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

Smooth Euler-symmetric varieties generated by a single polynomial

Cong Ding School of Mathematical Sciences
Shenzhen University
Guangdong
China
[email protected]
 and  Zhijun Luo Center for complex geometry
Institute for Basic Science
55 Expo-ro
Yuseong-gu
Daejeon
34126
Republic of Korea.
[email protected]
[email protected]
Abstract.

We classify smooth Euler-symmetric varieties corresponding to the symbol system generated by a single reduced polynomial.

Keywords. Euler-symmetric variety, equivariant compactification, fundamental form, symbol system.

2020 MSC 14M27.

1. Introduction

The notion of Euler-symmetric varieties was first introduced by Hwang and Fu in [FH20]. They are nondegenerate projective varieties admitting many \mathbb{C}^{*}-actions of Euler type. More precisely we have

Definition 1.1.

Let ZVZ\subset\mathbb{P}V be a nondegenerate projective variety. For a nonsingular point xZx\in Z, a \mathbb{C}^{*}-action on ZZ coming from a multiplicative subgroup of GL(V)\operatorname{GL}(V) is said to be of Euler type at xx if

  1. (1)

    xx is an isolated fixed point of the restricted \mathbb{C}^{*}-action on ZZ;

  2. (2)

    the induced \mathbb{C}^{*}-action on the tangent space Tx(Z)T_{x}(Z) is by scalar multiplication.

We say that ZVZ\subset\mathbb{P}V is Euler-symmetric if for a general point xZx\in Z, there exists a \mathbb{C}^{*}-action on ZZ of Euler type at xx.

There are many examples of Euler-symmetric varieties. For instance, Hermitian symmetric spaces under the equivariant projective embedding with respect to their automorphism groups, blow-up of n\mathbb{P}^{n} along a subvariety inside a hyperplane (where the projective embeddings have not been classified) are Euler-symmetric. More examples can be found in [FH20]. Characterizing or classifying Euler-symmetric varieties under specific conditions presents a fascinating challenge. In [FH20], the authors propose a conjecture that a Fano manifold with Picard number 1, realizable as an equivariant compactification of a vector group, can also be realized as an Euler-symmetric projective variety under a suitable projective embedding. This conjecture has been solved in [Sha23] under the assumption that the Fano manifold is toric. In [Luo23], the second author demonstrates that any Euler-symmetric complete intersection is necessarily a complete intersection of hyperquadrics.

From [FH20], Euler-symmetric varieties are determined by the symbol systems. We first recall

Definition 1.2.

Let WW be a vector space. For wWw\in W, define

ιw:Symk+1WSymkW,φφ(w,,,).\iota_{w}:\operatorname{Sym}^{k+1}W^{*}\to\operatorname{Sym}^{k}W^{*},\,\varphi\mapsto\varphi(w,\cdot,\ldots,\cdot).

For a subspace FkSymkWF^{k}\subset\operatorname{Sym}^{k}W^{*} of symmetric kk-linear forms on WW, define its prolongation prolong(Fk)Symk+1W\textbf{prolong}(F^{k})\subset\operatorname{Sym}^{k+1}W^{*} by the following

prolong(Fk):=wWιw1(Fk).\textbf{prolong}(F^{k}):=\bigcap_{w\in W}\iota_{w}^{-1}(F^{k}).

Then a symbol system is defined as follows.

Definition 1.3.

Let WW be a vector space. Fix a natural number rr. A subspace

F=k0FkSym(W):=k0SymkW\textbf{F}=\oplus_{k\geq 0}F^{k}\subset\operatorname{Sym}(W^{*}):=\oplus_{k\geq 0}\operatorname{Sym}^{k}W^{*}

with

F0==Sym0W,F1=W,Fr0,andFr+i=0i1,F^{0}=\mathbb{C}=\operatorname{Sym}^{0}W^{*},\;F^{1}=W^{*},\;F^{r}\neq 0,\;\text{and}\;F^{r+i}=0\;\forall\;i\geq 1,

is called a symbol system of rank rr, if Fk+1prolong(Fk)F^{k+1}\subset\textbf{prolong}(F^{k}) for each 1kr1\leq k\leq r.

