfrak{s}_{2}]+[\mathfrak{s}_{1},\mathfrak{h}_{2}])

Since [𝔥1,𝔥2]𝒦[\mathfrak{h}_{1},\mathfrak{h}_{2}]\in{\cal K}, there is an element [𝔥1,𝔥2]+ϵ𝔰𝒦¯[\mathfrak{h}_{1},\mathfrak{h}_{2}]+\epsilon\mathfrak{s}\in\overline{\cal K} with the condition that 𝔰db¯([𝔥1,𝔥2])\mathfrak{s}\in\overline{d_{b}}(-[\mathfrak{h}_{1},\mathfrak{h}_{2}]). We must check that

𝔰([𝔥1𝔰2]+[𝔰1,𝔥2])𝒦.\mathfrak{s}-([\mathfrak{h}_{1}\mathfrak{s}_{2}]+[\mathfrak{s}_{1},\mathfrak{h}_{2}])\in{\cal K}.

Now, as elements of 𝒢/𝒦{\cal G}/{\cal K}, we have:

[𝔥1,𝔰2]+[𝔰1,𝔥2]=[𝔥1,𝔰2][𝔥2,𝔰1]=𝔥1db¯(𝔥2)𝔥2db¯(𝔥1)=db¯([𝔥1,𝔥2]) since db¯ is a derivation=𝔰+𝒦\begin{array}[]{rcl}[\mathfrak{h}_{1},\mathfrak{s}_{2}]+[\mathfrak{s}_{1},\mathfrak{h}_{2}]&=&[\mathfrak{h}_{1},\mathfrak{s}_{2}]-[\mathfrak{h}_{2},\mathfrak{s}_{1}]\\ &=&\mathfrak{h}_{1}\cdot\overline{d_{b}}(-\mathfrak{h}_{2})-\mathfrak{h}_{2}\cdot\overline{d_{b}}(-\mathfrak{h}_{1})\\ &=&\overline{d_{b}}(-[\mathfrak{h}_{1},\mathfrak{h}_{2}])\mbox{ since $\overline{d_{b}}$ is a derivation}\\ &=&\mathfrak{s}+{\cal K}\\ \end{array}

Thus 𝔰[𝔥1,𝔰2][𝔰1,𝔥2]𝒦\mathfrak{s}-[\mathfrak{h}_{1},\mathfrak{s}_{2}]-[\mathfrak{s}_{1},\mathfrak{h}_{2}]\in{\cal K}. Whence:

[𝔥1+ϵ𝔰1,𝔥2+ϵ𝔰2]=[𝔥1,𝔥2]+ϵ𝔰+ϵ𝔥 with 𝔥𝒦[\mathfrak{h}_{1}+\epsilon\mathfrak{s}_{1},\mathfrak{h}_{2}+\epsilon\mathfrak{s}_{2}]=[\mathfrak{h}_{1},\mathfrak{h}_{2}]+\epsilon\mathfrak{s}+\epsilon\mathfrak{h}\mbox{ with }\mathfrak{h}\in{\cal K}

This proves that 𝒦¯\overline{\cal K} in indeed a Lie subalgebra. Note that ϵ𝒦ϵ𝒦¯𝒦¯\epsilon{\cal K}\subseteq\epsilon\overline{\cal K}\subseteq\overline{\cal K} and thus if 𝔥+ϵ𝔰𝒦¯\mathfrak{h}+\epsilon\mathfrak{s}\in\overline{\cal K} then so is 𝔥+ϵ𝔰\mathfrak{h}+\epsilon\mathfrak{s}^{\prime}, where 𝔰𝔰+𝒦\mathfrak{s}^{\prime}\in\mathfrak{s}+{\cal K}.

Conversely, if 𝒦¯={𝔥+ϵ𝔰|𝔥𝒦}\overline{\cal K}=\{\mathfrak{h}+\epsilon\mathfrak{s}|\mathfrak{h}\in{\cal K}\} is such an extension, then put db¯(𝔥)=𝔰+𝒦\overline{d_{b}}(\mathfrak{h})=-\mathfrak{s}+{\cal K}. Since 𝒦¯p\overline{\cal K}\subseteq{\cal H}_{p}, we have 𝔰+=db(𝔥)-\mathfrak{s}+{\cal H}=d_{b}(\mathfrak{h}) as well. Thus db¯(𝔥)db(𝔥)\overline{d_{b}}(\mathfrak{h})\subseteq d_{b}(\mathfrak{h}) and db¯\overline{d_{b}} extends dbd_{b}.

We must now show that db¯\overline{d_{b}} is a derivation. Towards this, let 𝔥i+ϵ𝔰i\mathfrak{h}_{i}+\epsilon\mathfrak{s}_{i} and [𝔥1,𝔥2]+ϵ𝔰[\mathfrak{h}_{1},\mathfrak{h}_{2}]+\epsilon\mathfrak{s} be elements of 𝒦¯\overline{\cal K}. By the closure of 𝒦¯\overline{\cal K} under Lie bracket, we have:

[𝔥1,𝔰2]+[𝔰1,𝔥2]𝔰+𝒦[\mathfrak{h}_{1},\mathfrak{s}_{2}]+[\mathfrak{s}_{1},\mathfrak{h}_{2}]\in\mathfrak{s}+{\cal K}

Let us verify the derivation condition:

𝔥1db¯(𝔥2)𝔥2db¯(𝔥1)=[𝔥1,𝔰2]+[𝔥2,𝔰1]+𝒦=([𝔥1,𝔰2]+[𝔰1,𝔥2])+𝒦=𝔰+𝒦=db¯([𝔥1,𝔥2])\begin{array}[]{rcl}\mathfrak{h}_{1}\circ\overline{d_{b}}(\mathfrak{h}_{2})-\mathfrak{h}_{2}\circ\overline{d_{b}}(\mathfrak{h}_{1})&=&-[\mathfrak{h}_{1},\mathfrak{s}_{2}]+[\mathfrak{h}_{2},\mathfrak{s}_{1}]+{\cal K}\\ &=&-([\mathfrak{h}_{1},\mathfrak{s}_{2}]+[\mathfrak{s}_{1},\mathfrak{h}_{2}])+{\cal K}\\ &=&-\mathfrak{s}+{\cal K}\\ &=&\overline{d_{b}}([\mathfrak{h}_{1},\mathfrak{h}_{2}])\end{array}

This proves the proposition. \Box

Remark 6.13

We have the exact sequence of 𝒦{\cal K}-modules [Wag10]:

0/𝒦𝒢/𝒦𝒢/00\longrightarrow{\cal H}/{\cal K}\longrightarrow{\cal G}/{\cal K}\longrightarrow{\cal G}/{\cal H}\longrightarrow 0

and the corresponding long exact sequence of cohomology modules:

0H0(𝒦,/𝒦)H0(𝒦,𝒢/𝒦)H0(𝒦,𝒢/)H1(𝒦,/𝒦)H1(𝒦,𝒢/𝒦)H1(𝒦,𝒢/)H2(𝒦,/𝒦)\begin{array}[]{rcl}0&\longrightarrow&H^{0}({\cal K},{\cal H}/{\cal K})\longrightarrow H^{0}({\cal K},{\cal G}/{\cal K})\longrightarrow H^{0}({\cal K},{\cal G}/{\cal H})\\ &\longrightarrow&H^{1}({\cal K},{\cal H}/{\cal K})\longrightarrow H^{1}({\cal K},{\cal G}/{\cal K})\longrightarrow H^{1}({\cal K},{\cal G}/{\cal H})\\ &\longrightarrow&H^{2}({\cal K},{\cal H}/{\cal K})\longrightarrow\ldots\end{array}

Since both dbd_{b} and db¯\overline{d_{b}} are derivations, they belong to the spaces H1(𝒦,𝒢/)H^{1}({\cal K},{\cal G}/{\cal H}) and H1(𝒦,𝒢/𝒦)H^{1}({\cal K},{\cal G}/{\cal K}) respectively (but they may be 0). Whence, parts of the long exact sequence which are relevant to the extension of dbd_{b} to db¯\overline{d_{b}} are:

H1(𝒦,/𝒦)H1(𝒦,𝒢/𝒦)H1(𝒦,𝒢/)H2(𝒦,/𝒦)\longrightarrow H^{1}({\cal K},{\cal H}/{\cal K})\longrightarrow H^{1}({\cal K},{\cal G}/{\cal K})\longrightarrow H^{1}({\cal K},{\cal G}/{\cal H})\\ \longrightarrow H^{2}({\cal K},{\cal H}/{\cal K})\longrightarrow
Remark 6.14

Our construction of the ϵ\epsilon-extension 𝒦¯\overline{\cal K} through a module leads to derivations as the direction of infinitesimal extension. This is a variation of the ideas of Nijenhuis and Richardson, [NR67], where the 2-cocyles are the infinitesimal directions of deformations. All the same, it is likely that 𝒦0{\cal K}_{0} is not rigid (again, [NR67]) while 𝒦(t0){\cal K}(t_{0}) are, for generic t0t_{0}\in\mathbb{C}.

We now come to the main proposition of this section:

Definition 6.15

Let A(t),fA(t),f and gg be such that:

A(t)f=f(t)=tag+tbfb+ higher termsA(t)\cdot f=f(t)=t^{a}g+t^{b}f_{b}+\mbox{ higher terms}

We say that gg is a regular limit of ff via A(t)A(t) if (i) in the above expression fc0f_{c}\neq 0 only when cac-a is a multiple of bab-a, and (ii) if {\cal H} is the stabilizer of gg, then the stabilizer 𝒦(1){\cal K}(1) of ff is not contained inside {\cal H}.

Proposition 6.16

Let gg be a regular limit of ff via A(t)A(t) and f(t)f(t) and fbf_{b} be as above. Let 𝒦(t){\cal K}(t) be the stabilizing Lie algebra of f(t)f(t) and {\cal H} that of gg. Moreover, let 𝒦0b{\cal K}_{0}\subseteq{\cal H}_{b} be the limit of 𝒦(t){\cal K}(t), as t0t\rightarrow 0. Let db:𝒦0𝒢/d_{b}:{\cal K}_{0}\rightarrow{\cal G}/{\cal H} be the derivation as above. Then there is a derivation db¯:𝒦0𝒢/𝒦0\overline{d_{b}}:{\cal K}_{0}\rightarrow{\cal G}/{\cal K}_{0} which extends dbd_{b}.

Proof: Let δ=ba\delta=b-a. Condition (i) of regularity implies that that after removal of tat^{a}, all non-zero terms in the expression for f(t)f(t) have powers of tt as tkδt^{k\delta} for some k0k\geq 0. We may absorb tat^{a} into A(t)A(t) and write:

A(t)f=f(t)=g+tδfb+ higher terms in powers of tδA(t)\cdot f=f(t)=g+t^{\delta}f_{b}+\mbox{ higher terms in powers of $t^{\delta}$}

This, in turn, implies that the matrices MN(t)M_{N}(t) and M𝒮(t)M_{\cal S}(t) have elements in [tδ]\mathbb{C}[t^{\delta}]. Thus the basis for 𝒦(t){\cal K}(t) may be written in the form {𝔨i(tδ)}i=1k\{\mathfrak{k}_{i}(t^{\delta})\}_{i=1}^{k}. Now condition (ii) that 𝒦{\cal K}\not\subseteq{\cal H} implies that NN{\cal H}\cdot N\not\subseteq N and that there is a 𝔨𝒦\mathfrak{k}\in{\cal K} with 𝔨=𝔥+𝔰\mathfrak{k}=\mathfrak{h}+\mathfrak{s}, with 𝔥\mathfrak{h}\in{\cal H} and 𝔰𝒮\mathfrak{s}\in{\cal S}. This implies that λ𝒮(𝔥fb)0\lambda_{\cal S}(\mathfrak{h}\cdot f_{b})\neq 0 and thus λ𝒮(𝔥f+(t))\lambda_{\cal S}(\mathfrak{h}\cdot f^{+}(t)) must be of the form tδ𝔰1+higher termst^{\delta}\mathfrak{s}_{1}+\mbox{higher terms}. Thus some element 𝔨i(t)=𝔥0+𝔤1tδ+ higher terms\mathfrak{k}_{i}(t)=\mathfrak{h}_{0}+\mathfrak{g}_{1}t^{\delta}+\mbox{ higher terms} is such that 𝔤10\mathfrak{g}_{1}\neq 0. This implies that if (αijk(T))(\alpha_{ij}^{k}(T)) are the structure constants of 𝒦(t){\cal K}(t), then there are some constants i,j,ki^{\prime},j^{\prime},k^{\prime} such that αijk=c1tδ+c2t2δ+\alpha_{i^{\prime}j^{\prime}}^{k^{\prime}}=c_{1}t^{\delta}+c_{2}t^{2\delta}+\ldots, where c10c_{1}\neq 0. Whence, there is a non-trivial ϵ\epsilon-extension 𝒦1{\cal K}_{1} of 𝒦0{\cal K}_{0}. By Prop. 6.12, this is equivalent to the assertion that there is a derivation db¯:𝒦0𝒢/𝒦0\overline{d_{b}}:{\cal K}_{0}\rightarrow{\cal G}/{\cal K}_{0} which extends dbd_{b}.

Remark 6.17

The Lie alegbra 𝒦0¯\overline{{\cal K}_{0}} may not yield the “additional semisimplcity” which may be present in 𝒦(t0){\cal K}(t_{0}) (and absent in 𝒦0{\cal K}_{0}), but its structure coefficients do approximate those of 𝒦(t){\cal K}(t) better as polynomials in tt. See Example 6.1, where 𝒦0{\cal K}_{0} is the 11-dimensional algebra generated by AA below, while, 𝒦0¯\overline{{\cal K}_{0}} is 22-dimensional (over \mathbb{C}) and generated by A(ϵ)A(\epsilon). On the other hand 𝒦(t0){\cal K}(t_{0}) is 11-dimensional and generated by A(t0)A(t_{0}).

A=[0100] A(ϵ)=[01ϵ0] A(t0)=[01t020]A=\left[\begin{array}[]{cc}0&1\\ 0&0\end{array}\right]\mbox{ }A(\epsilon)=\left[\begin{array}[]{cc}0&1\\ -\epsilon&0\end{array}\right]\mbox{ }A(t_{0})=\left[\begin{array}[]{cc}0&1\\ -t_{0}^{2}&0\end{array}\right]
Remark 6.18

The existence of the limit limt0f(t)=g\lim_{t\rightarrow 0}f(t)=g allows the construction of 𝒦(t){\cal K}(t), the limit 𝒦0{\cal K}_{0} and the tangent of approach fbf_{b}. We also have the additional \star-action of {\cal H} and that 𝒦0b{\cal K}_{0}\subseteq{\cal H}_{b}, the stabilizer of fb¯\overline{f_{b}} under this \star-action. Thus, we have the picture below:

𝒦(t)lim𝒦0?limb\begin{array}[]{ccc}{\cal K}(t)&\stackrel{{\scriptstyle lim}}{{\rightarrow}}&{\cal K}_{0}\\ \downarrow&&\downarrow\\ ?&\stackrel{{\scriptstyle lim}}{{\rightarrow}}&{\cal H}_{b}\\ \end{array}

In other words, we already know that there is a local 𝒢{\cal G}-action in the vicinity of the orbit O(g)O(g) of gg. Does this action allow us to compute the stabilizer 𝒢x,n{\cal G}_{x,n} ( possibly over (t)\mathbb{C}(t)), of the tangent vector nTVgn\in TV_{g}? This may then be applied to the vector fbf_{b} to obtain a possible localization of 𝒦(t){\cal K}(t). The current recipe of 𝒦0¯\overline{{\cal K}_{0}} over [ϵ]\mathbb{C}[\epsilon] is not entirely satisfactory.

Remark 6.19

The codimension 11 case This case yields two important simplifications. Firstly, the equality of b{\cal H}_{b} and 𝒦0{\cal K}_{0} simplifies the extension problem as below:

𝒢/b{{{\cal G}/{\cal H}_{b}}}𝒢/{{{\cal G}/{\cal H}}}b{{\cal H}_{b}}db¯\scriptstyle{\overline{d_{b}}}db\scriptstyle{d_{b}}

The second simplification comes from the theorem of Hoffman[Hof65], stated below:

Theorem 6.20

Let 𝒦{\cal K}\subseteq{\cal H} be Lie alegbras over \mathbb{R} such that 𝒦{\cal K} is of codimension 11. Then exactly one of the following three cases must be true:

  1. 1.

    Let PP be the 22-dimensional parabolic subgroup of upper triangular matrices in sl2sl_{2}. Then there is an ideal {\cal I} such that 𝒦/P{\cal K}/{\cal I}\cong P and /sl2{\cal H}/{\cal I}\cong sl_{2}.

  2. 2.

    Let DPD\subseteq P the subalgebra of diagonal matrices in PP. Then there is an ideal {\cal I} such that 𝒦/D{\cal K}/{\cal I}\cong D and /P{\cal H}/{\cal I}\cong P

  3. 3.

    𝒦{\cal K} is an ideal of {\cal H}.

Assuming that our Lie algebras are complex extensions of real Lie algebras, we see that b{\cal H}_{b}\subset{\cal H} must fall into one of these. Whence, the 11-dimensional b{\cal H}_{b}-module /b{\cal H}/{\cal H}_{b} has a simple structure in the exact sequence below:

H1(b,/b)H1(b,𝒢/b)H1(b,𝒢/)H2(b,/b)\longrightarrow H^{1}({\cal H}_{b},{\cal H}/{\cal H}_{b})\longrightarrow H^{1}({\cal H}_{b},{\cal G}/{\cal H}_{b})\longrightarrow H^{1}({\cal H}_{b},{\cal G}/{\cal H})\longrightarrow H^{2}({\cal H}_{b},{\cal H}/{\cal H}_{b})\longrightarrow

7 The single matrix under conjugation

In this section we use the local model to solve a different problem - computing the projective limits of stable and polystable points of the vector space of n×nn\times n matrices under conjugation. This problem has been studied extensively in the (affine) setting of semisimple Lie groups under the adjoint action. The proof of the main theorem in this section, Theorem 7.22, illustrates the use of the local parametrization and the map θ\theta, to obtain projective closures. The statement of the main theorem was inspired by calculations performed using the local model of the neighbourhood of a matrix with a single Jordon block resulting in Theorem 7.7.

Remark 7.1

Experts in the area inform us that the statement of Theorem 7.22 does not surprise them and that it probably follows from earlier work. However we are not aware of any literature where this question has been discussed.

Before proceeding to describe our results we set up some notation and recall some well known results about closures of orbits in affine space, in type A.

Let VV be the GL(X)GL(X)-module End(X)End(X) where GL(X)GL(X) acts by conjugation. Thus, given AGL(X)A\in GL(X) and YVY\in V, we have AY=AYA1A\cdot Y=AYA^{-1}. Polystable points in VV are the diagonalizable matrices, and a generic stable point is a diagonalizable matrix with distinct eigenvalues, also called a regular semisimple element.