One typical example of a symbol system comes from the system of fundamental forms of a non-degenerate projective variety. Moreover we know

Theorem 1.4 (Classical result due to E. Cartan).

Let ZPVZ\subset PV be a nondegenerate subvariety and let xZx\in Z be a general point. Then the system of fundamental forms Fx=k0Fxk\textbf{F}_{x}=\oplus_{k\geq 0}F_{x}^{k} is a symbol system of rank rr for some natural number r1r\geq 1.

Definition 1.5.

Given a symbol system F of rank rr, define a rational map

ΦF:(W)(W(F2)(Fr)),\Phi_{\textbf{F}}:\mathbb{P}(\mathbb{C}\oplus W)\dashrightarrow\mathbb{P}(\mathbb{C}\oplus W\oplus(F^{2})^{*}\oplus\cdots\oplus(F^{r})^{*}),

by

[t:w][tr:tr1w:tr2ιw2::tιwr1:ιwr].[t:w]\mapsto[t^{r}:t^{r-1}w:t^{r-2}\iota^{2}_{w}:\cdots:t\,\iota^{r-1}_{w}:\iota^{r}_{w}].

Write VF:=W(F2)(Fr)V_{\textbf{F}}:=\mathbb{C}\oplus W\oplus(F^{2})^{*}\oplus\cdots\oplus(F^{r})^{*}. We will denote the closure of the image of the rational map ΦF\Phi_{\textbf{F}} by M(F)VFM(\textbf{F})\subset\mathbb{P}V_{\textbf{F}}. We say the projective variety M(F)M(\textbf{F}) associated to the symbol system F has rank rr, denoted by rank(M(F))\operatorname{rank}(M(\textbf{F})).

The following theorem gives the relation between Euler-symmetric varieties and symbol systems.

Theorem 1.6.

[FH20, Theorem 3.3] Let o=[1:0::0]M(F)o=[1:0:\cdots:0]\in M(\textbf{F}) be the point ΦF([t=1:w=0])\Phi_{\textbf{F}}([t=1:w=0]). Then:

  1. (1)

    The natural action of the vector group WW on (W)\mathbb{P}(\mathbb{C}\oplus W) can be extended to an action of WW on VF\mathbb{P}V_{\textbf{F}} preserving M(F)M(\textbf{F}) such that the orbit of oo is an open subset biregular to WW.

  2. (2)

    The \mathbb{C}^{*}-action on WW with weight 11 induces a \mathbb{C}^{*}-action on M(F)M(\textbf{F}) of Euler type at oo, making M(F)M(\textbf{F}) Euler-symmetric.

  3. (3)

    The system of fundamental forms of M(F)VFM(\textbf{F})\subset\mathbb{P}V_{\textbf{F}} at oo is isomorphic to the symbol system F.

Conversely, any Euler-symmetric projective variety is of the form M(F)M(\textbf{F}) for some symbol system F on a vector space WW.

Remark 1.7.

The action of WW on V𝐅\mathbb{P}V_{\mathbf{F}} coming from the natural action of the vector group WW on (W)\mathbb{P}(\mathbb{C}\oplus W) can be explicitly written as follows. For vWv\in W and x=[t:w:f2::fr]x=[t:w:f^{2}:\cdots:f^{r}],

gvx=[t:w+tv:gvxf2::gvxfr],g_{v}\cdot x=[t:w+tv:g_{v}^{x}\cdot f^{2}:\cdots:g_{v}^{x}\cdot f^{r}],

where for each 2kr2\leq k\leq r,

gvxfk==2kCkfιvk+kιwιvk1+tιvk,g_{v}^{x}\cdot f^{k}=\sum_{\ell=2}^{k}C_{k}^{\ell}f^{\ell}\circ\iota_{v}^{k-\ell}+k\iota_{w}\circ\iota_{v}^{k-1}+t\iota_{v}^{k},

denoting the binomial coefficients by CkC^{\ell}_{k}.

The action of \mathbb{C}^{*} on V𝐅\mathbb{P}V_{\mathbf{F}} that induced by the \mathbb{C}^{*}-action on WW with weight 11 can be explicitly written as follows. For λ\lambda\in\mathbb{C}^{*} and x=[t:w:f2::fr]x=[t:w:f^{2}:\cdots:f^{r}],

λx=[t:λw:λ2f2::λrfr].\lambda\cdot x=[t:\lambda w:\lambda^{2}f^{2}:\cdots:\lambda^{r}f^{r}].
Definition 1.8.

Let UU be the open orbit biregular to WW, and the boundary divisor D=M(F)\UD=M(\textbf{F})\backslash U.

In this article we are going to characterize certain special Euler-symmetric varieties associated to symbol systems generated by a single reduced homogeneous polynomial.

Definition 1.9.

Given a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a homogeneous polynomial of degree rr. The associated symbol system FP:=i=0rFPi\textbf{F}_{P}:=\oplus_{i=0}^{r}F^{i}_{P} is defined as follows:

FP2=dr2P,,FPk=drkP,,FPr=P.F^{2}_{P}=\langle d^{r-2}P\rangle,\cdots,F^{k}_{P}=\langle d^{r-k}P\rangle,\cdots,F^{r}_{P}=\langle P\rangle.

In this case, we denote VP:=VFPV_{P}:=V_{\textbf{F}_{P}}, and the Euler-symmetric variety associated to this symbol system is denoted by MPVPM_{P}\subset\mathbb{P}V_{P}, correspondingly the rational map is denoted by ΦP\Phi_{P}.

Our main result is the following.

Theorem 1.10.

Given a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a reduced homogeneous polynomial of degree rr. If the Euler-symmetric variety MPM_{P} is smooth, then MPM_{P} is a homogeneous projective variety.

Moreover, MP=X1××XkM_{P}=X_{1}\times\cdots\times X_{k} where kk is the number of irreducible components of PP, and each XiX_{i} is one of the following irreducible Hermitian symmetric spaces, and the embedding MPVPM_{P}\subset\mathbb{P}V_{P} is given by the line bundle 𝒪(1,,1)\mathcal{O}(1,\cdots,1).

  1. (1)

    the Grassmannian variety Gr(n,2n)Gr(n,2n);

  2. (2)

    the Spinor variety 𝕊n\mathbb{S}_{n} with nn being even;

  3. (3)

    the Lagrangian Grassmannian LG(n,2n)\operatorname{LG}(n,2n);

  4. (4)

    the hyperquadric n\mathbb{Q}^{n};

  5. (5)

    the 2727-dimensional E7/P7E_{7}/P_{7}.

Conversely, for any X=X1××XkNX=X_{1}\times\cdots\times X_{k}\subset\mathbb{P}^{N} such that XiX_{i} is one of the above irreducible Hermitian symmetric spaces with the embedding in N\mathbb{P}^{N} given by the line bundle 𝒪(1,,1)\mathcal{O}(1,\cdots,1), there exists a reduced homogeneous polynomial PP such that (XN)(MPVP)(X\subset\mathbb{P}^{N})\cong(M_{P}\subset\mathbb{P}V_{P}), up to projective isomorphism.

The key point of the proof lies in demonstrating that the system of fundamental forms at the terminal point z=[0,,0,1]z=[0,\cdots,0,1] constitutes a symbol system. This presents a challenge because Cartan’s Theorem, which holds on general points, does not readily provide this information. Algebraically, Proposition 2.10 allows us to easily obtain only the information of the (r1)(r-1)-th fundamental form. This form is precisely captured by the derivatives of PP_{*}, the so-called multiplicative Legendre transform of PP (Definition 2.6). However, it is not easy to check that the system of fundamental forms of MPM_{P} at zz is exactly coincident with the symbol system generated by PP_{*}. Instead of this tedious algebraic verification, we use a theorem recently proved by Lawrence Ein and Wenbo Niu (see [EN23, Theorem 3.3]). This theorem establishes that the rank equality between the system of fundamental forms and the symbol system generated by PP (see Proposition 2.5) will imply that the system of fundamental forms at zz is a symbol system. Consequently, the existence of another vector group action on MPM_{P} is guaranteed, with its orbit through zz being biregular to the vector group.