It is well known that the null-cone 𝒩{\cal N} for this action is the set of all nilpotent matrices,see [New78]. The closure of nilpotent matrices in affine space is well understood, see [Ger61],[Hes76]. The main theorem of this section has a description similar to that of closures of nilpotent matrices in affine space, which we review in Remark  7.17.

If NVN\in V is a nilpotent matrix, in its GG-orbit is the Jordan canonical form of NN. Let us quickly recall this. The b×bb\times b-matrix JbJ_{b} is given by the conditions: Jb(i,j)=0J_{b}(i,j)=0 for all (i,j)(i,j) except for the tuple (i,i+1)(i,i+1), for i=1,,b1i=1,\ldots,b-1, where it is 11. J4J_{4} is shown below:

J4=[0100001000010000]J_{4}=\left[\begin{array}[]{cccc}0&1&0&0\\ 0&0&1&0\\ 0&0&0&1\\ 0&0&0&0\end{array}\right]

JbJ_{b} is determined by the properties that (i) it is a b×bb\times b matrix, (ii) Jba0J_{b}^{a}\neq 0 for a<ba<b, and (iii) Jbb=0J_{b}^{b}=0. Given NN, any n×nn\times n nilpotent matrix, its Jordan canonical form is given as the block diagonal matrix with blocks Jλ1,,JλkJ_{\lambda_{1}},\ldots,J_{\lambda_{k}}, with λ1λk>0\lambda_{1}\geq\ldots\geq\lambda_{k}>0 and iλi=n\sum_{i}\lambda_{i}=n, or in other words, a partition λ\lambda of nn. This partition λ\lambda will be called the nilpotent signature of NN.

Let us assume now that XX is an nn-dimensional complex vector space and that we have chosen a basis of XX. With respect to this basis every element in VV can be identified with an n×nn\times n matrix.

Let JnJ_{n} be the nilpotent matrix given by Jn(i,i+1)=1J_{n}(i,i+1)=1 for i=1,,n1i=1,\ldots,n-1, and zero otherwise. This is the single largest orbit in 𝒩{\cal N}, and 𝒩{\cal N} is the closure of the orbit of JnJ_{n}.

In the next section we study the neighbourhood of JnJ_{n} in a direction normal to the orbit of JnJ_{n} under conjugation. We use the local model to show that a polystable point whose projective orbit closure contains JnJ_{n} must be necessarily stable.

In Section 7.2 we consider the case Ja,bJ_{a,b} and describe the semisimple and nilpotent matrices in its neighborhood.

In Section 7.3, motivated by the above two computations, we state our main theorem which describes the nilpotent matrices which are in the projective closure of polystable points. This is described by the combinatorics of multiplicities of the eigenvalues of the matrix.

To simplify notation we don’t use fraktur symbols to denote Lie algebra elements in the next two sections. Instead we use letters A,B,CA,B,C\ldots to denote matrices.

7.1 The neighbourhood of JnJ_{n}

Recall that

Jn=[010001000010000].J_{n}=\left[\begin{array}[]{ccccc}0&1&0&\ldots&\\ 0&0&1&0&\ldots\\ &\vdots&&\vdots&\\ 0&0&\ldots&0&1\\ 0&0&\ldots&0&0\end{array}\right].

The action ρ\rho of the Lie algebra gl(X)gl(X) on VV is given by ρ(A)(Y)=AYYA\rho(A)(Y)=A\cdot Y-Y\cdot A, where Agl(X)A\in gl(X) and YVY\in V. For k=(n1),,1,0,1,,n1k=-(n-1),\ldots,-1,0,1,\ldots,n-1, let ZkZ_{k} be the n×nn\times n-matrix where Zk(i,j)=0Z_{k}(i,j)=0 unless j=i+kj=i+k, and then Zi,i+k=1Z_{i,i+k}=1. Thus Jn=Z1J_{n}=Z_{1}. We see that ZiZj=Zi+jZ_{i}Z_{j}=Z_{i+j} if i,j0i,j\leq 0 or i,j0i,j\geq 0.

Lemma 7.2

Let JnJ_{n} be as above. Then:

  1. 1.

    The tangent space of the orbit OO of p=Jnp=J_{n} under the GL(X)GL(X) action is

    TOp={BEnd(X)|i=k+1nbi,i+k=0 for (n1)k0}TO_{p}=\{B\in End(X)|\sum_{i=-k+1}^{n}b_{i,i+k}=0\mbox{ for $-(n-1)\leq k\leq 0$}\}
  2. 2.

    A complement to TOpTO_{p} within TVp=End(X)TV_{p}=End(X) is the space of matrices 𝒞n{\cal C}_{n} is

    𝒞n(c)=[cn100cn2000c1000c0000]{\cal C}_{n}(c)=\left[\begin{array}[]{lcccc}-c_{n-1}&0&0&\ldots&\\ -c_{n-2}&0&0&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -c_{1}&0&\ldots&0&0\\ -c_{0}&0&\ldots&0&0\end{array}\right]

    where c=[c0,,cn1]Tnc=[c_{0},\ldots,c_{n-1}]^{T}\in\mathbb{C}^{n} is a column vector. Note that the affine space Jn+𝒞nJ_{n}+{\cal C}_{n} is the space of all companion matrices:

    Jn+𝒞n=[cn110cn2010c1001c0000]J_{n}+{\cal C}_{n}=\left[\begin{array}[]{lcccc}-c_{n-1}&1&0&\ldots&\\ -c_{n-2}&0&1&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -c_{1}&0&\ldots&0&1\\ -c_{0}&0&\ldots&0&0\end{array}\right]

    where c0,,cn1c_{0},\ldots,c_{n-1}\in\mathbb{C}.

  3. 3.

    The stabilizer Lie algebra {\cal H} of JnJ_{n} is the collection of all matrices i=0n1aiZi\sum_{i=0}^{n-1}a_{i}Z_{i}, with aia_{i}\in\mathbb{C}

  4. 4.

    A complement 𝒮{\cal S} to {\cal H} within gl(X)gl(X) is the collection of all matrices Bgl(X)B\in gl(X) such that B1,i=0B_{1,i}=0 for i=1,,ni=1,\ldots,n and for every element vTOpv\in TO_{p}, there is a unique element a𝒮a\in{\cal S} such that ρ(a)(Jn)=v\rho(a)(J_{n})=v.

Proof: The action of the Lie algebra gl(X)gl(X) on pp is given by AAJnJnA=BA\rightarrow AJ_{n}-J_{n}A=B. But B=(bij)=(ai,j1ai+1,j)B=(b_{ij})=(a_{i,j-1}-a_{i+1,j}). We thus see that

(I)i=1nkbi,i+k=a1,kank+1,nfor 1kn1(II)i=k+1nbi,i+k=0for (n1)k0\begin{array}[]{lrcll}(I)&\sum_{i=1}^{n-k}b_{i,i+k}&=&a_{1,k}-a_{n-k+1,n}&\mbox{for $1\leq k\leq n-1$}\\ (II)&\sum_{i=-k+1}^{n}b_{i,i+k}&=&0&\mbox{for $-(n-1)\leq k\leq 0$}\\ \end{array}

In fact, it is easily shown that for any matrix BB such that the nn equations constituting (II)(II) hold, we may find an Agl(X)A\in gl(X) such that B=AJnJnAB=AJ_{n}-J_{n}A. This proves the first assertion. The second assertion follows from the fact that 𝒞nTOp=0¯{\cal C}_{n}\cap TO_{p}=\overline{0}, and that dim(𝒞n)=ndim({\cal C}_{n})=n. Assertion (3) is classical. Assertion (4) follows from (1) and (2). \Box

Remark 7.3

The minimal polynomial of Jn+𝒞n(c)J_{n}+{\cal C}_{n}(c) is p(X)=Xn+cn1Xn1++c0X0p(X)=X^{n}+c_{n-1}X^{n-1}+\ldots+c_{0}X^{0}.

Proof: This is classical. Let us use TT for the matrix 𝒞n(c){\cal C}_{n}(c). If en=[0,,0,1]Tne_{n}=[0,\ldots,0,1]^{T}\in\mathbb{C}^{n} is the nn-th column vector, then it is easily seen that Tien=eniT^{i}\cdot e_{n}=e_{n-i} and that

Tnen=cn1e1cn2e2+c0en=cn1Tn1encn2Tn2en+c0T0en\begin{array}[]{rcl}T^{n}\cdot e_{n}&=&-c_{n-1}e_{1}-c_{n-2}e_{2}+\ldots-c_{0}e_{n}\\ &=&-c_{n-1}T^{n-1}\cdot e_{n}-c_{n-2}T^{n-2}\cdot e_{n}+\ldots-c_{0}T^{0}\cdot e_{n}\\ \end{array}

Thus, we have (Tn+cn1Tn1++c0T0)en=0(T^{n}+c_{n-1}T^{n-1}+\ldots+c_{0}T^{0})\cdot e_{n}=0. This, and the fact that {Tien\{T^{i}\cdot e_{n} are linearly independent vectors show that the polynomial pp is the minimial polynomial of TT. \Box

We now prove the following lemma.

Lemma 7.4

θi(n)\theta^{i}(n) is identically zero for i2i\geq 2.

Proof: Since θ(n)\theta(n) is zero on NVN\subset V, it suffices to calculate it on the orbit which we have identified with 𝒮{\cal S}. Given a matrix B𝒮B\in{\cal S} such that B(1,i)=0B(1,i)=0 for all ii and an element C𝒞nC\in{\cal C}_{n} shown below, we see that BCCBBC-CB is precisely BCBC since CB=0CB=0.

C=[cn100cn2000c1000c0000]C=\left[\begin{array}[]{lcccc}-c_{n-1}&0&0&\ldots&\\ -c_{n-2}&0&0&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -c_{1}&0&\ldots&0&0\\ -c_{0}&0&\ldots&0&0\end{array}\right]

We may thus write B=[0,b2T,,bnT]TB=[0,b_{2}^{T},\ldots,b_{n}^{T}]^{T}, where bib_{i} are row matrices and C=[c,0,,0]C=[c,0,\ldots,0], where cc is a column matrix, to get

BCCB=[000b2c00bnc00]BC-CB=\left[\begin{array}[]{cccc}0&0&\ldots&0\\ b_{2}\cdot c&0&\ldots&0\\ \vdots&&\vdots&\\ b_{n}\cdot c&0&\ldots&0\\ \end{array}\right]

Whence BCCBNBC-CB\in N Since θi(n)\theta^{i}(n), i2i\geq 2, is the composition of projection of θi1(n)\theta^{i-1}(n) onto 𝒮{\cal S} followed by θ\theta, it follows that θi(n)\theta^{i}(n) is zero for every i>1i>1.\Box

To carry out the program of the local model outlined in Theorem 2.7 we need to understand how {\cal H} acts on the space NN, as given in the sequence given below. Note from the description of {\cal H} in Lemma 7.2, {\cal R}, the reductive part of {\cal H} is zero, so in fact =𝒬{\cal H}={\cal Q}.

We show

nVλ𝒮𝒮n0{\cal H}\stackrel{{\scriptstyle\cdot n}}{{\longrightarrow}}V\stackrel{{\scriptstyle\lambda_{\cal S}}}{{\longrightarrow}}{\cal S}\stackrel{{\scriptstyle\cdot n}}{{\longrightarrow}}0

For AA\in{\cal H} and CNC\in N we have the first arrow:

CAAC=[cn1a0cn1a1cn1an1cn2a0cn2a1cn2an1c0a0c0a1c0an1][i=0n1cn1iai00i=0n2cn2iai00c0a000]CA-AC=\left[\begin{array}[]{cccc}c_{n-1}a_{0}&c_{n-1}a_{1}&\ldots&c_{n-1}a_{n-1}\\ c_{n-2}a_{0}&c_{n-2}a_{1}&\ldots&c_{n-2}a_{n-1}\\ \vdots&&\vdots&\\ c_{0}a_{0}&c_{0}a_{1}&\ldots&c_{0}a_{n-1}\\ \end{array}\right]-\left[\begin{array}[]{cccc}\sum_{i=0}^{n-1}c_{n-1-i}a_{i}&0&\ldots&0\\ \sum_{i=0}^{n-2}c_{n-2-i}a_{i}&0&\ldots&0\\ \vdots&&\vdots&\\ c_{0}a_{0}&0&\ldots&0\\ \end{array}\right]

Since the expression is linear for {\cal H}, let us assume that ai=1a_{i}=1 for a particular ii, and aj=0a_{j}=0 for jij\neq i. Whence, we have:

CAiAiC=[cni000000cni10000000c0000cni000000cni100000000c0000]CA_{i}-A_{i}C=\left[\begin{array}[]{ccccccccc}-c_{n-i}&0&\ldots&0\ldots&&0&0&0&0\\ -c_{n-i-1}&0&\ldots&0\ldots&&0&0&0&0\\ \vdots&\vdots&&\vdots\ldots&&\vdots&0\\ -c_{0}&\vdots&\vdots&&&\vdots&0&0&0\\ \vdots&&\vdots&&&c_{n-i}&0&0&0\\ 0&0&\ldots&0\ldots&&c_{n-i-1}&0&0&0\\ \vdots&&\vdots&&&\vdots&0&0&0\\ 0&0&\ldots&&&c_{0}&0&0&0\\ \end{array}\right]

The sum of the entries in the diagonals j,(n1)j0j,-(n-1)\leq j\leq 0 are all zero, so from Lemma 7.2 this is in the orbit TOJnTO_{J_{n}}.

Thus there is an S𝒮S\in{\cal S} such that SJnJnS=OiS\cdot J_{n}-J_{n}\cdot S=O_{i}. Since SCCS=0S\cdot C-C\cdot S=0 we have

𝒬nVλ𝒮𝒮nV=0{\cal Q}\stackrel{{\scriptstyle\cdot n}}{{\longrightarrow}}V\stackrel{{\scriptstyle\lambda_{\cal S}}}{{\longrightarrow}}{\cal S}\stackrel{{\scriptstyle\cdot n}}{{\longrightarrow}}V=0

\Box

Our final computation is the dimension of the stabilizer of an arbitrary point sitting normally over JnJ_{n}.

Lemma 7.5

The dimension of the stabilizer of any point Jn+𝒞n(c)J_{n}+{\cal C}_{n}(c) is nn.

Proof: Let T=𝒞n(c)T={\cal C}_{n}(c). By Theorem 2.7, the stabilizer of the point Jn+TJ_{n}+T is governed by the equation:

ΔrT+λN(1+θ(n))1(ΔqT)=0\Delta r\cdot T+\lambda_{N}\circ(1+\theta(n))^{-1}(\Delta q\cdot T)=0

where 𝔯,𝔮𝒬{\mathfrak{r}}\in{\cal R},\mathfrak{q}\in{\cal Q}, the reductive and nilpotent Lie subalgebras of {\cal H}. As mentioned earlier =0{\cal R}=0. From the calculations in the preceeding paragraph, the image of 𝒬n{\cal Q}\cdot n is in 𝒮{\cal S} and so λN(1+θ(n))1(ΔqT)=0\lambda_{N}\circ(1+\theta(n))^{-1}(\Delta q\cdot T)=0 for all 𝔮{\mathfrak{q}}. So every elements of 𝒬{\cal Q} may be supplemented by an element of 𝒮{\cal S} to given an element of the stabilier of T+JnT+J_{n}. \Box

Remark 7.6

That the matrix T+JnT+J_{n} is in companion form actually tells us that its minimal polynomial is of degree nn

We now come to the main theorem:

Theorem 7.7

Let AEnd(X)A\in End(X) be polystable under the action of SL(X)SL(X) by conjugation. If JnJ_{n} is a projective limit of a matrix AA under the adjoint action, then AA has nn distinct eigenvalues and is stable. Moreover, for any matrix AA such that AA has distinct eigenvalues λ1,,λn\lambda_{1},\ldots,\lambda_{n}, there is a family 𝒜(t){\cal A}(t) such that (i) the eigenvalues of 𝒜(t){\cal A}(t) are tλ1,,tλnt\lambda_{1},\ldots,t\lambda_{n} and (ii) limt0𝒜(t)=Jn\lim_{t\rightarrow 0}{\cal A}(t)=J_{n}.

Proof: Under the hypothesis, we know that there is a GL(X)GL(X)-conjugate AA^{\prime} such that AJn<ϵ\|A^{\prime}-J_{n}\|<\epsilon. By the local model construction, we may further assume that AA^{\prime} is of the form Jn+TJ_{n}+T, with TNT\in N. Since AA is polystable, so is AA^{\prime} and so it is diagonalizable. Since A=Jn+TA^{\prime}=J_{n}+T as above, its minimal polynomial is of degree nn. Thus, AA^{\prime} and therefore AA has nn distinct eigenvalues. This proves the first part.

For the second part, let 𝒞n(c){\cal C}_{n}(c) be the companion form of AA. Thus, the vector c=(cn1,,c0)c=(c_{n-1},\ldots,c_{0}) are the coefficients of the characteristic polynomial of AA. Define c(t)=(tcn1,t2cn2t2,,tnc0)c(t)=(tc_{n-1},t^{2}c_{n-2}t^{2},\ldots,t^{n}c_{0}) and let 𝒜(t)=Jn+𝒞(c(t)){\cal A}(t)=J_{n}+{\cal C}(c(t)), i.e., the companion matrix with column vector c(t)c(t). Then 𝒜(t){\cal A}(t) has eigenvalues (tλ1,,tλn)(t\lambda_{1},\ldots,t\lambda_{n}) and limt0𝒜(t)=Jn\lim_{t\rightarrow 0}{\cal A}(t)=J_{n}. \Box

Example 7.8

Let us compute explicitly through 𝒮{\cal S}-completion, the stabilizer of the matrix Z4=J4+C4Z_{4}=J_{4}+C_{4} sitting ”normally” above J4J_{4}:

Z4=[0100001000011000]C4=[0000000000001000]Z_{4}=\left[\begin{array}[]{cccc}0&1&0&0\\ 0&0&1&0\\ 0&0&0&1\\ 1&0&0&0\\ \end{array}\right]\>\>\>C_{4}=\left[\begin{array}[]{cccc}0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ 1&0&0&0\\ \end{array}\right]

The matrix Z4Z_{4} has the minimal polynomial X41X^{4}-1. The stabilizing Lie algebra of Z4Z_{4} is given by the basis elements Z4iZ_{4}^{i}, for i=0,,3i=0,\ldots,3.

The complementary space 𝒮{\cal S} is the 12-dimensional space of matrices shown below:

𝒮=[0000]{\cal S}=\left[\begin{array}[]{cccc}0&0&0&0\\ &*&*&*\\ &*&*&*\\ &*&*&*\\ \end{array}\right]

The elementary matrices EijE_{ij}, for i=2,3,4i=2,3,4 and j=1,2,3,4j=1,2,3,4 form a basis for 𝒮{\cal S}. The tangent space to the orbit of J4J_{4} is given by the basis of 12 matrices EijJ4J4EijE_{ij}\cdot J_{4}-J_{4}\cdot E_{ij}. The stabilizer {\cal H} of J4J_{4} is the space of matrices generated by J4iJ_{4}^{i}, i=0,,3i=0,\dots,3. The stabilizer space of Z4Z_{4} is again 44-dimensional and given by the elements h+sh+s where hh\in{\cal H} and s𝒮s\in{\cal S} such that [s,J4]+[h,C4]=0[s,J_{4}]+[h,C_{4}]=0. Let us select the element h=J43h=J_{4}^{3} to obtain [h,C4][h,C_{4}] as shown below:

h=[0001000000000000][h,C4]=[1000000000000001]h=\left[\begin{array}[]{cccc}0&0&0&1\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ \end{array}\right]\>\>\>[h,C_{4}]=\left[\begin{array}[]{cccc}1&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&-1\\ \end{array}\right]

One may check that [s,J4][s,J_{4}] below and verify that [s,J4]=[h,C4][s,J_{4}]=-[h,C_{4}]:

s=E21+E32+E43=[0000100001000010][s,J4]=[1000000000000001]s=E_{21}+E_{32}+E_{43}=\left[\begin{array}[]{cccc}0&0&0&0\\ 1&0&0&0\\ 0&1&0&0\\ 0&0&1&0\\ \end{array}\right]\>\>\>[s,J_{4}]=\left[\begin{array}[]{cccc}-1&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&1\\ \end{array}\right]

Thus s+hs+h shown below, stabilizes Z4Z_{4}.

s+h=[0001100001000010]s+h=\left[\begin{array}[]{cccc}0&0&0&1\\ 1&0&0&0\\ 0&1&0&0\\ 0&0&1&0\\ \end{array}\right]

7.2 The Ja,bJ_{a,b} case

We now consider the partition (a,b)(a,b) where a+b=na+b=n and aba\geq b. The matrix Ja,bJ_{a,b} may be written in the block-diagonal form as below:

Ja,b=[Ja00Jb]J_{a,b}=\left[\begin{array}[]{cc}J_{a}&0\\ 0&J_{b}\end{array}\right]
Lemma 7.9

The stabilizer {\cal H} of Ja,bJ_{a,b} is of the form:

[ZaYabYbaZb]\left[\begin{array}[]{cc}Z_{a}&Y_{ab}\\ Y_{ba}&Z_{b}\end{array}\right]

Where ZmZ_{m} is m×mm\times m such that (i) Zm(i,j)=0Z_{m}(i,j)=0 for all i>ji>j, and (ii) Zm(i+1,j+1)=Zm(i,j)Z_{m}(i+1,j+1)=Z_{m}(i,j). The forms for YabY_{ab} and YbaY_{ba} are given below:

Yab=[Zb0]Yba=[0Zb′′]Y_{ab}=\left[\begin{array}[]{c}Z^{\prime}_{b}\\ 0\end{array}\right]\>\>\>\>Y_{ba}=\left[\begin{array}[]{cc}0&Z^{\prime\prime}_{b}\end{array}\right]

The dimension of {\cal H} is a+3ba+3b, aa coming from ZaZ_{a} and bb each from the three matrices Zb,Zb,Zb′′Z_{b},Z^{\prime}_{b},Z^{\prime\prime}_{b}.

Lemma 7.10

A normal space to the orbit of Ja,bJ_{a,b} is 𝒞{\cal C} below:

𝒞=[𝒞aWabWba𝒞b]{\cal C}=\left[\begin{array}[]{cc}{\cal C}_{a}&W_{ab}\\ W_{ba}&{\cal C}_{b}\end{array}\right]

where 𝒞a{\cal C}_{a} and 𝒞b{\cal C}_{b} are as shown below:

𝒞a(c)=[ca100ca2000c1000c0000]𝒞b(d)=[db100db2000d1000d0000]{\cal C}_{a}(c)=\left[\begin{array}[]{lcccc}-c_{a-1}&0&0&\ldots&\\ -c_{a-2}&0&0&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -c_{1}&0&\ldots&0&0\\ -c_{0}&0&\ldots&0&0\end{array}\right]\>\>\>{\cal C}_{b}(d)=\left[\begin{array}[]{lcccc}-d_{b-1}&0&0&\ldots&\\ -d_{b-2}&0&0&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -d_{1}&0&\ldots&0&0\\ -d_{0}&0&\ldots&0&0\end{array}\right]

and WabW_{ab} and WbaW_{ba} are of the following form:

Wab=[0(a1)×bα1αb]Wba=[β10b×(a1)βb]W_{ab}=\left[\begin{array}[]{ccc}&&\\ &0_{(a-1)\times b}&\\ &&\\ \hline\cr\alpha_{1}&\ldots&\alpha_{b}\end{array}\right]\>\>\>W_{ba}=\left[\begin{array}[]{c|c}\beta_{1}&\\ \vdots&0_{b\times{(a-1)}}\\ \beta_{b}&\end{array}\right]

where ci,dj,αr,βsc_{i},d_{j},\alpha_{r},\beta_{s}\in\mathbb{C} and Or×sO_{r\times s} are zero matrices. The dimension of 𝒞{\cal C} is also a+3ba+3b.

The proofs of these lemmas are straightforward. We now come to two main lemmas.

Lemma 7.11

For any C𝒞C\in{\cal C}, let pCp_{C} be the minimal polynomial of T=Ja,b+CT=J_{a,b}+C. Then the degree of pCp_{C} is at least aa.

Proof: We may write TT in the jumbo-size matrix below:

[ca110ca20100(a1)×bc1001c0000α1α2αbβ1db110β2db20100b×(a1)d1001βbd0000]\left[\begin{array}[]{lcccc|lcccc}-c_{a-1}&1&0&\ldots&&&\\ -c_{a-2}&0&1&0&\ldots&&&&\\ \vdots&\vdots&&\vdots&&&\lx@intercol\hfil 0_{(a-1)\times b}\hfil\lx@intercol&\\ -c_{1}&0&\ldots&0&1&&&&\\ -c_{0}&0&\ldots&0&0&\alpha_{1}&\alpha_{2}&\ldots&&\alpha_{b}\\ \hline\cr\beta_{1}&&&&&-d_{b-1}&1&0&\ldots&\\ \beta_{2}&&&&&-d_{b-2}&0&1&0&\ldots\\ \vdots&\lx@intercol\hfil 0_{b\times(a-1)}\hfil\lx@intercol&&\vdots&\vdots&&\vdots&\\ &&&&&-d_{1}&0&\ldots&0&1\\ \beta_{b}&&&&&-d_{0}&0&\ldots&0&0\end{array}\right]

Let e1,,ea,ea+1,,ea+be_{1},\ldots,e_{a},e_{a+1},\ldots,e_{a+b} be a basis of n\mathbb{C}^{n} (as column vectors). From the matrix structure of TT, it is clear that for i=2,,ai=2,\ldots,a we have T(ei)=ei1T(e_{i})=e_{i-1}. Thus if Q=Tm+γm1Tm1++γ0T0Q=T^{m}+\gamma_{m{-1}}T^{m-1}+\ldots+\gamma_{0}T^{0} is any operator, where m<am<a, then Q(ea)=eam+γm1eam1++γ0eaQ(e_{a})=e_{a_{m}}+\gamma_{m-1}e_{a-m-1}+\ldots+\gamma_{0}e_{a}. When Q(ea)0Q(e_{a})\neq 0. Thus, TT cannot have a minimal polynomial of degree less than aa. \Box

Lemma 7.12

Let TT be as above and λC\lambda\in C, be any number. Let ker(TλI)ker(T-\lambda I) be the space of all vv such that T(v)=λvT(v)=\lambda v, then dim(ker(TλI))2dim(ker(T-\lambda I))\leq 2. Moreover, any vv in this space is determined by v1v_{1} and va+1v_{a+1}.

Proof: The condition that (TλI)v=0(T-\lambda I)v=0 gives us the following conditions:

cak+1v1λvk1+vk=0for k=2,,aβi1v1dbi+1va+1λva+i1+va+i=0for i=2,,b\begin{array}[]{rl}-c_{a-k+1}v_{1}-\lambda v_{k-1}+v_{k}=0&\mbox{for $k=2,\ldots,a$}\\ \beta_{i-1}v_{1}-d_{b-i+1}v_{a+1}-\lambda v_{a+i-1}+v_{a+i}=0&\mbox{for $i=2,\ldots,b$}\\ \end{array}

This proves the claim. \Box

For TT as above, we write

Ta=[ca110ca2010c1001c0000]Tb=[db110db2010d1001d0000]T_{a}=\left[\begin{array}[]{lcccc}-c_{a-1}&1&0&\ldots&\\ -c_{a-2}&0&1&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -c_{1}&0&\ldots&0&1\\ -c_{0}&0&\ldots&0&0\end{array}\right]\>\>\>{T}_{b}=\left[\begin{array}[]{lcccc}-d_{b-1}&1&0&\ldots&\\ -d_{b-2}&0&1&0&\ldots\\ \vdots&\vdots&&\vdots&\\ -d_{1}&0&\ldots&0&1\\ -d_{0}&0&\ldots&0&0\end{array}\right]

Note that, by Remark 7.3, the minimal polynomial of TaT_{a} (respectively TbT_{b}) is Ta+ca1Ta1++c0T0T^{a}+c_{a-1}T^{a-1}+\ldots+c_{0}T^{0} (respectively Tb+db1Ta1++d0T0T^{b}+d_{b-1}T^{a-1}+\ldots+d_{0}T^{0}).

Lemma 7.13

Suppose that the minimal polynomial pp of TT has degree aa. Then pp is the minimal polynomial of TaT_{a} and the minimal polynomial of TbT_{b} divides pp.

Proof: Let p(T)=Ta+γa1Ta1++γ0T0p(T)=T^{a}+\gamma_{a{-1}}T^{a-1}+\ldots+\gamma_{0}T^{0} be the minimal polynomial of TT of degree aa. Using the notation as in the proof of Lemma 7.11, p(T)(ea)=0p(T)(e_{a})=0 implies that γi=ci\gamma_{i}=c_{i} for i=0,,a1i=0,\ldots,a{-1} and βj=0\beta_{j}=0 for j=1,,bj=1,\ldots,b. Thus, p(T)=Ta+ca1Ta1++c0T0p(T)=T^{a}+c_{a{-1}}T^{a-1}+\ldots+c_{0}T^{0} is also the minimal polynomial of TaT_{a}.

We now write TT in block upper triangular form

T=[Ta0b×aTb]T=\left[\begin{array}[]{c|c}T_{a}&*\\ \hline\cr 0_{b\times a}&T_{b}\end{array}\right]

It is easy to see that for any polynomial QQ, Q(T)Q(T) is of the form

Q(T)=[Q(Ta)0b×aQ(Tb)]Q(T)=\left[\begin{array}[]{c|c}Q(T_{a})&*\\ \hline\cr 0_{b\times a}&Q(T_{b})\end{array}\right]

As p(T)=0p(T)=0, this shows that p(Tb)=0p(T_{b})=0. Therefore, the minimal polynomial of TbT_{b} divides pp. \Box

Proposition 7.14

The nilpotent matrices in the space Ja,b+𝒞J_{a,b}+{\cal C} have signature (a,b)(a^{\prime},b^{\prime}) with aaa^{\prime}\geq a. Moreover, Ja,bJ_{a,b} is in the projective orbit closure of Ja,bJ_{a^{\prime},b^{\prime}}.

Proof: Let JJ be a nilpotent matrix of the above form and let aa^{\prime} be the largest component of its signature. The conditions on the minimum degree of the minimal polynomial implies that aaa^{\prime}\geq a. The dimension of the kernel of JJ, with λ=0\lambda=0 in lemma 7.12, is only 22. That forces JJ to have only two components. Thus, J=Ja,bJ=J_{a^{\prime},b^{\prime}} with aaa^{\prime}\geq a. That every such Ja,bJ_{a^{\prime},b^{\prime}} is obtained is verified by putting all cc’s, dd’s, β\beta’s and α\alpha’s to zero except that αi=t\alpha_{i}=t, where t0t\neq 0. One may verify that this matrix Ji(t)J_{i}(t) is nilpotent with signature a=a+bi+1a^{\prime}=a+b-i+1 and b=i1b^{\prime}=i-1. We also see that limt0Ji(t)=Ja,b\lim_{t\rightarrow 0}J_{i}(t)=J_{a,b}, and thus Ja,bJ_{a,b} is in the projective orbit closure of Ja,bJ_{a^{\prime},b^{\prime}}. \Box

Proposition 7.15

The diagonalizable matrices in the space Ja,b+𝒞J_{a,b}+{\cal C} have at least aa distinct eigenvalues. Moreover, each eigenvalue has multiplicity at most 22. Moreover, for any such matrix AA, the matrix Ja,bJ_{a,b} is in its projective closure.

Proof: For a diagonalizable matrix TT, the number of distinct eigenvalues is equal to the degree of the minimal polynomial. Further, for an eigenvalue λ\lambda of TT, the (algebraic) multiplicity of λ\lambda is equal to its geometric multiplicity, namely dim(ker(TλI))dim(ker(T-\lambda I)). The claim now follows from Lemmas 7.11 and 7.12.

Coming to the second assertion, let {λ1,,λb,λb+1,,λa}\{\lambda_{1},\ldots,\lambda_{b^{\prime}},\lambda_{b^{\prime}+1},\ldots,\lambda_{a^{\prime}}\} be an ordering of the eigenvalues of TT such that λ1,,λb\lambda_{1},\ldots,\lambda_{b^{\prime}} have multiplicity 22 and thus a+b=na^{\prime}+b^{\prime}=n. By extending the argument of Theorem 7.7, Ja,bJ_{a^{\prime},b^{\prime}} is in the projective orbit closure of TT. By prop. 7.14, we have aaa^{\prime}\geq a, and thus Ja,bJ_{a,b} is in its projective orbit closure of Ja,bJ_{a^{\prime},b^{\prime}}, and therefore of TT. \Box

7.3 Orbit closure of polystable points

We end the section with a complete characterization of the orbit closure of polystable points in the vicinity of a single nilpotent matrix. The closure is completely determined by the combinatorics of the spectrum of the polystable point under consideration. We give an algebraic proof which is motivated by the constructions of the previous section.

Before we begin, let us recall some definitions. A partition α\alpha of nn is a sequence (α1αr>0)(\alpha_{1}\geq\cdots\geq\alpha_{r}>0) such that i=1rαi=n\sum_{i=1}^{r}\alpha_{i}=n.

Definition 7.16

Let α\alpha and β\beta be partitions of nn. The transpose γ\gamma of α\alpha is given by the numbers γ1γs>0\gamma_{1}\geq\ldots\geq\gamma_{s}>0 where γi=|{j|αji}|\gamma_{i}=|\{j|\alpha_{j}\geq i\}|, and is denoted by αT\alpha^{T}. It is easy to check that γ\gamma too is a partition of nn. We say that α\alpha dominates β\beta iff for all i>0i>0, j=1iαjj=1iβj\sum_{j=1}^{i}\alpha_{j}\geq\sum_{j=1}^{i}\beta_{j}. This is denoted by αβ\alpha\unrhd\beta or, equivalently, βα\beta\unlhd\alpha. It is easy to see that αβ\alpha\unrhd\beta iff βTαT\beta^{T}\unrhd\alpha^{T}.

The order described above on partitions of nn is a partial order called the dominance order.

Remark 7.17

Let XX be a nilpotent matrix with signature λ\lambda. The orbit closure of XX under conjugation contains all nilpotent matrices with partition signature λ\lambda^{\prime} with λλ\lambda\unrhd\lambda^{\prime}. This is a celebrated result proved by Gerstenhaber, [Ger61].

If an n×nn\times n-diagonalizable matrix has eigenvalues μ1,μ2,,μs\mu_{1},\mu_{2},\ldots,\mu_{s} with multiplicities λ:=λ1λ2λs\lambda:=\lambda_{1}\geq\lambda_{2}\geq\ldots\geq\lambda_{s}, we call λ\lambda, the spectrum partition of that matrix.

Proposition 7.18

Let 𝒳kr{\cal X}_{k}^{r} denote the set of all n×nn\times n matrices xx for which there exist λ1,,λk\lambda_{1},\ldots,\lambda_{k} such that rank((xλ1I)(xλ2I)(xλkI))r\mathrm{rank}((x-\lambda_{1}I)(x-\lambda_{2}I)\ldots(x-\lambda_{k}I))\leq r and det(xλiI)=0\mathrm{det}(x-\lambda_{i}I)=0 for all i=1,,ki=1,\ldots,k. Then 𝒳kr{\cal X}_{k}^{r} is a projective variety. Moreover, 𝒳kr{\cal X}_{k}^{r} is invariant under the action of GLnGL_{n} on xx by conjugation.

Proof: Let XX be an n×nn\times n matrix of variables and let Y={y1,,yk}Y=\{y_{1},\ldots,y_{k}\} be another set of indeterminates.

Let the affine variety 𝒱n2+k{\cal V}\subseteq{\mathbb{C}}^{n^{2}+k} be defined by the homogeneous ideal JJ in [X,Y]{\mathbb{C}}[X,Y] generated by det(XyiI)=0\mathrm{det}(X-y_{i}I)=0 for i=1,,ki=1,\ldots,k. We define the projection morphism ϕ:𝒱n2\phi:{\cal V}\to{\mathbb{C}}^{n^{2}} as follows: ϕ(A,b1,,bk)=A\phi(A,b_{1},\ldots,b_{k})=A. The induced map of the co-ordinate rings is simply the inclusion of [X][𝒱]=[X,Y]/J{\mathbb{C}}[X]\hookrightarrow{\mathbb{C}}[{\cal V}]={\mathbb{C}}[X,Y]/J. As det(XyiI)=0\mathrm{det}(X-y_{i}I)=0, each yi[𝒱]y_{i}\in{\mathbb{C}}[{\cal V}] is integral over [X]{\mathbb{C}}[X]. This shows that ϕ\phi is a finite map and hence maps closed sets to closed sets.

Consider the matrix Z=(Xy1I)(Xy2I)(XykI))Z=(X-y_{1}I)(X-y_{2}I)\ldots(X-y_{k}I)). Let P,QP,Q be subsets of [n][n] such that |P|=|Q|=r+1|P|=|Q|=r+1 and let fP,Qf_{P,Q} be the homogeneous form det(ZP,Q)\mathrm{det}(Z_{P,Q}), i.e., the determinant of the (P,Q)th(P,Q)^{\mathrm{th}} minor of ZZ. Consider the ideal I[𝒵]I\subseteq{\mathbb{C}}[{\cal Z}] generated by all such forms. This is a homogeneous ideal in the variables of XYX\cup Y. Its variety, say 𝒰𝒱{\cal U}\subseteq{\cal V} is indeed of points (A,λ1,,λk)𝒱(A,\lambda_{1},\ldots,\lambda_{k})\in{\cal V} such that rank((Aλ1I)(Aλ2I)(AλkI))r\mathrm{rank}((A-\lambda_{1}I)(A-\lambda_{2}I)\ldots(A-\lambda_{k}I))\leq r. The set 𝒳kr{\cal X}_{k}^{r} is precisely the image of the closed set 𝒰{\cal U} under the finite map ϕ\phi. As a result, 𝒳kr{\cal X}_{k}^{r} is also a closed set. The fact that 𝒳kr{\cal X}_{k}^{r} is closed under homothety and invariant under the conjugation action of GLnGL_{n} is obvious. \Box

Proposition 7.19

Let xx be an n×nn\times n diagonalizable matrix with spectrum partition λ=(λ1λ2λkλs)\lambda=(\lambda_{1}\geq\lambda_{2}\geq\cdots\lambda_{k}\geq\cdots\geq\lambda_{s}). The x𝒳krx\in{\cal X}_{k}^{r} iff λ1++λknr\lambda_{1}+\ldots+\lambda_{k}\geq n-r.