The organization of the rest of this paper is as follows: In §2 we discuss the homaloidal polynomial and its relation with the smoothness of the Euler symmetric variety MPM_{P}, which is sufficient to obtain the rank equality. In §3 we discuss some properties on the fixed point components of MPM_{P} in the Białynicki-Birula decomposition and the boundary divisor. In §4 we finish the proof of our main result (Theorem 1.10).

2. EKP-Homaloidal polynomials and smoothness of Euler-symmetric varieties

We first recall

Definition 2.1.

Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a homogeneous polynomial. We say PP is a homaloidal polynomial, if the induced derivative map

P:WW,w(dP)w,P^{\prime}:\mathbb{P}W\to\mathbb{P}W^{*},w\mapsto(dP)_{w},

is birational map.

We list some cubic homaloidal polynomials here.

Example 2.2.
  1. (1)

    W=W=\mathbb{C}, P=x3P=x^{3};

  2. (2)

    W=mW=\mathbb{C}^{m}\oplus\mathbb{C}, P=Q(v)xP=Q(v)x, where QQ is any non-degenerate quadratic form on m\mathbb{C}^{m};

  3. (3)

    W=3W=\mathbb{C}^{3}, P=x0x1x2P=x_{0}x_{1}x_{2};

  4. (4)

    WW is the space of 33 by 33 symmetric matrices, PP is the determinant function;

  5. (5)

    W=Mat3()W=\operatorname{Mat}_{3}(\mathbb{C}), PP is the determinant function;

  6. (6)

    W=Λ2(6)W=\Lambda^{2}(\mathbb{C}^{6}), PP is the Pffafian polynomial;

  7. (7)

    W=Mat3()×Mat3()×Mat3()W=\operatorname{Mat}_{3}(\mathbb{C})\times\operatorname{Mat}_{3}(\mathbb{C})\times\operatorname{Mat}_{3}(\mathbb{C}), the polynomial PP is the Cartan cubic P(A,B,C)=|A|+|B|+|C|Tr(ABC)P(A,B,C)=|A|+|B|+|C|-\operatorname{Tr}(ABC).

Moreover, the last four examples correspond to irreducible Hermitian symmetric spaces listed in Theorem 1.10 with rank=3\operatorname{rank}=3.

Proposition 2.3.

Let PP be a homogeneous polynomial, then

dim(FPj)=dim(FPrj),j1.\operatorname{dim}(F^{j}_{P})=\operatorname{dim}(F^{r-j}_{P}),\forall j\neq 1.

Moreover, if PP is homaloidal, then

dim(FPj)=dim(FPrj),j.\operatorname{dim}(F^{j}_{P})=\operatorname{dim}(F^{r-j}_{P}),\forall j.
Proof.

Let SymdW\operatorname{Sym}^{d}W be the vector space of degree dd homogeneous differential operator with constant coefficients. Consider the follow jj-th partial derivative map of PP:

Pj,rj:SymjWSymrjW;DD(P).P_{j,r-j}:\operatorname{Sym}^{j}W\to\operatorname{Sym}^{r-j}W^{*};\;D\mapsto D(P).

By the definition of the symbol system FP\textbf{F}_{P}, we have that Im(Pj,rj)=FPrj\operatorname{Im}(P_{j,r-j})=F_{P}^{r-j} for jr1j\neq r-1. We claim that

rank(Pj,rj)=rank(Prj,j).\operatorname{rank}(P_{j,r-j})=\operatorname{rank}(P_{r-j,j}).

This was mentioned in [GL19] without explicit proof. To prove the claim, we write

P=I=(a1,,am),ai0a1++am=rAI1a1!am!x1a1xmam.P=\sum_{\begin{subarray}{c}I=(a_{1},\cdots,a_{m}),a_{i}\geq 0\\ a_{1}+\cdots+a_{m}=r\end{subarray}}A_{I}\dfrac{1}{a_{1}!\cdots a_{m}!}x^{a_{1}}_{1}\cdots x^{a_{m}}_{m}.