The proof is obvious.

Proposition 7.20

Let x=Jθ1Jθsx=J_{\theta_{1}}\oplus\ldots\oplus J_{\theta_{s}} be a nilpotent matrix with partition signature (θ1θs)(\theta_{1}\geq\ldots\geq\theta_{s}). Then x𝒳krx\in{\cal X}_{k}^{r} iff rank(xk)rrank(x^{k})\leq r. This is given by the condition i:θi>k(θik)r\sum_{i:\theta_{i}>k}(\theta_{i}-k)\leq r.

It will be useful to restate this condition in terms of the transpose partition.

Proposition 7.21

Let x=Jθx=J_{\theta} be a nilpotent matrix with partition signature θ=(θ1θs)\theta=(\theta_{1}\geq\ldots\geq\theta_{s}) whose transpose partition is θT=(θ1θt)\theta^{T}=(\theta^{\prime}_{1}\geq\cdots\geq\theta^{\prime}_{t}). Then x𝒳krx\in{\cal X}_{k}^{r} iff θ1++θknr\theta^{\prime}_{1}+\ldots+\theta^{\prime}_{k}\geq n-r.

We state our main theorem.

Theorem 7.22

Let xx be a diagonalizable matrix in End(X)End(X) with spectrum partition λ\lambda. The projective orbit closure of xx under the conjugation action by GL(X)GL(X) in (End(X)){\mathbb{P}}(End(X)) contains nilpotent matrices whose partition signature is θ\theta iff θλT\theta\unlhd\lambda^{T}.

Proof Let yy be a nilpotent matrix with partition signature θ\theta such that θλT\theta\not\!\!\unlhd~{}\lambda^{T}. This implies that λθT=(θ1θs)\lambda\not\!\!\unlhd~{}\theta^{T}=(\theta^{\prime}_{1}\geq\cdots\geq\theta^{\prime}_{s}). As a result, there is an index kk such that

j=1kλj>j=1kθj\sum_{j=1}^{k}\lambda_{j}>\sum_{j=1}^{k}\theta^{\prime}_{j}

Now we choose rr such that j=1kλj=nr\sum_{j=1}^{k}\lambda_{j}=n-r and apply Propositions 7.19 and 7.21 to conclude that x𝒳krx\in{\cal X}_{k}^{r} and y𝒳kry\not\in{\cal X}_{k}^{r}. This allows us to conclude that yy is not in the projective orbit closure of xx.

In view of Remark 7.17, it only remains to show that a nilpotent matrix JJ with partition signature λT=(λ1λs)\lambda^{T}=(\lambda^{\prime}_{1}\geq\cdots\geq\lambda^{\prime}_{s}) belongs to the projective orbit closure of xx. Towards this, we can easily adapt the argument in the proof of Theorem 7.7. by writing xx as a block diagonal matrix x1xsx_{1}\oplus\cdots\oplus x_{s} where diagonal matrix xix_{i} is of size λi\lambda^{\prime}_{i} and has distinct eigenvalues, namely those eigenvalues of xx which have multiplicity at least ii. We skip the straightforward details. \Box

Now we extend the above theorem to all matrices. Towards this, fix a matrix xx in a Jordan canonical form with ss distinct eigenvalues μ1,μ2,,μs\mu_{1},\mu_{2},\ldots,\mu_{s}. For eigenvalue μi\mu_{i}, we associate a partition λi=(λi1λi2)\lambda_{i}=(\lambda_{i1}\geq\lambda_{i2}\geq\ldots) which records the sizes of Jordan blocks with diagonal entries μi\mu_{i}, in non-increasing order. We now define the transpose block-spectrum partition of xx to be the partition χ=(χ1χ2)\chi=(\chi_{1}\geq\chi_{2}\geq\ldots) where χj=λ1j+λ2j+λsj\chi_{j}=\lambda_{1j}+\lambda_{2j}+\ldots\lambda_{sj}. Thus, χ\chi is simply the “sum” of the partitions λ1,,λs\lambda_{1},\ldots,\lambda_{s}. Note that when xx is a diagonal matrix, its transpose block-spectrum partition is nothing but the transpose of its spectrum partition.

Theorem 7.23

Let xx be a matrix in End(X)End(X) with transpose block-spectrum partition χ\chi. The projective orbit closure of xx under the conjugation action by GL(X)GL(X) in (End(X)){\mathbb{P}}(End(X)) contains nilpotent matrices whose partition signature is θ\theta iff θχ\theta\unlhd\chi.

Proof Let yy be a nilpotent matrix with partition signature θ=(θ1θ2)\theta=(\theta_{1}\geq\theta_{2}\geq\ldots) such that θχ\theta\not\!\!\unlhd~{}\chi. This means, there is an index \ell with

j=1θj>j=1χj\sum_{j=1}^{\ell}\theta_{j}>\sum_{j=1}^{\ell}\chi_{j}

Henceforth we fix the least index \ell with this property. This further implies that θ>χ\theta_{\ell}>\chi_{\ell}.

Now we set k=χ+1k=\chi_{\ell+1}, r=i=1(χiχ+1)r=\sum_{i=1}^{\ell}(\chi_{i}-\chi_{\ell+1}) and claim that x𝒳krx\in{\cal X}_{k}^{r} and y𝒳kry\not\in{\cal X}_{k}^{r}. We first focus on showing that y𝒳kry\not\in{\cal X}_{k}^{r}. By Proposition 7.20, it suffices to show that

i:θi>k(θik)>r\sum_{i:\theta_{i}>k}(\theta_{i}-k)>r

As k=χ+1k=\chi_{\ell+1} and θ>χχ+1\theta_{\ell}>\chi_{\ell}\geq\chi_{\ell+1}, we have

i:θi>k(θik)i=1(θiχ+1)>i=1(χiχ+1)=r\sum_{i:\theta_{i}>k}(\theta_{i}-k)\geq\sum_{i=1}^{\ell}(\theta_{i}-\chi_{\ell+1})>\sum_{i=1}^{\ell}(\chi_{i}-\chi_{\ell+1})=r

Now we turn our focus on showing that x𝒳krx\in{\cal X}_{k}^{r}. We make use of the notation developed for defining χ\chi just before the proof of the theorem. In this notation, we show that

rank((xμ1I)λ1,+1(xμ2I)λ2,+1(xμsI)λs,+1)=r\mathrm{rank}\left((x-\mu_{1}I)^{\lambda_{1,{\ell+1}}}(x-\mu_{2}I)^{\lambda_{2,{\ell+1}}}\ldots(x-\mu_{s}I)^{\lambda_{s,{\ell+1}}}\right)=r

Note that, this would imply that x𝒳krx\in{\cal X}_{k}^{r} as k=χ+1=λ1,+1++λs,+1k=\chi_{\ell+1}=\lambda_{1,{\ell+1}}+\ldots+\lambda_{s,{\ell+1}}. As the generalized eigenspaces corresponding to different eigenvalues are independent and (xμiI)(x-\mu_{i}I) is nilpotent on the generalized eigenspace corresponding to eigenvalue μi\mu_{i} and invertible on other generalized eigenspaces, we can use Proposition 7.20 to conclude that

rank((xμiI)λi,+1)=(λi,1λi,+1)++(λi,λi,+1)=j=1(λi,jλi,+1)\mathrm{rank}\left((x-\mu_{i}I)^{\lambda_{i,{\ell+1}}}\right)=(\lambda_{i,1}-\lambda_{i,\ell+1})+\ldots+(\lambda_{i,\ell}-\lambda_{i,\ell+1})=\sum_{j=1}^{\ell}(\lambda_{i,j}-\lambda_{i,\ell+1})

and hence,

rank((xμ1I)λ1,+1(xμ2I)λ2,+1(xμsI)λs,+1)=i=1sj=1(λi,jλi,+1)=j=1i=1s(λi,jλi,+1)=j=1((i=1sλi,j)(i=1sλi,+1))=j=1(χjχ+1)=r\begin{array}[]{ccl}\mathrm{rank}\left((x-\mu_{1}I)^{\lambda_{1,{\ell+1}}}(x-\mu_{2}I)^{\lambda_{2,{\ell+1}}}\ldots(x-\mu_{s}I)^{\lambda_{s,{\ell+1}}}\right)&=&\sum_{i=1}^{s}\sum_{j=1}^{\ell}(\lambda_{i,j}-\lambda_{i,\ell+1})\\ &=&\sum_{j=1}^{\ell}\sum_{i=1}^{s}(\lambda_{i,j}-\lambda_{i,\ell+1})\\ &=&\sum_{j=1}^{\ell}\left((\sum_{i=1}^{s}\lambda_{i,j})-(\sum_{i=1}^{s}\lambda_{i,\ell+1})\right)\\ &=&\sum_{j=1}^{\ell}(\chi_{j}-\chi_{\ell+1})\\ &=&r\end{array}

We have thus proved that x𝒳krx\in{\cal X}_{k}^{r} and y𝒳kry\not\in{\cal X}_{k}^{r}. This shows that yy is not in the projective orbit closure of xx.

In order to complete the proof, we finally show that a nilpotent matrix of partition signature χ=(χ1,,χs)\chi=(\chi_{1},\ldots,\chi_{s}) belongs to the projective orbit closure of xx. We first write xx as a block diagonal matrix x1xsx_{1}\oplus\cdots\oplus x_{s} where matrix xjx_{j} is of size χj\chi_{j} and xj=y1,jy2,jys,jx_{j}=y_{1,j}\oplus y_{2,j}\cdots\oplus y_{s,j} where yi,jy_{i,j} is a Jordan block of size λi,j\lambda_{i,j} with diagonal entry μi\mu_{i}. The crucial property of each xjx_{j} is that different Jordan blocks in it correspond to different eigenvalues and hence the minimal polynomial of xjx_{j} equals its characteristic polynomial. As a result, each xjx_{j} can be conjugated to obtain xjx^{\prime}_{j} which is a companion matrix. We now consider x=x1xsx^{\prime}=x^{\prime}_{1}\oplus\cdots\oplus x^{\prime}_{s} to which xx can be block-conjugated.

Now we use the 1-PS family A(t)=diagonal(t,t2,,tn)A(t)=\mathrm{diagonal}(t,t^{2},\ldots,t^{n}) and observe that the leading term of A(t).xA(t).x^{\prime} is the matrix t1Jχt^{-1}J_{\chi} where Jχ=Jχ1JχsJ_{\chi}=J_{\chi_{1}}\oplus\ldots\oplus J_{\chi_{s}} is a nilpotent matrix whose partition signature is χ\chi. This shows that JχJ_{\chi} belongs to the projective orbit closure of xx and completes the proof. \Box.

We now conclude this section with an example.

Example 7.24

Consider the following matrices in Jordan canonical forms

x1=[100010001]x2=[110010001]y1=[010001000]y2=[010000000]x_{1}=\left[\begin{array}[]{ccc}1&0&0\\ 0&1&0\\ 0&0&-1\\ \end{array}\right]\>\>\>x_{2}=\left[\begin{array}[]{ccc}1&1&0\\ 0&1&0\\ 0&0&-1\\ \end{array}\right]\>\>\>y_{1}=\left[\begin{array}[]{ccc}0&1&0\\ 0&0&1\\ 0&0&0\\ \end{array}\right]\>\>\>y_{2}=\left[\begin{array}[]{ccc}0&1&0\\ 0&0&0\\ 0&0&0\\ \end{array}\right]

Clearly, x1x_{1} is diagonal and its spectrum partition as well as transpose spectrum partition is (2,1)(2,1). The transpose block-spectrum partition of x2x_{2} is (3)(3). Note that x1x_{1} is in the affine orbit closure of x2x_{2}. Further, y1y_{1} and y2y_{2} are nilpotent matrices with partition signatures (3)(3) and (2,1)(2,1) respectively.

The above results allow us to conclude that the projective orbit closure of x1x_{1} contains y2y_{2} but not y1y_{1}, and the projective orbit closure of x2x_{2} contains both y1y_{1} and y2y_{2}.

8 Local Differential Geometry of Orbits

So far we have explored algebraic techniques which relate properties of the point yy, the tangent vector yTyO(y)\ell\cdot y\in T_{y}O(y) and the ultimate limit xx for paths which arise from 1-PS. A natural question is whether these paths satisfy additional differential geometric properties, or if there are other special paths which take us to specific limit points xx and if the tangent vectors of exit can be determined using the local geometry at yy. We formulate this as the following path problem.

Definition 8.1

Given a point yVy\in V and an xO(y)¯x\in\overline{O(y)}, the path problem is to determine a path π\pi from yy to xx satisfying specific properties derived from those of yy and xx.

This question is intriguing even when the orbit of yy is affine and we wish to reach representative points of codimension 11 components, i.e., the boundary of the orbit O(y)VO(y)\subseteq\mathbb{P}V. Specifically, is there an optimization problem formulation to arrive at the codimension 11 components and the tangents of exit from yy leading to these components.

An important step in this direction would be to develop the local manifold geometry of the orbit. We have already seen in Remark 2.11, that the function Φ:𝒮×N𝒮\Phi:{\cal S}\times N\to{\cal S} of the local model is intimately connected to the 2-form Π\Pi. However, one technical issue is that our interest lies in the closure in projective space. In Lemma 8.10 we extend Remark 2.11 to compute the Riemannian curvature tensor on a chart of the projective space containing yy. This is developed further in Section 8.3, where we combine the optimization approach and the computation of the curvatures for a specific problem.

Let us begin the discussion with the familiar 1-PS situation with a GL(X)GL(X)-module VV, and x,yVx,y\in V with the action of a 1-PS λ(t)\lambda(t) as shown below:

λ(t)y=xata+xbtb++xdtd\lambda(t)\cdot y=x_{a}\cdot t^{a}+x_{b}t^{b}+\ldots+x_{d}t^{d}

Here x=xax=x_{a} is the projective limit of yy via the 1-PS λ(t)\lambda(t). We will call aa as the degree of the limit xx under λ\lambda. If =logt(λ(t))\ell=\log_{t}(\lambda(t)) then y\ell\cdot y is the tangent of exit of the curve λ(t)y\lambda(t)\cdot y as it exits yy and approaches its limit x=xax=x_{a}. We see how the two may be related by a smooth optimization problem.

8.1 Degree and the structure of the orbit closure

In this subsection, O(y)O(y) will denote the SL(X)SL(X)-orbit of yy, O(y)¯\overline{O(y)} its closure in VV, and O(y)¯V\mathbb{P}\overline{O(y)}\subset\mathbb{P}V will denote the projective closure of the orbit O(y)O(y). The structure of O(y)¯\mathbb{P}\overline{O(y)} consists of closures O(x)¯\mathbb{P}\overline{O(x)} of various lower dimensional orbits O(x)O(x) of limit points xx of the orbit O(y)O(y). If O(y)O(y) is closed in VV and the stabilizer KK of yy is reductive, then in fact, the set O(y)¯O(y)\mathbb{P}\overline{O(y)}-\mathbb{P}O(y) is the union of a finite number of co-dimension 1 projective varieties O(xi)¯\mathbb{P}\overline{O(x_{i})}. This offers the first decomposition of the limit points of O(y)O(y).

And yet, while the boundary of O(y)\mathbb{P}O(y) may be of small co-dimension, the limit xx of a typical 1-PS λ(t)\lambda(t) acting on yy may be embedded deep within this boundary as the following examples show.

Example 8.2

Let AA be an n×nn\times n-matrix such that A(1,n)0A(1,n)\neq 0. Let λ(t)\lambda(t) be the 1-PS consisting of diagonal matrices diag([1,t,,tn1])diag([1,t,\ldots,t^{n-1}]). Then the projective limit of λ(t)Aλ(t)1\lambda(t)A\lambda(t)^{-1} is the matrix E1nE_{1n} with degree (n1)-(n-1). On the other hand, we know that if AA has distinct eigenvalues then the orbit O(Jn)¯\mathbb{P}\overline{O(J_{n})} of JnJ_{n} is of co-dimension 1 within O(A))¯\mathbb{P}\overline{O(A))}. The point JnJ_{n} is approached by a special 1-PS which is conjugate to λ(t)\lambda(t) and where the degree of JnJ_{n} is 1-1.

Example 8.3

Consider the variables X={x1,,x9}X=\{x_{1},\ldots,x_{9}\} and the form det3(X)Sym3(X)det_{3}(X)\in Sym^{3}(X). The dimension of the stabilizer KGL(X)K\subseteq GL(X) of det3det_{3} is 1616. There is another basis 𝒳={X1,,X9}{\cal X}=\{X_{1},\ldots,X_{9}\} of X\mathbb{C}X and a 1-PS λ2(t)\lambda_{2}(t) with the following properties ( see 5.2 for details). Let Y={X1,X2,X3}Y=\{X_{1},X_{2},X_{3}\} and Z={X4,,X9}Z=\{X_{4},\ldots,X_{9}\}. Let λ2(t)\lambda_{2}(t) be such that λ2(t)(y)=y\lambda_{2}(t)(y)=y for yYy\in Y, while λ2(t)(z)=tz\lambda_{2}(t)(z)=tz for all tZt\in Z. Then λ2det3(𝒳)=tQ2(𝒳)+t3Q3(𝒳)\lambda_{2}\cdot det_{3}({\cal X})=tQ_{2}({\cal X})+t^{3}Q_{3}({\cal X}), where Q2(𝒳)=X12X4+X22X5+X32X6+X1X2X7+X2X3X8+X1X3X9Q_{2}({\cal X})=X_{1}^{2}X_{4}+X_{2}^{2}X_{5}+X_{3}^{2}X_{6}+X_{1}X_{2}X_{7}+X_{2}X_{3}X_{8}+X_{1}X_{3}X_{9}. Moreover, the degree of Q2Q_{2} is 11 and its orbit is one of the co-dimension 1 forms in the projective closure of the orbit of det3det_{3}, and the dimension of its stabilizer is 17.

On the other hand, for λ2(t)\lambda^{\prime}_{2}(t) defined as λ2(t)xi=xi\lambda^{\prime}_{2}(t)x_{i}=x_{i} for i=1,2,3i=1,2,3 and λ2(t)(xj)=txj\lambda^{\prime}_{2}(t)(x_{j})=tx_{j}, for j=4,,9j=4,\ldots,9, we have the limit Q2=x1x2x3Q^{\prime}_{2}=x_{1}x_{2}x_{3} of degree 0. The stabilizer of Q2Q^{\prime}_{2} is of dimension 1919. Indeed, for most μ(t)\mu(t) conjugate to λ2\lambda_{2}, we will have μ(t)(det3(X))=t0P0+t1P1+t2P2+t3P3\mu(t)(det_{3}(X))=t^{0}P_{0}+t^{1}P_{1}+t^{2}P_{2}+t^{3}P_{3}, i.e., with the degree of the leading term P0P_{0} as zero, which is lower than that of Q2Q_{2}. Moreover, the dimension of the orbit of this generic limit is lower than that of Q2Q_{2}.