For DjSymjWD^{j}\in\operatorname{Sym}^{j}W, it can be written as

Dj=J=(b1,,bm),bi0b1++bm=jDJjjx1b1xmbm.D^{j}=\sum_{\begin{subarray}{c}J=(b_{1},\cdots,b_{m}),b_{i}\geq 0\\ b_{1}+\cdots+b_{m}=j\end{subarray}}D^{j}_{J}\dfrac{\partial^{j}}{\partial x^{b_{1}}_{1}\cdots\partial x^{b_{m}}_{m}}.

Then we have

Dj(P)=I,J,aibiAIDJj1(a1b1)!(ambm)!x1a1b1xmambm.D^{j}(P)=\sum_{I,J,a_{i}\geq b_{i}}A_{I}D^{j}_{J}\dfrac{1}{(a_{1}-b_{1})!\cdots(a_{m}-b_{m})!}x^{a_{1}-b_{1}}_{1}\cdots x^{a_{m}-b_{m}}_{m}.

Define Φ:Im(Pj,rj)Im(Prj,j)\Phi:\operatorname{Im}(P_{j,r-j})\rightarrow\operatorname{Im}(P_{r-j,j}) as

Φ(Dj(P))=I,J,aibiAIDJj1(b1)!(bm)!x1b1xmbm.\Phi(D^{j}(P))=\sum_{I,J,a_{i}\geq b_{i}}A_{I}D^{j}_{J}\dfrac{1}{(b_{1})!\cdots(b_{m})!}x^{b_{1}}_{1}\cdots x^{b_{m}}_{m}.

It is easy to see that Φ\Phi is a bijection and hence the claim holds. Therefore, dim(FPj)=dim(FPrj)\operatorname{dim}(F^{j}_{P})=\operatorname{dim}(F^{r-j}_{P}) for any j1j\neq 1.

If PP is homaloidal, then dim(FPr1)=m=dim(W)\operatorname{dim}(F^{r-1}_{P})=m=\operatorname{dim}(W^{*}). ∎

We fix some conventions here. Consider the WW-action on M𝐅M_{\mathbf{F}} as stated in Theorem 1.6. For a vector vWv\in W, denoting the corresponding group element by gvg_{v}, the limiting point

limt0gt1voM𝐅\lim_{t\to 0}g_{t^{-1}v}\cdot o\in M_{\mathbf{F}}

is called the terminal point of the rational curve {gtvot}¯MP\overline{\{g_{tv}\cdot o\mid t\in\mathbb{C}\}}\subset M_{P}, and oo is called the original point. If vv is a general vector, the limiting points are called the terminal points of the WW-action.

Homaloidal polynomials are closely related to the smoothness of MPM_{P}. More precisely we have

Proposition 2.4.

If the Euler-symmetric variety MPM_{P} is smooth, then PP is a homaloidal polynomial.

Proof.

Consider the terminal point z=[0::0:1]MPVPz=[0:\cdots:0:1]\in M_{P}\subset\mathbb{P}V_{P}. Firstly, for any point wWw\in W such that P(w)0P(w)\neq 0, we have that the projective tangent cone at zz contains the direction [0::0:dP(w):0][0:\cdots:0:dP(w):0]. We claim that the the map wdP(w)w\mapsto dP(w) is not degenerate, namely the image is not contained in a hyperplane. If not, by choosing a suitable coordinate, we can assume that P/wm0\partial P/\partial w_{m}\equiv 0. Hence MPM_{P} is degenerate. It contradicts with the non-degeneracy of Euler-symmetric projective varieties. Therefore, the projective tangent cone must contain the projective subspace L={[0::0::]}L=\{[0:\cdots:0:*:*]\} of dimension mm.

Consider the tangential projection of MPM_{P} from zz:

ψ:MP\displaystyle\psi:M_{P} L\displaystyle\dashrightarrow L
[tr:tr1w::tdP(w):P(w)]\displaystyle[t^{r}:t^{r-1}w:\cdots:tdP(w):P(w)] [tdP(w):P(w)].\displaystyle\mapsto[tdP(w):P(w)].