Thus, an important problem is to locate directions, which in the algebraic case, are 1-PS, to specific limit points. A foundation for such analysis was provided by Kempf, [Kem78], in the case when λ(t)SL(X)\lambda(t)\subseteq SL(X) and the degree aa at which the limit is found is non-negative. The case of a>0a>0 corresponds to yy being in the nullcone with the limit x=0x=0. The case a=0a=0 corresponds to a limit x0x\neq 0, with the orbit of O(x)O(x) lying in the affine closure of the SL(X)SL(X)-orbit of yy.

A key construction in Kempf’s analysis is the computation of an optimal 1-PS λ(t)\lambda(t) within a fixed torus TT. The uniqueness of this λ\lambda within one torus allows its comparsion with optimal points in other tori, and the identification of a canonical parabolic subgroup P(λ)GL(X)P(\lambda)\subseteq GL(X) to be associated with yy and xx. The first computation has a generalization to projective limits, which we outline below.

Let us fix a maximal torus TGL(X)T\subseteq GL(X). For simplicity, let us assume that it is the space of invertible diagonal matrices DnD_{n}. Let 𝒟n{\cal D}_{n} be the Lie algebra of DnD_{n}. Let

On2={pn such that |ipi2=1 and ipi=0}O_{n-2}=\{p\in\mathbb{R}^{n}\mbox{ such that }|\sum_{i}p_{i}^{2}=1\mbox{ and }\sum_{i}p_{i}=0\}

For any pnp\in\mathbb{R}^{n}, let us denote by p¯\overline{p} the n×nn\times n diagonal matrix with pp as its diagonal. We regard the action of p¯\overline{p} as an element of gl(x)gl(x), the Lie algebra of GL(X)GL(X). Let tpt^{p} denote the diagonal matrix with diagonal [tp1,,tpn][t^{p_{1}},\ldots,t^{p_{n}}]. Note that for pOn2p\in O_{n-2}, we have p¯𝒟nsl(X)\overline{p}\in{\cal D}_{n}\subseteq sl(X) and tpDnSL(X)t^{p}\in D_{n}\subseteq SL(X).

Let Ξ𝒟n\Xi\subseteq{\cal D}_{n}^{*} be the weight-space of the module VV under TT. For any vVv\in V let v=χΞvχv=\sum_{\chi\in\Xi}v_{\chi} be the decomposition of vv by this weight-space. Let Ξ(v)\Xi(v) be those χΞ\chi\in\Xi for which vχ0v_{\chi}\neq 0. Thus Ξ(v)\Xi(v) is the support of vv. To simplify notation we omit the subscripts in 𝒟n,Dn{\cal D}_{n},D_{n}.

Lemma 8.4

For any element 𝒟\ell\in{\cal D}, we have:

v=χ,χvχ via the Lie algebra actiontv=χt,χvχ via the group action\begin{array}[]{rcll}\ell\cdot v&=&\sum_{\chi}\langle\ell,\chi\rangle v_{\chi}&\mbox{ via the Lie algebra action}\\ t^{\ell}\cdot v&=&\sum_{\chi}t^{\langle\ell,\chi\rangle}v_{\chi}&\mbox{ via the group action}\\ \end{array}

We now state Kempf’s basic optimization result.

Proposition 8.5

(Kempf) We say that vVv\in V is the the null-cone (with respect to 𝒟{\cal D}), if there is an 𝒟\ell\in{\cal D} such that for all χΞ(v)\chi\in\Xi(v) we have ,χ>0\langle\ell,\chi\rangle>0. Let μ(,v)=min{,χ|χΞ(v)}\mu(\ell,v)=\min\{\langle\ell,\chi\rangle|\chi\in\Xi(v)\}. If vv is unstable then there is a unique 0sl(X)\ell_{0}\in sl(X) such that μ(0,v)>0\mu(\ell_{0},v)>0 and μ(0,v)>μ(,v)\mu(\ell_{0},v)>\mu(\ell,v) for all 0\ell\neq\ell_{0}.

We have the following alternate optimization formulation.

Proposition 8.6

Let vVv\in V be arbitrary. Let α(v)={𝒟|μ(,v)>α}{\cal L}_{\alpha}(v)=\{\ell\in{\cal D}|\mu(\ell,v)>\alpha\}. Suppose that α(v){\cal L}_{\alpha}(v) is non-empty. Consider the function

f(t,)=χΞvχ2t,χf(t,\ell)=\sum_{\chi\in\Xi}\|v_{\chi}\|^{2}t^{-\langle\ell,\chi\rangle}

For fixed vVv\in V and tt\in\mathbb{R} the above function is a smooth function on the compact set On2O_{n-2}. Let f(t)\ell_{f}(t) minimize ff on On2O_{n-2}. Then there is a constant t0>1t_{0}>1 depending on vv, such that for all t>t0t>t_{0} we have f(t)α(v)\ell_{f}(t)\in{\cal L}_{\alpha}(v).

Proof: Let Kα(v)\ell_{K}\in{\cal L}_{\alpha}(v) be any element and let μK=μ(K,v)\mu_{K}=\mu(\ell_{K},v). Note that μK>α\mu_{K}>\alpha. Whence, for t>1t>1, we have f(t,K)<v2tμKf(t,\ell_{K})<\|v\|^{2}t^{-\mu_{K}}.

Let Ξ(t)Ξ(v)\Xi_{-}(t)\subseteq\Xi(v) be those χΞ(v)\chi\in\Xi(v) such that f(t),χα\langle\ell_{f}(t),\chi\rangle\leq\alpha. Similarly, let Ξ+(t)\Xi_{+}(t) be those χΞ(v)\chi\in\Xi(v) such that f(t),χ>α\langle\ell_{f}(t),\chi\rangle>\alpha. The claim is that Ξ(t)\Xi_{-}(t) is empty for tt sufficiently large.

Suppose not and there is a ΞΞ(v)\Xi_{-}\subseteq\Xi(v) such that Ξ=Ξ(t)\Xi_{-}=\Xi_{-}(t) for a divergent sequence (ti)i=1(t_{i})_{i=1}^{\infty} with ti>1t_{i}>1 for all ii. Then there is a νΞΞ(v)\nu\in\Xi_{-}\subseteq\Xi(v) such that f(ti),να\langle\ell_{f}(t_{i}),\nu\rangle\leq\alpha for all ii. Whence f(ti,i)vν2tα>0f(t_{i},\ell_{i})\geq\|v_{\nu}\|^{2}t^{-\alpha}>0 for all tit_{i}. But f(t,K)<v2tμKf(t,\ell_{K})<\|v\|^{2}t^{-\mu_{K}} for all tt. Thus we have:

vν2tαf(ti,i)f(ti,K)<v2tiμK\|v_{\nu}\|^{2}t^{-\alpha}\leq f(t_{i},\ell_{i})\leq f(t_{i},\ell_{K})<\|v\|^{2}t_{i}^{-\mu_{K}}

This gives us tiμkα<v2/vν2t_{i}^{\mu_{k}-\alpha}<\|v\|^{2}/\|v_{\nu}\|^{2}, with μK>α\mu_{K}>\alpha.

This is a contradiction since we can choose a tit_{i} large enough for this to not hold. This proves the proposition. \Box

Remark 8.7

For any α(v)\ell\in{\cal L}_{\alpha}(v), let α0=μ(,v)\alpha_{0}=\mu(\ell,v) and Ξ={χΞ(v)|,χ=α0}\Xi_{\ell}=\{\chi\in\Xi(v)|\langle\ell,\chi\rangle=\alpha_{0}\}. Let v=χΞvχv_{\ell}=\sum_{\chi\in\Xi_{\ell}}v_{\chi}. Then vv_{\ell} is a projective limit of vv. We do see that larger values of α\alpha do lead to more fundamental limit points of vv.

Thus, if xx is the minimum degree limit for the point yy, and is approachable by a 1-PS within TT, then the optimization function f(t,)f(t,\ell) on 𝒟sl(X){\cal D}\subseteq sl(X) does indeed offer an algorithm to discover this. However, in the general case, this must be done across all tori within sl(X)=𝒢sl(X)={\cal G} in a unified manner. An example where this has been achieved is in designing efficient algorithms to decide intersections of orbit closures in the space (Matn)k(Mat_{n})^{k}, i.e., kk-tuples of matrices with GLn×GLnGL_{n}\times GL_{n} acting diagonally on the the left and right.  [AZGL+18] formulate this problem as a geodesic convex optimization problem on the space of positive definite matrices with their natural hyperbolic metric.

Thus, what is needed are global optimization functions f~(𝔤,y)\tilde{f}(\mathfrak{g},y), where 𝔤𝒢\mathfrak{g}\in{\cal G}. or more specifically on 𝒢/𝒦{\cal G}/{\cal K} where 𝒦{\cal K} is the stabilizer of yy. Since 𝒢/𝒴{\cal G}/{\cal Y} is the tangent space of the GL(X)GL(X)-orbit of yy, with a 𝒦{\cal K}-structure, the differential geometric structure of this space needs closer examination. The Riemannian approach, and its associated tensors and the geometry of geodesic paths offers an alternate analysis of this algebraic object. We begin with curvature of the orbit O(y)O(y) in the next subsection.

8.2 The Curvature Tensors

We now look at the differential geometry of yy as a point in V\mathbb{P}V.

Remark 2.11 already connects the algebraic gadgets θ\theta and Φ\Phi with the local 2-form Π\Pi of the orbit O(y)VO(y)\subset V. Thus, it is worthwhile to investigate if there is a differential geometric equivalent of the triple stabilizer condition. The hope is that the local geometry may guide us to identify potential tangent vectors and standard paths (e.g., along 1-PS or geodesics) which lead us to various limit points.

However, one difficulty in pursuing this further is that while the Lie algebra acts on VV, the topology and the computation of the limit happens in V\mathbb{P}V. Whence, we must develop a local chart 𝒞{\cal C} for the point yy as a member of V\mathbb{P}V and transfer the action of 𝒢{\cal G} to this chart.

The approach we adopt here is to restrict this computation to the real case. Whence, we assume that VV is equipped with a positive definite metric ,\langle\cdot,\cdot\rangle. Moreover, let us assume that a basis for VV is chosen such that v,v=vTv\langle v,v^{\prime}\rangle=v^{T}v^{\prime}, for any v,vVv,v^{\prime}\in V (treated as column vectors).

The chart 𝒞{\cal C} is constructed from the ball Bϵ(y)B_{\epsilon}(y) as follows:

𝒞={vBϵ(y) such that v=y}{\cal C}=\{v\in B_{\epsilon}(y)\mbox{ such that }\|v\|=\|y\|\}

We define the projection π𝒞:Bϵ(y)𝒞\pi_{\cal C}:B_{\epsilon}(y)\rightarrow{\cal C} as

π𝒞(v)=yvv\pi_{\cal C}(v)=\frac{\|y\|}{\|v\|}v
Definition 8.8

We say that the representation ρ:GL(X)GL(V)\rho:GL(X)\rightarrow GL(V) is of (A) degree zero if IX\mathbb{R}^{*}I_{X} is in the kernel of this map. Otherwise, (B) we say that ρ\rho has a non-zero degree. In this case, and if VV is irreducible, then there is an integer d0d\neq 0 such ρ(tIX)=tdIV\rho(tI_{X})=t^{d}I_{V}.

We say that a point vv is simple if in the case (A) above, the orbit O(y)O(y) does not contain any other multiple of yy, or in the case (B) above, the condition 𝔤𝒢\mathfrak{g}\in{\cal G} and 𝔤vv\mathfrak{g}\cdot v\in\mathbb{R}v implies that 𝔤IX\mathfrak{g}\in\mathbb{R}I_{X}. In this section, in case (B), we restrict 𝒢{\cal G} to sl(X)sl(X) and the orbit O(v)O(v) to be the SL(X)SL(X)-orbit of vv.

Let us define O~(y)\tilde{O}(y) as the image of the projection of O(y)Bϵ(y)O(y)\cap B_{\epsilon}(y) onto 𝒞{\cal C}. We have the following simple lemma:

Lemma 8.9

Suppose that yy is a simple point. Then we may choose ϵ>0\epsilon>0 such that all points vBϵ(y)v\in B_{\epsilon}(y) are also simple. Then the projection π𝒞:O(y)O~(y)\pi_{\cal C}:O(y)\rightarrow\tilde{O}(y) is a diffeomorphism at all points.

Proof: The condition of simplicity is equivalent to v𝒢vv\not\in{\cal G}\cdot v for (A) 𝒢=gl(X){\cal G}=gl(X) or (B) 𝒢=sl(X){\cal G}=sl(X). This is clearly an open condition. The second condition is immediate. \Box

The above lemma allows us to define ρ𝒞:𝒢Vec(𝒞)\rho_{\cal C}:{\cal G}\rightarrow Vec({\cal C}) given by ρ𝒞(𝔤)(v)=π𝒞(ρ(𝔤)(v))\rho_{\cal C}(\mathfrak{g})(v)=\pi_{\cal C}^{*}(\rho(\mathfrak{g})(v)). Our next task is to compute the second fundamental form Π𝒞\Pi_{\cal C} on 𝒞{\cal C}. As before, we will use the vector fields ρ𝒞(𝔰)\rho_{\cal C}(\mathfrak{s}) for 𝔰𝒮\mathfrak{s}\in{\cal S}, to compute this.

Lemma 8.10

Let yO~(y)𝒞Vy\in\tilde{O}(y)\subset{\cal C}\subset V be a simple point with stabilizer 𝒦{\cal K} and complement 𝒮{\cal S} of dimension KK. Let X=ρ(𝔰)(v)=SvX=\rho(\mathfrak{s})(v)=\nabla Sv, be the linear vector field, in the notation of Remark  2.11. For i=1,,Ki=1,\ldots,K, let 𝔰i\mathfrak{s}_{i} be a basis of 𝒮{\cal S}, Xi=ρ(𝔰i)=SivX_{i}=\rho(\mathfrak{s}_{i})=\nabla S_{i}v and Xi~=π𝒞(Xi)=ρ𝒞(𝔰i)\tilde{X_{i}}=\pi^{*}_{\cal C}(X_{i})=\rho_{\cal C}(\mathfrak{s}_{i}). Then:

  1. 1.

    The normal space N~\tilde{N} to O~(y)\tilde{O}(y) within 𝒞{\cal C} is π𝒞(y)(N)\pi^{*}_{\cal C}(y)(N).

  2. 2.

    Let π𝒞(y)\pi^{*}_{\cal C}(y) be the projection at point yy and S~i=π𝒞(y)Si\tilde{S}_{i}=\pi^{*}_{\cal C}(y)S_{i}, then:

    (DX~jX~i)(y)=yTSiyyTyS~jy+S~iS~jy(D_{\tilde{X}_{j}}\tilde{X}_{i})(y)=-\frac{y^{T}S_{i}y}{y^{T}y}\tilde{S}_{j}y+\tilde{S}_{i}\tilde{S}_{j}y
  3. 3.

    Then the second fundamental form Π𝒞\Pi_{\cal C} of O~(y)\tilde{O}(y) as a submanifold of 𝒞{\cal C} is given as follows:

    Π𝒞(X~j,X~i)(y)=λN~(DX~jX~i)(y)=λN~S~iS~jy\Pi_{\cal C}(\tilde{X}_{j},\tilde{X}_{i})(y)=\lambda_{\tilde{N}}(D_{\tilde{X}_{j}}\tilde{X}_{i})(y)=\lambda_{\tilde{N}}\tilde{S}_{i}\tilde{S}_{j}y
  4. 4.

    Finally, if TO(y)=TO~(y)TO(y)=T\tilde{O}(y), i.e., the orbit and its projection osculate at yy, and DXiy,y=0\langle D_{X_{i}}y,y\rangle=0, i.e., yTSiy=0y^{T}S_{i}y=0 for all ii, then:

    Π𝒞(X~j,X~i)(y)=π𝒞(y)(Π(Xj,Xi)(y))\Pi_{\cal C}(\tilde{X}_{j},\tilde{X}_{i})(y)=\pi_{\cal C}^{*}(y)(\Pi(X_{j},X_{i})(y))

Proof: The first part is easily proved by taking up cases (A) and (B) separately.

Next, the field X~i=ρ𝒞(𝔰i)\tilde{X}_{i}=\rho_{\cal C}(\mathfrak{s}_{i}) is given by:

X~i=π𝒞(v)(Siv)=(I1vTvvvT)Siv\tilde{X}_{i}=\pi_{\cal C}^{*}(v)(\nabla S_{i}v)=\nabla(I-\frac{1}{v^{T}v}vv^{T})S_{i}v

Since π𝒞(y)=(I1yTyyyT)\pi^{*}_{\cal C}(y)=(I-\frac{1}{y^{T}y}yy^{T}), we have:

X~j(y)=π𝒞(y)Siy=Si~y\tilde{X}_{j}(y)=\nabla\pi^{*}_{\cal C}(y)S_{i}y=\nabla\tilde{S_{i}}y

We then have:

(DX~jX~i)(y)=π𝒞(y)[[X~j(y)(I1vTvvvT)(y)]Siy+(I1yTyyyT)[X~j(y)(Siv)(y)]]=π𝒞(y)[[2(S~jyTy)yyTSiy1yTy[X~j(y)(vvT)(y)]Siy+π𝒞(y)Siπ𝒞(y)Sjy]]=π𝒞(y)[1yTyS~jyyTSiy1yTyyyTS~jTSiy+S~iS~jy]=π𝒞(y)[yTSiyyTyS~jyyTS~jSiyyTyy+S~iS~jy]=yTSiyyTyS~jy+S~iS~jy\begin{array}[]{rcl}(D_{\tilde{X}_{j}}\tilde{X}_{i})(y)&=&\pi^{*}_{\cal C}(y)\circ[[{\tilde{X}_{j}}(y)(I-\frac{1}{v^{T}v}vv^{T})(y)]S_{i}y+(I-\frac{1}{y^{T}y}yy^{T})[{\tilde{X}_{j}}(y)(S_{i}v)(y)]]\\ &=&\pi^{*}_{\cal C}(y)\circ[[2(\tilde{S}_{j}y^{T}y)yy^{T}S_{i}y-\frac{1}{y^{T}y}[\tilde{X}_{j}(y)(vv^{T})(y)]S_{i}y+\pi^{*}_{\cal C}(y)S_{i}\pi^{*}_{\cal C}(y)S_{j}y]]\\ &=&\pi^{*}_{\cal C}(y)[-\frac{1}{y^{T}y}\tilde{S}_{j}yy^{T}S_{i}y-\frac{1}{yTy}yy^{T}\tilde{S}_{j}^{T}S_{i}y+\tilde{S}_{i}\tilde{S}_{j}y]\\ &=&\pi^{*}_{\cal C}(y)[-\frac{y^{T}S_{i}y}{y^{T}y}\tilde{S}_{j}y-\frac{y^{T}\tilde{S}_{j}S_{i}y}{yTy}y+\tilde{S}_{i}\tilde{S}_{j}y]\\ &=&-\frac{y^{T}S_{i}y}{y^{T}y}\tilde{S}_{j}y+\tilde{S}_{i}\tilde{S}_{j}y\\ \end{array}

This proves (2). Now, since N~=π𝒞(y)(N)\tilde{N}=\pi^{*}_{\cal C}(y)(N), we have S~jyTyO(y)~\tilde{S}_{j}y\in T_{y}\tilde{O(y)}, we have λN~(S~jy)=0\lambda_{\tilde{N}}(\tilde{S}_{j}y)=0. Whence we have:

Π𝒞(X~j,X~i)(y)=λN~(DX~jX~i)(y)=λN~S~iS~jy\Pi_{\cal C}(\tilde{X}_{j},\tilde{X}_{i})(y)=\lambda_{\tilde{N}}(D_{\tilde{X}_{j}}\tilde{X}_{i})(y)=\lambda_{\tilde{N}}\tilde{S}_{i}\tilde{S}_{j}y

This proves (3). In the case when O(y)O(y) and O~(y)\tilde{O}(y) osculate, i.e., TOy=TO~yTO_{y}=T\tilde{O}_{y}, i.e., the tangent planes match, we have S~i=π𝒞(y)Si=Si\tilde{S}_{i}=\pi^{*}_{\cal C}(y)S_{i}=S_{i}, whence

Π𝒞(X~j,X~i)(y)=λN~SiSjy=π𝒞(y)λNSiSjy=π𝒞(Π(Xj,Xi)(y))\Pi_{\cal C}(\tilde{X}_{j},\tilde{X}_{i})(y)=\lambda_{\tilde{N}}S_{i}S_{j}y=\pi^{*}_{\cal C}(y)\lambda_{N}S_{i}S_{j}y=\pi_{\cal C}^{*}(\Pi(X_{j},X_{i})(y))

This proves (4). \Box

We now recall the fundamental result on Riemannian curvature tensor for smooth manifolds.

Proposition 8.11

Let MM be a submanifold of a Riemannian manifold PP. Let Xi,XjX_{i},X_{j} and XkX_{k} be vector fields on MM. Then the Riemannian curvature tensor of MM is given by:

R(Xi,Xj)Xk=D¯XiD¯XjXkD¯XjD¯XiXkD¯[Xi,Xj]XkR(X_{i},X_{j})X_{k}=\overline{D}_{X_{i}}\overline{D}_{X_{j}}X_{k}-\overline{D}_{X_{j}}\overline{D}_{X_{i}}X_{k}-\overline{D}_{[X_{i},X_{j}]}X_{k}

Moreover, if XlX_{l} is another vector field on MM then The quantity rijkl=R(Xi,Xj)(Xk)(x),Xl(x))r_{ijkl}=\langle R(X_{i},X_{j})(X_{k})(x),X_{l}(x))\rangle equals the quantity Π(Xi,Xl)(x),Π(Xj,Xk)(x)Π(Xj,Xl)(x),Π(Xi,Xk)(x)\langle\Pi(X_{i},X_{l})(x),\Pi(X_{j},X_{k})(x)\rangle-\langle\Pi(X_{j},X_{l})(x),\Pi(X_{i},X_{k})(x)\rangle.

The proof of this is standard.

Example 8.12

Let us consider n\mathbb{R}^{n} with the standard basis e1,,ene_{1},\ldots,e_{n} and under the action of SO(n)SO(n) and Lie algebra so(n)so(n). Consider the vector x=(r,0,,0)=re1x=(r,0,\ldots,0)=r\cdot e_{1}, whose orbit is S=rSn1S=rS^{n-1}, the sphere of radius rr in n\mathbb{R}^{n}. Note that in this case, O(x)=O~(x)O(x)=\tilde{O}(x). The tangent space TxST_{x}S is the span of e2,,ene_{2},\ldots,e_{n}. We have n=e1TxS\mathbb{R}^{n}=\mathbb{R}e_{1}\oplus T_{x}S and thus the normal space is only 11-dimensional. Let 𝔰iso(n)\mathfrak{s}_{i}\in so(n) be the element with matrix SiS^{i} with the only non-zero entries as Si(1,i)=1/r,Si(i,1)=1/rS^{i}(1,i)=-1/r,S^{i}(i,1)=1/r. Thus Xi=Siv¯X_{i}=\nabla S^{i}\overline{v} is the corresponding vector field such that Xi(x)=eiX_{i}(x)=e_{i}.

For i=2,,ni=2,\ldots,n and j=2,,nj=2,\ldots,n, we may build that data cjic^{i}_{j}, i.e., the “normal” part of jj-th column of the matrix SiS_{i}. Since the normal part is only 11-dimensional, i.e., the first row, we have:

Π(Xj,Xi)=cji={0if ij1/rif i=j\Pi(X_{j},X_{i})=c^{i}_{j}=\left\{\begin{array}[]{lr}0&\mbox{if $i\neq j$}\\ -1/r&\mbox{if $i=j$}\end{array}\right.

The Ricci curvature RjkcR^{c}_{jk} is the sum below:

Rjkc=irijki=iciicjkcjicikR^{c}_{jk}=\sum_{i}r_{ijki}=\sum_{i}c^{i}_{i}c^{k}_{j}-c^{i}_{j}c^{k}_{i}

This gives us:

Rjkc={0if jk(n1)/r2if j=kR^{c}_{jk}=\left\{\begin{array}[]{lr}0&\mbox{if $j\neq k$}\\ (n-1)/r^{2}&\mbox{if $j=k$}\end{array}\right.
Example 8.13

Let us consider the adjoint action of GLnGL_{n} on MatnnMatn_{n}, i.e., ρ(A)(X)=AXA1\rho(A)(X)=AXA^{-1}, where AGLnA\in GL_{n} and XMatnnX\in Matn_{n}, and the corresponding Lie algebra action of glngl_{n}, given by ρ(a)(X)=aXXa\rho(a)(X)=aX-Xa, where aglna\in gl_{n}.

Let x=diag(λ¯)x=diag(\overline{\lambda}), the diagonal matrix with distinct eigenvalues λ¯=(λ1,,λn)\overline{\lambda}=(\lambda_{1},\ldots,\lambda_{n}). Then =𝒟{\cal H}={\cal D}, the Lie algebra of all diagonal matrices and 𝒮{\cal S} is the linear span of {Eij|ij}\{E_{ij}|i\neq j\}, the non-diagonal elementary matrices. Note that [Eij,x]=(λjλi)x[E_{ij},x]=(\lambda_{j}-\lambda_{i})x and thus 𝒮{\cal S} is both the complement to 𝒟{\cal D} within glngl_{n} and to complement in MatnnMatn_{n} to TxOT_{x}O the tangent space of the orbit. Note that in the standard inner product on matrices, we have Tr([Eij,x]xT)=0Tr([E_{ij},x]x^{T})=0 and thus the point xx satisfies the osculation conditions of Lemma 8.10 (4).

Let eij=Eij/(λjλi)e_{ij}=E_{ij}/(\lambda_{j}-\lambda_{i}) and note that ρ(eij)(x)=eij\rho(e_{ij})(x)=e_{ij}. We now compute Π(ers,epq)=crspq(x)=λNDersρ(epq)(x)\Pi(e_{rs},e_{pq})=c^{pq}_{rs}(x)=\lambda_{N}D_{e_{rs}}\rho(e_{pq})(x). This is given by λN\lambda_{N} of the column vector corresponding to erse_{rs} in the matrix ρ(epq)\rho(e_{pq}). Since NN consists of diagonal matrices, this equals the diagonal entries in the action of epqe_{pq} on MatnMat_{n}, in other words, the diagonal matrix in [epq,ers][e_{pq},e_{rs}]. Thus, the only (r,s)(r,s) for which this happens is (q,p)(q,p) and thus we see that:

crspq={0if (rs)(qp)dpqif (rs)=(qp)c^{pq}_{rs}=\left\{\begin{array}[]{lr}0&\mbox{if $(rs)\neq(qp)$}\\ d_{pq}&\mbox{if $(rs)=(qp)$}\end{array}\right.

where dpqd_{pq} is the diagonal matrix below:

{dpq(p,p)=(λqλp)2dpq(q,q)=(λqλp)2dpq(i,j)=0 for all other tuples (i,j)\left\{\begin{array}[]{ll}d_{pq}(p,p)&=-(\lambda_{q}-\lambda_{p})^{2}\\ d_{pq}(q,q)&=(\lambda_{q}-\lambda_{p})^{2}\\ d_{pq}(i,j)&=0\mbox{ for all other tuples $(i,j)$}\\ \end{array}\right.

Finally, note that dpq,x0\langle d_{pq},x\rangle\neq 0 unles λpλq\lambda_{p}\neq\lambda_{q}, Π𝒞(crspq)\Pi_{\cal C}(c_{rs}^{pq}) will need a final projection which is skipped here.

Example 8.14

Let us consider the case when n=2mn=2m in the example above and

x=[λIm00μIm] with λ,μ0x=\left[\begin{array}[]{cc}\lambda I_{m}&0\\ 0&\mu I_{m}\\ \end{array}\right]\mbox{ with $\lambda,\mu\neq 0$}

Clearly, the stabilizer {\cal H} of xx is the Lie algebra glm×glmgl_{m}\times gl_{m} and 𝒮{\cal S} is the space below:

𝒮={=[0XY0]|X,Yglm}{\cal S}=\left\{=\left[\begin{array}[]{cc}0&X\\ Y&0\\ \end{array}\right]|X,Y\in gl_{m}\right\}

Note that 𝒮=TxO{\cal S}=T_{x}O and that NN is the space of all block-diagonal matrices:

N={=[Z00W]|Z,Wglm}N=\left\{=\left[\begin{array}[]{cc}Z&0\\ 0&W\\ \end{array}\right]|Z,W\in gl_{m}\right\}

We use the {\cal H}-invariant inner product A,B=Tr(AB)\langle A,B\rangle=Tr(AB) on the space glngl_{n}. We decompose 𝒮=𝒮+𝒮{\cal S}={\cal S}^{+}\oplus{\cal S}^{-}, as given below:

𝒮+={=[0X00]|Xglm} and 𝒮={=[00Y0]|Yglm}{\cal S}^{+}=\left\{=\left[\begin{array}[]{cc}0&X\\ 0&0\\ \end{array}\right]|X\in gl_{m}\right\}\mbox{ and }{\cal S}^{-}=\left\{=\left[\begin{array}[]{cc}0&0\\ Y&0\\ \end{array}\right]|Y\in gl_{m}\right\}

For matrices X,Y𝒮±X,Y\in{\cal S}^{\pm}, it is easy to see that Π(X,Y)(x)=λNDXY(x)\Pi(X,Y)(x)=\lambda_{N}D_{X}Y(x) is precisely the block diagonal component of [X,Y][X,Y]. Whence we have:

Π(X,Y)(x)={[XY00YX]when X𝒮+,Y𝒮[YX00XY]when X𝒮,Y𝒮+0otherwise\Pi(X,Y)(x)=\left\{\begin{array}[]{rl}\left[\begin{array}[]{ll}XY&0\\ 0&-YX\end{array}\right]&\mbox{when $X\in{\cal S}^{+},Y\in{\cal S}^{-}$}\\ \vskip 8.53581pt\cr\left[\begin{array}[]{cc}-YX&0\\ 0&XY\end{array}\right]&\mbox{when $X\in{\cal S}^{-},Y\in{\cal S}^{+}$}\\ \vskip 8.53581pt\cr 0&\mbox{otherwise}\end{array}\right.

8.3 The Cyclic Shift Matrix

We now illustrate an optimization formulation of the path problem and its relationship with local curvatures.

Let XX be the space MatnMat_{n} as before acted upon by glngl_{n} my conjugation. Let us examine the case of the approach of a special matrix to its projective limit point expressed in a basis in which the required 1-PS is easily written.

Let n={0,,n1}\mathbb{Z}_{n}=\{0,\ldots,n-1\} be the set of integers under modulo nn addition. This will also be an index set for defining the matrix 𝔠n\mathfrak{c}_{n} (or simply 𝔠)\mathfrak{c}) as the cyclic shift matrix:

𝔠(i,j)={1 if ji=10 otherwise\mathfrak{c}(i,j)=\left\{\begin{array}[]{rl}1&\mbox{ if $j-i=1$}\\ 0&\mbox{ otherwise}\end{array}\right.

Thus 𝔠\mathfrak{c} is the matrix below:

𝔠=[010001000011000]\mathfrak{c}=\left[\begin{array}[]{ccccc}0&1&\ldots&&0\\ 0&0&1&\ldots&0\\ \vdots&&\vdots&&\\ 0&\ldots&0&0&1\\ 1&\ldots&0&0&0\\ \end{array}\right]

Let λ(t)\lambda(t) be the 11-PS below:

λ(t)=[t000t200000tn]\lambda(t)=\left[\begin{array}[]{ccccc}t&0&\ldots&&0\\ 0&t^{2}&0&\ldots&0\\ \vdots&&\vdots&&\\ 0&\ldots&0&0&t^{n}\\ \end{array}\right]

We the see that log(λ(t))\log(\lambda(t)) is the matrix \ell below:

=[1000200000n]\ell=\left[\begin{array}[]{ccccc}1&0&\ldots&&0\\ 0&2&0&\ldots&0\\ \vdots&&\vdots&&\\ 0&\ldots&0&0&n\\ \end{array}\right]

The action of a generic element 𝔞=(aij)\mathfrak{a}=(a_{ij}) of glngl_{n} on 𝔠\mathfrak{c} is given by:

tn=𝔞𝔠=a𝔠𝔠a=[a0,n1a10a00a11a0,n2a1,n1an1,n1a0,0an1,0a0,1an1,n1a0,n1]t_{n}=\mathfrak{a}\cdot\mathfrak{c}=a\cdot\mathfrak{c}-\mathfrak{c}\cdot a=\left[\begin{array}[]{ccccc}a_{0,n-1}-a_{10}&a_{00}-a_{11}&\ldots&&a_{0,n-2}-a_{1,n-1}\\ \vdots&&\vdots&&\\ a_{n-1,n-1}-a_{0,0}&a_{n-1,0}-a_{0,1}&\ldots&&a_{n-1,n-1}-a_{0,n-1}\\ \end{array}\right]

Thus the matrix tnt_{n} is given by:

tn(i,j)=ai,j1ai+1,jt_{n}(i,j)=a_{i,j-1}-a_{i+1,j}

We use the standard inner product on both MatnMat_{n} as well as glngl_{n}, viz., x,y=Tr(xy¯T)\langle x,y\rangle=Tr(x\overline{y}^{T}), for any matrices x,yx,y. Since (𝔠i)T=𝔠ni(\mathfrak{c}^{i})^{T}=\mathfrak{c}^{n-i}, we have the following straightforward lemma:

Lemma 8.15

The tangent space TnXT_{n}X to 𝔠\mathfrak{c} is perpendicular to 𝔠\mathfrak{c}. In other words Tr(tn𝔠T)=0Tr(t_{n}\mathfrak{c}^{T})=0 for all 𝔞gln\mathfrak{a}\in gl_{n}.

The stabilizer {\cal H} of 𝔠\mathfrak{c} is the space i=0n1𝔠i\oplus_{i=0}^{n-1}\mathfrak{c}^{i}. Let us compute the complement 𝒮{\cal S} to {\cal H} under the above inner product. For k=0,,n1k=0,\ldots,n-1, let DkD_{k} be the indices

Dk={(i,i+k)|i=1,,n}D_{k}=\{(i,i+k)|i=1,\ldots,n\}

Thus DkD_{k} is kk-th shifted diagonal, i.e., the support cnkc_{n}^{k}. Thus

𝒮={𝔰=(sij) such that (i,j)Dksij=0 for all k}{\cal S}=\{\mathfrak{s}=(s_{ij})\mbox{ such that }\sum_{(i,j)\in D_{k}}s_{ij}=0\mbox{ for all $k$}\}

Let us look at γ(𝔰)=𝔰𝔠/𝔰\gamma(\mathfrak{s})=\|\mathfrak{s}\cdot\mathfrak{c}\|/\|\mathfrak{s}\|, the compression achieved by 𝔰\mathfrak{s}. We show that \ell above is one of the elements of 𝒮{\cal S} which minimize γ\gamma. For convenience, let us denote by 𝒮k{\cal S}_{k} as the elements (s=(sij)𝒮(s=(s_{ij})\in{\cal S} where ji=kj-i=k, i.e., matrices in 𝒮{\cal S} with support in DkD_{k}. Thus, 𝒮=k=0n1𝒮i{\cal S}=\oplus_{k=0}^{n-1}{\cal S}_{i}.

Let us assume that γ\gamma achieves a minima at 𝔵𝒮\mathfrak{x}\in{\cal S} and that 𝔱=𝔵𝔠\mathfrak{t}=\mathfrak{x}\cdot\mathfrak{c}. We may decompose these elements as above, by their support and express 𝔵\mathfrak{x} and 𝔱\mathfrak{t} as a sum 𝔵=i=0n1𝔵i\mathfrak{x}=\sum_{i=0}^{n-1}\mathfrak{x}_{i} and 𝔱=i=0n1𝔱i\mathfrak{t}=\sum_{i=0}^{n-1}\mathfrak{t}_{i}. We observe that:

γ(𝔵)=𝔱02++𝔱n12𝔵02++𝔵n12\gamma(\mathfrak{x})=\frac{\sqrt{\|\mathfrak{t}_{0}\|^{2}+\ldots+\|\mathfrak{t}_{n-1}\|^{2}}}{\sqrt{\|\mathfrak{x}_{0}\|^{2}+\ldots+\|\mathfrak{x}_{n-1}\|^{2}}}

Since 𝔵i𝔠=𝔱i+1\mathfrak{x}_{i}\cdot\mathfrak{c}=\mathfrak{t}_{i+1}, the minima is also achieved where 𝔵=𝔵i\mathfrak{x}=\mathfrak{x}_{i} for some ii. We choose i=0i=0; the other cases are similar. Thus 𝔵𝒮0\mathfrak{x}\in{\cal S}_{0} is a diagonal matrix (xii)(x_{ii}) such that ixii=0\sum_{i}x_{ii}=0. We see then that:

γ(𝔵)=i=1n(xiixi+1,i+1)2i=1nxii2\gamma(\mathfrak{x})=\frac{\sum_{i=1}^{n}(x_{ii}-x_{i+1,i+1})^{2}}{\sum_{i=1}^{n}x_{ii}^{2}}

The condition that ixii=0\sum_{i}x_{ii}=0 requires that some xiixi+1,i+1x_{ii}\neq x_{i+1,i+1} for some ii. Suppose then that xnnx11=1x_{nn}-x_{11}=1. It is then easily seen that the optimal distribution of the intermediate xiix_{ii} is obtained by evenly spacing them, i.e., xi+1,i+1xii=1/(n1)x_{i+1,i+1}-x_{ii}=1/(n-1), for i=1,,n2i=1,\ldots,n-2.