Since the Euler-symmetric variety MPM_{P} is smooth, the projective subspace LL coincides with projective tangent space of MPM_{P} at point zz. Then ψ\psi must be a local isomorphism in the complex topology. That is, the map ψ\psi must be a local isomorphism at zz, in particular it must be locally injective.

On the other hand, suppose that we can find two non-colinear vectors w1w_{1} and w2w_{2} with P(w1),P(w2)0P(w_{1}),P(w_{2})\neq 0 such that the vectors dP(w1)dP(w_{1}) and dP(w2)dP(w_{2}) are colinear. After multiplying them by a suitable constant, we may suppose that P(w1)1dP(w1)=P(w2)1dP(w2)P(w_{1})^{-1}dP(w_{1})=P(w_{2})^{-1}dP(w_{2}). Then for tt small enough, [tr:tr1w1::tdP(w1):P(w1)][t^{r}:t^{r-1}w_{1}:\cdots:tdP(w_{1}):P(w_{1})] and [tr:tr1w2::tdP(w2):P(w2)][t^{r}:t^{r-1}w_{2}:\cdots:tdP(w_{2}):P(w_{2})] are two distinct points in MPM_{P} close to zz. But [tdP(w1):P(w1)][tdP(w_{1}):P(w_{1})] and [tdP(w2):P(w2)][tdP(w_{2}):P(w_{2})] coincide, a contradiction with the local injectivity of ψ\psi. We conclude that the rational map [w]W[dP(w)]W[w]\in\mathbb{P}W\mapsto[dP(w)]\in\mathbb{P}W^{*} must be injective on the open subset P(w)0P(w)\neq 0. Hence, the polynomial PP is homaloidal. ∎

Since MPM_{P} is smooth, we consider the system of fundamental forms at zz, which is denoted by Fz=k=0rzFzk\textbf{F}_{z}=\oplus_{k=0}^{r_{z}}F^{k}_{z}. Note that we can not use Theorem 1.4 to deduce that Fz\textbf{F}_{z} is a symbol system directly since zz might not be a general point. However, the dimensions match as follows.

Proposition 2.5.

rz=rr_{z}=r and dim(Fzk)=dim(FPrk)=dim(FPk)\operatorname{dim}(F_{z}^{k})=\operatorname{dim}(F_{P}^{r-k})=\operatorname{dim}(F_{P}^{k}), for any 0kr0\leq k\leq r.

Proof.

Let w\mathscr{L}_{w} be the image of the line Lw={[a:bw][a:b]1,P(w)0}L_{w}=\{[a:bw]\mid[a:b]\in\mathbb{P}^{1},P(w)\neq 0\} under the rational map ΦP\Phi_{P}. The tangent direction of w\mathscr{L}_{w} at zz is

[0::dP(w)P(w):0],[0:\cdots:\frac{dP(w)}{P(w)}:0],

and the coordinate of the neighborhood of zz is

[trP(w):tr1wP(w):tr2ιw2P(w)::tdP(w)P(w):1].[\frac{t^{r}}{P(w)}:\frac{t^{r-1}w}{P(w)}:\frac{t^{r-2}\iota^{2}_{w}}{P(w)}:\cdots:\frac{tdP(w)}{P(w)}:1].

Therefore, the tangent space TzMPT_{z}M_{P} can be identified with WW^{*}. Hence by [FH20, Lemma 2.5] and the smoothness of MPM_{P}, we can take the coordinate

(zmr(r),,z1(r),,zm2(2),,z1(2),zm,,z1),(z_{m_{r}}^{(r)},\cdots,z_{1}^{(r)},\cdots,z_{m_{2}}^{(2)},\cdots,z_{1}^{(2)},z_{m},\cdots,z_{1}),

such that zj=tjP(w)P(w)z_{j}=\frac{t\partial_{j}P(w)}{P(w)}, and analytic functions hjk(zm,,z1)h^{k}_{j}(z_{m},\cdots,z_{1}) satisfy

zj(k)=hjk(zm,,z1),2kr,1jmi.z_{j}^{(k)}=h^{k}_{j}(z_{m},\cdots,z_{1}),2\leq k\leq r,1\leq j\leq m_{i}.