Recall \ell as before and let ¯\overline{\ell} be obtained by subtracting constant (n+1)2\frac{(n+1)}{2} from each entry of \ell. This makes ¯\overline{\ell} an element of 𝒮{\cal S}. We have the following lemma, which is easily proved.

Lemma 8.16

Let ij=𝔠i¯𝔠j\ell_{ij}=\mathfrak{c}^{i}\overline{\ell}\mathfrak{c}^{-j}.

  • Then ij𝒮ij\ell_{ij}\in{\cal S}_{i-j} and the set Lk={ij|ij=k}L_{k}=\{\ell_{ij}|i-j=k\} is a basis for 𝒮i{\cal S}_{i}. L=kLkL=\cup_{k}L_{k} is a basis for 𝒮{\cal S}.

  • The elements 𝔵𝒮k\mathfrak{x}\in{\cal S}_{k}, for some kk, optimizing γ\gamma above are scalar multiples of some element of LL.

The above lemma is not a completely happy situation. For 𝔯𝒮0\mathfrak{r}\in{\cal S}_{0} with ii=𝔠i¯𝔠i\ell_{ii}=\mathfrak{c}^{i}\overline{\ell}\mathfrak{c}^{-i}, we see that the leading term of ii(𝔠)=[ii,𝔠]=ii𝔠𝔠ii\ell_{ii}(\mathfrak{c})=[\ell_{ii},\mathfrak{c}]=\ell_{ii}\mathfrak{c}-\mathfrak{c}\ell_{ii} is 𝔠iJn𝔠i\mathfrak{c}^{i}J_{n}\mathfrak{c}^{-i}, which is a cyclic renumbering of the nilpotent flag of JnJ_{n} and thus ii\ell_{ii} is not only an optimal tangent of exit, but it also leads to the desired highest dimension limit point. This is the good part.

However, for iji\neq j, ij=𝔠i¯𝔠j𝒮ji\ell_{ij}=\mathfrak{c}^{i}\cdot\overline{\ell}\cdot\mathfrak{c}^{-j}\in{\cal S}_{j-i} does minimize γ\gamma, but we have the following:

LT([ij,𝔠])=ij𝔠𝔠ij=(𝔠iJn𝔠1)cijLT([\ell_{ij},\mathfrak{c}])=\ell_{ij}\mathfrak{c}-\mathfrak{c}\ell_{ij}=(\mathfrak{c}^{i}J_{n}\mathfrak{c}^{-1})c^{i-j}

This limit is not conjugate to JnJ_{n}, and need not even be a nilpotent matrix. Thus the direction ij𝒮ij\ell_{ij}\in{\cal S}_{i-j} is a false start from the point 𝔠\mathfrak{c}. The resolution of this situation lies in the analysis of the curvature form at the point 𝔠\mathfrak{c}. Note that thus form Π:𝒮×𝒮N\Pi:{\cal S}\times{\cal S}\rightarrow N, where we have the basis kLi\cup_{k}L_{i} for 𝒮{\cal S} and {𝔠i|i=0,,n1}\{\mathfrak{c}^{i}|i=0,\ldots,n-1\} for NN.

Our next task will be to find an expression for Π(i,j,k,l),𝔠r\langle\Pi(\ell_{i,j},\ell_{k,l}),\mathfrak{c}^{r}\rangle for general i,j,k,l,ri,j,k,l,r and to evaluate these for a suitable subset. For this, we will use the lemma:

Lemma 8.17

With the above notation, we have:

Π(i,j,k,l),𝔠r=λN(ρ(k,l)(ρ(i,j),(𝔠r))=1nTr([kl,[ij,𝔠1]](𝔠r)T)\langle\Pi(\ell_{i,j},\ell_{k,l}),\mathfrak{c}^{r}\rangle=\langle\lambda_{N}(\rho(\ell_{k,l})(\rho(\ell_{i,j}),(\mathfrak{c}^{r}))\rangle=\frac{1}{n}Tr([\ell_{kl},[\ell_{ij},\mathfrak{c}^{1}]]\cdot(\mathfrak{c}^{r})^{T})

Proof: The first equality follows from Lemma 8.10. For the second equality, note that ρ(k,l)(ρ(i,j)(𝔠r))=[kl,[ij,𝔠1]]\rho(\ell_{k,l})(\rho(\ell_{i,j})(\mathfrak{c}^{r}))=[\ell_{kl},[\ell_{ij},\mathfrak{c}^{1}]]. Next, since {𝔠i|i=0,,n1}\{\mathfrak{c}^{i}|i=0,\ldots,n-1\} is an orthogonal basis in the norm X,Y=Tr(XYT)\langle X,Y\rangle=Tr(XY^{T}), we have that the component along 𝔠r\mathfrak{c}^{r} is precisely ,𝔠r/𝔠r,𝔠r\langle\cdot,\mathfrak{c}^{r}\rangle/\langle\mathfrak{c}^{r},\mathfrak{c}^{r}\rangle. The result follows since 𝔠r,𝔠r=Tr(𝔠r(𝔠r)T)=n\langle\mathfrak{c}^{r},\mathfrak{c}^{r}\rangle=Tr(\mathfrak{c}^{r}(\mathfrak{c}^{r})^{T})=n. \Box

Lemma 8.18

Let Li=𝔠i=0,iL_{i}=\ell\cdot\mathfrak{c}^{i}=\ell_{0,-i} and let:

Π(Li,Lj)=k=0n1pijk𝔠k\Pi(L_{i},L_{j})=\sum_{k=0}^{n-1}p_{ij}^{k}\mathfrak{c}^{k}

Then pijkp_{ij}^{k} are given by:

pijk={(n1)(i+j) when k0 and k=i+j+1 in n0otherwisep_{ij}^{k}=\left\{\begin{array}[]{rl}(n-1)-(i+j)&\mbox{ when $k\neq 0$ and $k=i+j+1$ in $\mathbb{Z}_{n}$}\\ 0&\mbox{otherwise}\end{array}\right.

Proof: Let is use the index set n={0,,n1}\mathbb{Z}_{n}=\{0,\ldots,n-1\}. Let us first note that for any i,ji,j, there is at most one kk for which pijk0p_{ijk}\neq 0. This is simply because Π(Li,Lj)=[Lj,[Li,𝔠1]]\Pi(L_{i},L_{j})=[L_{j},[L_{i},\mathfrak{c}^{1}]] has support only in Di+j+1D_{i+j+1}. Thus, k=i+j+1k=i+j+1. Let us denote the inner term [𝔠i,𝔠1][\ell\mathfrak{c}^{i},\mathfrak{c}^{1}] as Δi\Delta_{i} which is given by:

Δi(r,s)={1if r{0,,n2} and s=r+i+1 in nn1r=n1 and s=r+i+1 in n0otherwise\Delta_{i}(r,s)=\left\{\begin{array}[]{rl}-1&\mbox{if $r\in\{0,\ldots,n-2\}$ and $s=r+i+1$ in $\mathbb{Z}_{n}$}\\ n-1&\mbox{$r=n-1$ and $s=r+i+1$ in $\mathbb{Z}_{n}$}\\ 0&\mbox{otherwise}\\ \end{array}\right.

LjL_{j} itself is given by:

Lj(r,s)={rif r{0,,n1} and s=r+j in n0otherwiseL_{j}(r,s)=\left\{\begin{array}[]{rl}r&\mbox{if $r\in\{0,\ldots,n-1\}$ and $s=r+j$ in $\mathbb{Z}_{n}$}\\ 0&\mbox{otherwise}\\ \end{array}\right.

Then the matrix of interest LjΔiΔiLjL_{j}\Delta_{i}-\Delta_{i}L_{j}. We see that:

Tr(LjΔi(𝔠k)T)=r=0n1(Lj)r,r+j(Δi)r+j,r+k=r=0,r+jn1n1(r)+r(n1)δr+j,n1=r=0n1(r)+r(n)δr+j,n1=n(n1)/2+n(n1j)=n(n1)/2jn\begin{array}[]{rcl}Tr(L_{j}\Delta_{i}(\mathfrak{c}^{k})^{T})&=&\sum_{r=0}^{n-1}(L_{j})_{r,r+j}(\Delta_{i})_{r+j,r+k}\\ &=&\sum_{r=0,r+j\neq n-1}^{n-1}(-r)+r\cdot(n-1)\delta_{r+j,n-1}\\ &=&\sum_{r=0}^{n-1}(-r)+r\cdot(n)\delta_{r+j,n-1}\\ &=&-n(n-1)/2+n(n-1-j)\\ &=&n(n-1)/2-j\cdot n\\ \end{array}

In the same manner, we have:

Tr(ΔiLj(𝔠k)T)=r=0n1(Δi)r,r+i+1(Lj)r+i+1,r+k=r=0n2(Δi)r,r+i+1(Lj)r+i+1,r+k+(n1)(Lj)n+i,n1+k=r=0n2irn2i+1n2(rn)+(n1)i=r=0n2(r+i+1)+r=n2i+1n2(r+i+1n)+(n1)i=(i+1+n+i1)(n1)/2+in+i(n1)=(n1)n/2+in\begin{array}[]{rcl}Tr(\Delta_{i}L_{j}(\mathfrak{c}^{k})^{T})&=&\sum_{r=0}^{n-1}(\Delta_{i})_{r,r+i+1}(L_{j})_{r+i+1,r+k}\\ &=&\sum_{r=0}^{n-2}(\Delta_{i})_{r,r+i+1}(L_{j})_{r+i+1,r+k}+(n-1)(L_{j})_{n+i,n-1+k}\\ &=&-\sum_{r=0}^{n-2-i}r-\sum_{n-2-i+1}^{n-2}(r-n)+(n-1)i\\ &=&-\sum_{r=0}^{n-2}(r+i+1)+\sum_{r=n-2-i+1}^{n-2}(r+i+1-n)+(n-1)i\\ &=&-(i+1+n+i-1)(n-1)/2+in+i(n-1)\\ &=&-(n-1)n/2+in\\ \end{array}

This gives us:

pijk=1nTr((LjΔiΔiLj)(𝔠k)T)=(n1)(i+j)p_{ij}^{k}=\frac{1}{n}Tr((L_{j}\Delta_{i}-\Delta_{i}L_{j})(\mathfrak{c}^{k})^{T})=(n-1)-(i+j)

where k=i+j+1k=i+j+1. This proves the lemma. \Box

Example 8.19

Let Πk\Pi^{k} be the matrices (pijk)(p^{k}_{ij}), then for n=5n=5 we have the following.

P1=[4000000001000100010001000]P2=[0300030000000020002000200]P^{1}=\left[\begin{array}[]{rrrrr}4&0&0&0&0\\ 0&0&0&0&-1\\ 0&0&0&-1&0\\ 0&0&-1&0&0\\ 0&-1&0&0&0\\ \end{array}\right]P^{2}=\left[\begin{array}[]{rrrrr}0&3&0&0&0\\ 3&0&0&0&0\\ 0&0&0&0&-2\\ 0&0&0&-2&0\\ 0&0&-2&0&0\\ \end{array}\right]
P3=[0020002000200000000300030]P4=[0001000100010001000000004]P^{3}=\left[\begin{array}[]{rrrrr}0&0&2&0&0\\ 0&2&0&0&0\\ 2&0&0&0&0\\ 0&0&0&0&-3\\ 0&0&0&-3&0\\ \end{array}\right]P^{4}=\left[\begin{array}[]{rrrrr}0&0&0&1&0\\ 0&0&1&0&0\\ 0&1&0&0&0\\ 1&0&0&0&0\\ 0&0&0&0&-4\\ \end{array}\right]

Note that Π(Li,Lj)\Pi(L_{i},L_{j}) will have no component along 𝔠0\mathfrak{c}^{0}, the identity matrix, simply because the matrix [Li[Lj,𝔠1]][L_{i}[L_{j},\mathfrak{c}^{1}]] will have zero trace. Next, the normal 𝔠1\mathfrak{c}^{1} is indeed special since, the in any local model of the projective orbit O(𝔠)\mathbb{P}O(\mathfrak{c}), the normal 𝔠\mathfrak{c} to O(𝔠)O(\mathfrak{c}) will collapse. We then have the following important lemma:

Lemma 8.20

For the element 𝔠Matn\mathfrak{c}\in Mat_{n}, we have Π(L0,L0)=0\Pi(L_{0},L_{0})=0 for the submanifold O(𝔠)V\mathbb{P}O(\mathfrak{c})\subseteq\mathbb{P}V. Moreover, if nn is odd then of all LiL_{i}, it is unique in having this property.

This analysis does show that the limiting 1-PS is indeed special in the local differential geometry of the starting point yy. However, the exact connection between this and the algebraic conditions which lead us to the limit xx, and its properties, is not clear.

9 Conclusion

Our motivating question has been to understand the conditions under which a point xVx\in V is in the projective orbit closure of yVy\in V, when both xx and yy have distinctive stabilizers, resp. HH and KK, with Lie algebras {\cal H} and 𝒦{\cal K}. Our primary objective has been to provide some initial Lie algebraic tools to analyse this important problem.

Towards this, we develop a local model for xx, i.e., an open neighborhood of xx and a Lie algebra action of 𝒢{\cal G} with certain useful properties. We show that this open neighborhood has a convenient product structure M×NM\times N, where MM corresponds to a neighborhood of the orbit O(x)O(x) identified with 𝒮{\cal S}, a complement to {\cal H} in 𝒢{\cal G}, and NN, a complementary space to TxO(x)VT_{x}O(x)\subseteq V, the tangent space of the orbit. We show that for points y=x+nM×Ny=x+n\in M\times N, the Lie algebra 𝒦{\cal K} of the stabilizer KK of yy is determined by a subspace of {\cal H}. Indeed, 𝒦{\cal K} is the 𝒮{\cal S}-completion of this subspace of {\cal H}, where 𝒮{\cal S} above is a complement to {\cal H} in 𝒢{\cal G}. Moreover, we show that the computation of 𝒦{\cal K} is made effective by a map θ\theta, which measures the “shear” due to 𝒬{\cal Q}, the unipotent radical of {\cal H}, and its close relative Φ\Phi, which is intimately connected with the curvature of the orbit at xx. This alignment between {\cal H} and 𝒦{\cal K} is an extension of Luna’s results on reductive stabilizers, but at the Lie algebra level.

Our main application is on forms, i.e., on the space V=Symd(X)V=Sym^{d}(X), and points x,yVx,y\in V, which have distinctive stabilizers, and where xx is obtained from yy by taking limits through 1-parameter subgroups λ(t)\lambda(t) of GL(X)GL(X). We show that that there is a leading term subalgebra 𝒦0{\cal K}_{0} of 𝒦{\cal K}, the stabilizer of yy, which appears as a subalgebra of {\cal H}, the stabilizer of xx. We also show that a generic λ\lambda leads to solvable 𝒦0{\cal K}_{0}, while non-generic λ\lambda connect the form xx, the tangent of approach, the tangent of exit and the form yy through triple-stabilizers.

The connection between 𝒦0{\cal K}_{0} and 𝒦{\cal K} is explored further by considering 𝒦{\cal K} as an extension of 𝒦0{\cal K}_{0}. Under a transversality assumption, there is an extension 𝒦(ϵ){\cal K}(\epsilon) of 𝒦0{\cal K}_{0} whose structure constants agree with that of 𝒦{\cal K} up to the first order. This leads us to derivations of Lie algebras and certain Lie cohomology conditions. These tools are applied to the co-dimension 1 subvarieties within orbit closures.

The local model is next applied to the celebrated problem of computing projective closures of orbits of a single matrix under conjugation. Here we reprove the classical result of [Ger61, Hes79] on the orbit closure of a nilpotent matrix, as well as prove, perhaps for the first time, a result on the projective closures of a general semisimple matrix.

Finally, we consider the path problem, of the choice of paths which lead out of yy and their limit points xx. We show how Kempf’s algebraic construction for a torus does have an unconstrained optimization formulation, and thus points to local gradient descent schemes at the local tangent space of yy. We then connect this with the map Φ\Phi above and compute curvatures at yy and show, through an example, that these may indicate good choices of paths out of yy.

Coming back to GCT, our motivation for the problems considered in this paper, identifying linkages between forms and their stabilizer groups has been an important theme in the GCT approach. The expectation is that such linkages would lead to obstructions, or witnesses of why a form gg with stabilizer HH, cannot be in the orbit closure of another form, ff, with stabilizer KK, where both HH and KK are distinctive. Indeed, in [MS01] and [MS08] the authors proposed representation theoretic obstructions which arose from a generalization of the Peter-Weyl theorem to this situation. These were GG-modules in coordinate rings of the orbit closure of gg and ff, and the obstruction arose from the multiplicities of these modules in the respective rings. Whether these multiplicities obstructions will be enough to separate gg from ff has been intensely analysed.

Perhaps the local model allows for another source of obstructions connecting {\cal H} and 𝒦{\cal K}, the corresponding Lie algebra stabilizers. If indeed gg is obtained as a limit of ff, then the local model gives us two intermediate structures which link {\cal H} and 𝒦{\cal K}. The first is a “tangent of approach”, a form fbf_{b}, along which ff approaches gg, and the tangent of exit f\ell f. The second is the Lie algebra 𝒦0{\cal K}_{0}, which is obtained from 𝒦{\cal K} through a limiting process. 𝒦0{\cal K}_{0} is also a sublagebra of {\cal H}, which stabilizes the vector fb¯\overline{f_{b}} in the “quotient” representation V/TOgV/TO_{g} of {\cal H}. This also requires an alignment in certain Lie algebra cohomology modules, which may be the place for potential obstructions. We believe that these obstructions may be especially computable when the orbit of gg is expected to be within the orbit of ff as a sub-variety of codimension one.

References

  • [AM95] DV Alekseevsky and Peter W Michor. Differential geometry of g-manifolds. Differential Geometry and its Applications, 5(4):371–403, 1995.
  • [AZGL+18] Zeyuan Allen-Zhu, Ankit Garg, Yuanzhi Li, Rafael Oliveira, and Avi Wigderson. Operator scaling via geodesically convex optimization, invariant theory and polynomial identity testing. In Proceedings of the 50th Annual ACM SIGACT Symposium on Theory of Computing, pages 172–181, 2018.
  • [BGO+18] Peter Bürgisser, Ankit Garg, Rafael Oliveira, Michael Walter, and Avi Wigderson. Alternating minimization, scaling algorithms, and the null-cone problem from invariant theory. In 9th Innovations in Theoretical Computer Science Conference (ITCS 2018). Schloss Dagstuhl-Leibniz-Zentrum fuer Informatik, 2018.
  • [BIP19] Peter Bürgisser, Christian Ikenmeyer, and Greta Panova. No occurrence obstructions in Geometric Complexity Theory. Journal of the American Mathematical Society, 32(1):163–193, 2019.
  • [BLMW11] Peter Bürgisser, Joseph M Landsberg, Laurent Manivel, and Jerzy Weyman. An overview of mathematical issues arising in the Geometric Complexity Theory approach to VP \neq VNP. SIAM Journal on Computing, 40(4):1179–1209, 2011.
  • [DM17] Harm Derksen and Visu Makam. Polynomial degree bounds for matrix semi-invariants. Advances in Mathematics, 310:44–63, 2017.
  • [Dol03] Igor Dolgachev. Lectures on invariant theory. Number 296. Cambridge University Press, 2003.
  • [Ger61] Murray Gerstenhaber. Dominance over the classical groups. Annals of Mathematics, pages 532–569, 1961.
  • [GGOW19] Ankit Garg, Leonid Gurvits, Rafael Oliveira, and Avi Wigderson. Operator scaling: theory and applications. Foundations of Computational Mathematics, pages 1–68, 2019.
  • [Gro15] Joshua A Grochow. Unifying known lower bounds via Geometric Complexity Theory. computational complexity, 24(2):393–475, 2015.
  • [Hes76] Wim Hesselink. Singularities in the nilpotent scheme of a classical group. Transactions of the american mathematical society, 222:1–32, 1976.
  • [Hes79] Wim H Hesselink. Desingularizations of varieties of nullforms. Inventiones mathematicae, 55(2):141–163, 1979.
  • [Hil93] David Hilbert. Über die vollen Invariantensysteme. Mathematische annalen, 42(3):313–373, 1893.
  • [HL16] Jesko Hüttenhain and Pierre Lairez. The boundary of the orbit of the 3-by-3 determinant polynomial. Comptes rendus Mathematique, 354(9):931–935, 2016.
  • [Hof65] Karl Heinrich Hofmann. Lie algebras with subalgebras of co-dimension one. Illinois Journal of Mathematics, 9(4):636–643, 1965.
  • [Hüt17] Jesko Hüttenhain. Geometric Complexity Theory and orbit closures of homogeneous forms. Ph.D Thesis,Technichen Universita¨\ddot{a}t, Berlin, 2017.
  • [IP17] Christian Ikenmeyer and Greta Panova. Rectangular Kronecker coefficients and plethysms in Geometric Complexity Theory. Advances in Mathematics, 319:40–66, 2017.
  • [IQS17] Gábor Ivanyos, Youming Qiao, and KV Subrahmanyam. Non-commutative Edmonds’ problem and matrix semi-invariants. computational complexity, 26(3):717–763, 2017.
  • [Kem78] George R Kempf. Instability in invariant theory. Annals of Mathematics, 108(2):299–316, 1978.
  • [Lan15] Joseph M Landsberg. Geometric Complexity Theory: an introduction for geometers. Annali dell’universita’di Ferrara, 61(1):65–117, 2015.
  • [Lee01] Dong Hoon Lee. The structure of complex Lie groups. CRC Press, 2001.
  • [LL89] Thomas Lehmkuhl and Thomas Lickteig. On the order of approximation in approximative triadic decompositions of tensors. Theoretical computer science, 66(1):1–14, 1989.
  • [Lun73] Domingo Luna. Slices étales. Bull. Soc. Math. France, 33:81–105, 1973.
  • [Mat60] Yozô Matsushima. Espaces homogenes de Stein des groupes de Lie complexes. Nagoya Mathematical Journal, 16:205–218, 1960.
  • [MFK94] David Mumford, John Fogarty, and Frances Kirwan. Geometric Invariant Theory, volume 34. Springer Science & Business Media, 1994.
  • [Mic08] Peter W Michor. Topics in differential geometry, volume 93. American Mathematical Soc., 2008.
  • [MLR13] Laurent Manivel, Joseph M Landsberg, and Nicolas Ressayre. Hypersurfaces with degenerate duals and the Geometric Complexity Theory program. Commentarii Mathematici Helvetici, 88(2):469–484, 2013.
  • [MR04] Thierry Mignon and Nicolas Ressayre. A quadratic bound for the determinant and permanent problem. International Mathematics Research Notices, 2004(79):4241–4253, 2004.
  • [MS01] Ketan D Mulmuley and Milind Sohoni. Geometric Complexity Theory i: An approach to the P vs. NP and related problems. SIAM Journal on Computing, 31(2):496–526, 2001.
  • [MS08] Ketan D Mulmuley and Milind Sohoni. Geometric Complexity Theory ii: Towards explicit obstructions for embeddings among class varieties. SIAM Journal on Computing, 38(3):1175–1206, 2008.
  • [New78] Peter E Newstead. Introduction to moduli problems. Tata Inst. Lecture Notes, 1978.
  • [NR67] Albert Nijenhuis and R.W. Richardson. Deformations of lie algebra structures. Journal of Mathematics and Mechanics, 17(1):89–105, 1967.
  • [Pop03] Vladimir Leonidovich Popov. The cone of Hilbert nullforms. Trudy Matematicheskogo Instituta imeni VA Steklova, 241:192–209, 2003.
  • [Pop09] Vladimir L Popov. Two orbits: When is one in the closure of the other? Proceedings of the Steklov Institute of Mathematics, 264(1):146–158, 2009.
  • [Val79] L. G. Valiant. Completeness classes in algebra. In Proceedings of the Eleventh Annual ACM Symposium on Theory of Computing, STOC ’79, page 249–261, New York, NY, USA, 1979. Association for Computing Machinery.
  • [Wag10] Fredrich Wagemann. Introduction to Lie algebra cohomology with a view towards BRST cohomology. Lecture notes, Université Nantes, 2010.

Appendix A Algebraic complexity and the geometric complexity theory - a brief introduction

Let YY denote a vector of indeterminates suitably arranged, (e.g., as an m×mm\times m-matrix) and let Symd(Y)Sym^{d}(Y) be the space of forms of degree dd. A natural model to compute forms are straight-line programs (also known as arithmetic circuits). Such a program consists of an ordered sequence of arithmetic operations involving +,,+,-,* (called lines), starting with variables yiYy_{i}\in Y and constants from {\mathbb{C}}. A form fSymd(Y)f\in Sym^{d}(Y) is said to have an efficient computation if there is straight-line program computing ff in which the number of lines is bounded by a polynomial in |Y||Y| and dd. The determinant form detm(Y)Symm(Y)det_{m}(Y)\in Sym^{m}(Y) (where YY is an m×mm\times m-matrix) is known be be efficiently computable. On the other hand, for the permanent form, permm(Y)perm_{m}(Y), no such efficient computation is known. The analysis of this divergence has been a central problem of the theory.

In a remarkable construction, Valiant[Val79] showed that for a form f(Y)Symd(Y)f(Y)\in Sym^{d}(Y), if there is a program111In this construction the program corresponds to an arithmetic formula. of LL lines then there is an O(L)×O(L)O(L)\times O(L)-matrix with entries as linear forms Aij(Y)A_{ij}(Y^{\prime}) in the variables Y=Y{z}Y^{\prime}=Y\cup\{z\} (where zz is merely a degree homogenizing variable) such that detL(A(Y))=z(Ld)f(Y)det_{L}(A(Y^{\prime}))=z^{(L-d)}f(Y).

We may thus define the determinantal complexity of a function fSymd(Y)f\in Sym^{d}(Y) as the smallest integer nn such that ff can be written as the determinant of an n×nn\times n matrix whose entries are linear forms in the variables Y=Y{z}Y^{\prime}=Y\cup\{z\}. Valiant’s construction shows that the determinant is universal, and that every form ff has a finite determinantal complexity.

Let us now connect determinantal complexity with the projective orbit closure of the form detn(X)Symn(X)det_{n}(X)\in Sym^{n}(X). Suppose that the determinantal complexity of fSymd(Y)f\in Sym^{d}(Y) is nn (naturally, with n>|Y|n>|Y|). Then (i) define X={xij,1i,jn}X=\{x_{ij},1\leq i,j\leq n\} as a set of new n2n^{2} variables, (ii) identify yijy_{ij} with xijx_{ij} for 1i,jm1\leq i,j\leq m and zz with xnnx_{nn}, and finally (iii) define the form fd,n(X):=xnnndf(Y)Symn(X)f_{d,n}(X):=x_{nn}^{n-d}f(Y)\in Sym^{n}(X). The condition that fd,n(X)=znddet(A(Y))f_{d,n}(X)=z^{n-d}det(A(Y^{\prime})) tell us that there is a (possibly singular) matrix AEnd(X)A\in End(\mathbb{C}\cdot X) such that det(AX)=fd,n(X)det(A\cdot X)=f_{d,n}(X). Whence this homogenized version of ff is in the GL(X)GL(X) projective orbit closure of detndet_{n}. Thus, computability of forms in various models of computation is intimately connected with the algebraic closures of orbits of universal functions associated with these models. This theme was first proposed as the Geometric Complexity Theoretic (GCT) approach in the study of computational complexity[MS01].

A central question in computational complexity has been to determine the determinantal complexity of permmperm_{m}, and thus to determine those nn for which permm,nO(detn)¯perm_{m,n}\in\overline{O(det_{n})}.

It is well known that invariant properties of these functions play an important part in the design of algorithms. For example, the invariance of the determinant to elementary row operations, has paved the way for its efficient evaluation. The group of symmetries of the permanent too is large, and also well known. In fact, both permm(Y)perm_{m}(Y) and detm(Y)det_{m}(Y) as forms in Symm(Y)Sym^{m}(Y) are determined by their stabilizers within GL(Y)GL(Y) - and yet this divergence in their computational complexity.

One approach to understanding this divergence is by a careful examination of their symmetries. More concretely, it is to ask if the question of [permm,n]O([detn])¯[perm_{m,n}]\not\in\overline{O([det_{n}])} can be cast in terms of the group theory, of linkages, or their absence, between the stabilizers of permm,nperm_{m,n} and detndet_{n}. This broad strategy of finding “obstructions” was also proposed as a method in the GCT approach through a sequence of papers, which we review briefly later.

A.1 Obstructions for membership in orbit closures

Let us now review a key approach of Geometric Complexity Theory, that of the hunt for obstructions for [x][x] to be in the orbit closure of [y][y]. This was formulated in [MS01], in the context of the permanent and determinant forms, both of which have large and distinctive stabilizers. We outline this briefly. Recall that XX denotes linear dual of the complex n2n^{2}-dimensional vector space of n×nn\times n-matrices over {\mathbb{C}}.

Let Δ[detn]\Delta[det_{n}] be the projective orbit closure of [detn][det_{n}] in (Symn(X)){\mathbb{P}}(Sym^{n}(X)). If [permm,n][perm_{m,n}] is in Δ[detn]\Delta[det_{n}] then there would be a GL(X)GL(X) equivariant embedding of the projective orbit closure of [permm,n][perm_{m,n}], , Zm,nZ_{m,n}, in Δ[detn]\Delta[det_{n}]. This would give rise to a surjective GL(X)GL(X)-morphism from the coordinate ring [Δ[detn]]{\mathbb{C}}[\Delta[det_{n}]] of Δ[detn]\Delta[det_{n}], to the coordinate ring [Zm,n]{\mathbb{C}}[Z_{m,n}] of Zm,nZ_{m,n}.

The distinctive nature of the stabilizers KK of [detn][det_{n}] and HH of [permm,n][perm_{m,n}] leads to a complete classification of the GL(X)GL(X)-modules which may appear in the above coordinate rings. Whence, the surjection implies that every irreducible GL(X)GL(X)-representation which occurs in [Zm,n]{\mathbb{C}}[Z_{m,n}] must occur in [Δ[detn]]{\mathbb{C}}[\Delta[det_{n}]]. An irreducible representation λ\lambda of GL(X)GL(X) which occurs in [Zm,n]{\mathbb{C}}[Z_{m,n}] and does not occur in [Δ[detn]]{\mathbb{C}}[\Delta[det_{n}]] is called an occurrence obstruction. Mulmuley and Sohoni conjectured that for every non negative integer cc and for infinitely many mm, there exist irreducible representations λ\lambda of GL(X)GL(X) (with n=mcn=m^{c}) which occur in [Zm,mc]{\mathbb{C}}[Z_{m,m^{c}}] but not in [Δ[detmc]]{\mathbb{C}}[\Delta[det_{m^{c}}]]. This conjecture was shown to be false for irreducible representations of GL(X)GL(X) of shape n×dn\times d in [IP17]. In [BIP19] the authors showed that the conjecture is false for all λ\lambda, when nm25n\geq m^{25}. Note that these results do not rule out the GCT approach. If one were to show that the multiplicity of some irreducible GL(X)GL(X)-module λ\lambda is more in [Zm,mc]{\mathbb{C}}[Z_{m,m^{c}}] than in [Δ[detmc]]{\mathbb{C}}[\Delta[det_{m^{c}}]] then an obstruction has been found. If one were to show this for infinitely many mm and every cc that would separate the complexity class VPVP from VNPVNP, the flagship problem of GCT.

Mignon and Ressayre [MR04] showed that the determinantal complexity of permmperm_{m} is Ω(m2)\Omega(m^{2}). They consider the zero locus of the determinant hypersurface detn=0det_{n}=0 in n2{\mathbb{C}}^{n^{2}} and the zero locus of the permanent hypersurface permm=0perm_{m}=0 in m2{\mathbb{C}}^{m^{2}}. They show that the Hessian of the determinant at any point on detn=0det_{n}=0 has rank at most 2n+12n+1 and this rank is m2m^{2} at a generic point on permm=0perm_{m}=0. Their proof follows from this beautiful calculation. Although their proof does not explicitly use obstructions, Growchow [Gro15] showed, using a construction of [MLR13], that their proof can also be seen through the lens of representation theoretic obstructions.

Appendix B Stabilizer conditions in special cases

Let’s analyse the stabilization conditions of Theorem 3.13 for an element 𝔥𝒦0\mathfrak{h}\in{\cal K}_{0} in a special case. We have:

𝔰g=𝔥fb\mathfrak{s}\cdot g=\mathfrak{h}\cdot f_{b}

We may split 𝔰\mathfrak{s} further as 𝔰=𝔰1+𝔰0+𝔰1\mathfrak{s}=\mathfrak{s}_{-1}+\mathfrak{s}_{0}+\mathfrak{s}_{1}, and 𝔥\mathfrak{h} as 𝔥=𝔥1+𝔥0+𝔥1\mathfrak{h}=\mathfrak{h}_{-1}+\mathfrak{h}_{0}+\mathfrak{h}_{1} to give us:

𝔰1g+𝔰0g+𝔰1g=𝔥1fb+𝔥0fb+𝔥1fb\mathfrak{s}_{-1}\cdot g+\mathfrak{s}_{0}\cdot g+\mathfrak{s}_{1}\cdot g=\mathfrak{h}_{-1}\cdot f_{b}+\mathfrak{h}_{0}\cdot f_{b}+\mathfrak{h}_{1}f_{b}

Now note that 𝔰0,𝔥0\mathfrak{s}_{0},\mathfrak{h}_{0} are elements of Hom(Y,Y)Hom(Z,Z)Hom(Y,Y)\oplus Hom(Z,Z). Typically, the dimension of Hom(Y,Y)Hom(Y,Y) may be much smaller that the dimension of 𝒦0{\cal K}_{0}. Thus, there is a subspace 𝒦𝒦0{\cal K}^{\prime}\subseteq{\cal K}_{0} of dimension k|Y|2k-|Y|^{2} where we have that for 𝔥𝒦\mathfrak{h}\in{\cal K}^{\prime}, we have 𝔥=𝔥1+𝔥Z+𝔥1\mathfrak{h}=\mathfrak{h}_{-1}+\mathfrak{h}_{Z}+\mathfrak{h}_{1}, where 𝔥ZHom(Z,Z)\mathfrak{h}_{Z}\in Hom(Z,Z), and that the stabilizer condition at degree bb and b1b-1 is satisfied as follows:

𝔰0g+𝔰1g=𝔥1fb+𝔥Zfb\mathfrak{s}_{0}\cdot g+\mathfrak{s}_{1}\cdot g=\mathfrak{h}_{-1}\cdot f_{b}+\mathfrak{h}_{Z}f_{b}

Noting that gVag\in V_{a} while fbVbf_{b}\in V_{b} gives us the following possibilities, (i) b=a+1b=a+1, (ii) b=a+2b=a+2, and (iii) b>a+2b>a+2. Let us take each of them one by one.

  1. 1.

    Case b=a+1b=a+1. We have 𝔥1fb=𝔰0g\mathfrak{h}_{-1}f_{b}=\mathfrak{s}_{0}\cdot g and 𝔰1g=𝔥Zfb\mathfrak{s}_{1}g=\mathfrak{h}_{Z}f_{b}.

  2. 2.

    Case b=a+2b=a+2. We have 𝔰1g=𝔥1fb\mathfrak{s}_{1}g=\mathfrak{h}_{-1}f_{b} and 𝔥Zfb=0\mathfrak{h}_{Z}f_{b}=0.

  3. 3.

    Case b>a+2b>a+2. We have 𝔰1g=0,𝔥1fb=0\mathfrak{s}_{1}g=0,\mathfrak{h}_{-1}f_{b}=0 and 𝔥Zfb=0\mathfrak{h}_{Z}f_{b}=0.

Each condition gives us certain stringent conditions on the relationship between gg and fbf_{b}. As an example, let us take the case b=a+1b=a+1 and an 𝔥𝒦\mathfrak{h}\in{\cal K}^{\prime}. Let us assume that |Z|=r|Z|=r and |Y|=s|Y|=s. We may find a basis for Z\mathbb{C}\cdot Z such that 𝔥Z\mathfrak{h}_{Z} is in Jordan canonical form. Assume for the moment that 𝔥Z\mathfrak{h}_{Z} is diagonal with eigenvalues μ1,,μr\mu_{1},\ldots,\mu_{r} where the first rr^{\prime} are non-zero. We may express fbf_{b} in this new basis as fb=αzαfbαf_{b}=\sum_{\alpha}z_{\alpha}f_{b}^{\alpha}, where zαz_{\alpha} are distinct monomials of degree a+1a+1 and fbαSymda1(Y)f^{\alpha}_{b}\in Sym^{d-a-1}(Y). By the same token, we may express gg as βzβgβ\sum_{\beta}z_{\beta}g^{\beta}, where zβSyma(Z)z^{\beta}\in Sym^{a}(Z) and gβSymda(Y)g^{\beta}\in Sym^{d-a}(Y). The condition 𝔰1g=𝔯Zb\mathfrak{s}_{1}g=\mathfrak{r}_{Z}b gives :

βi=1sLi(z)g/yi=αμαzαfbα\sum_{\beta}\sum_{i=1}^{s}L_{i}(z)\cdot\partial g/\partial y_{i}=\sum_{\alpha}\mu^{\alpha}z_{\alpha}f^{\alpha}_{b}

This tells us that those α{\alpha} such that μα0\mu^{\alpha}\neq 0, we have that fbαf_{b}^{\alpha} are \mathbb{C}-linear combinations of g/yi\partial g/\partial y_{i}, in other words, the “minors” of gg. The first condition, viz., 𝔮fb=0\mathfrak{q}f_{b}=0 gives us that there are linear relations between the fbαf^{\alpha}_{b}, i.e., the minors of gg, with coefficients in Sym1(Y)Sym^{1}(Y). These are akin to the Laplace’s conditions on the expansion of the determinant.