By the Taylor’s expansion of hjkh^{k}_{j} and comparing the degree w.r.t tt, we know that hjkh^{k}_{j} are homogeneous polynomials of degree kk. Hence rz=rr_{z}=r and the system of fundamental forms of MPM_{P} at zz is Fz=k=0rFzk\textbf{F}_{z}=\oplus_{k=0}^{r}F^{k}_{z} with Fzk=hjk1jmkSymkWF_{z}^{k}=\langle h^{k}_{j}\mid 1\leq j\leq m_{k}\rangle\subset\operatorname{Sym}^{k}W for k2k\geq 2. Together with Proposition 2.3 and Proposition 2.4 we know mk=dim(FPrk)=dim(FPk)m_{k}=\operatorname{dim}(F_{P}^{r-k})=\operatorname{dim}(F_{P}^{k}), for any 0kr0\leq k\leq r. ∎

Actually from the smoothness of MPM_{P}, we have more restrictions on the polynomial PP. Recall

Definition 2.6.

Let PP be a homogeneous function and det(Hess(ln(P)))\operatorname{det}(\operatorname{Hess}(\operatorname{ln}(P))) is not identically zero. In this case we can define a function PP_{*} by

P(dP(w)P(w))=1P(w).P_{*}(\frac{dP(w)}{P(w)})=\frac{1}{P(w)}.

If PP is homogeneous of degree dd then so is PP_{*}. We will call PP_{*} the multiplicative Legendre transform of PP.

Proposition 2.7.

[EKP02, Proposition 3.6] Fix a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a homogeneous polynomial of degree rr, and det(Hess(ln(P)))\operatorname{det}(\operatorname{Hess}(\operatorname{ln}(P))) is not identically zero. Then PP is a homaloidal polynomial if and only if its multiplicative Legendre transform PP_{*} is a rational function. Moreover, in this case,

dlnP=(dlnP)1.d\operatorname{ln}P_{*}=(d\operatorname{ln}P)^{-1}.
Definition 2.8.

We say a homogeneous polynomial PP is an EKP-homaloidal polynomial if its multiplicative Legendre transform PP_{*} is also a homogeneous polynomial.

From [EKP02, Lemma 3.5], we know P=PP_{**}=P. Hence PP is an EKP-homaloidal polynomial if and only if PP_{*} is an EKP-homaloidal polynomial. Following the proof of Proposition 2.5, we immediately have

Proposition 2.9.

If the Euler-symmetric variety MPM_{P} is smooth, then PP is an EKP-homaloidal polynomial.

Proof.

From the proof of Proposition 2.5, we know that 1P(w)=h1r(dP(w)P(w))\frac{1}{P(w)}=h^{r}_{1}(\frac{dP(w)}{P(w)}), namely P=h1rP_{*}=h^{r}_{1}. Hence PP is an EKP-homaloidal polynomial. ∎

Let π:BlzMPMP\pi:\textbf{Bl}_{z}M_{P}\to M_{P}, EE be the exceptional divisor. In fact, the inverse map ρ:=ΦP1:MP(W)\rho:=\Phi^{-1}_{P}:M_{P}\dashrightarrow\mathbb{P}(\mathbb{C}\oplus W) is the projection map. Then the composed rational map π¯:=ρπ:E(W)\overline{\pi}:=\rho\circ\pi:E\dashrightarrow\mathbb{P}(\mathbb{C}\oplus W) is defined by the (r1)(r-1)-th fundamental form of MPM_{P} at zz, namely, hjr1h^{r-1}_{j}, 1jm1\leq j\leq m. In particular, the image of the rational map π¯\overline{\pi} is contained in W\mathbb{P}W.

Proposition 2.10.

If the Euler-symmetric variety MPM_{P} is smooth, then Fzr1=FPr1F^{r-1}_{z}=F^{r-1}_{P_{*}}. In particular, π¯\overline{\pi} is birational map.

Proof.

By definition of hjr1h^{r-1}_{j}, we have that hm+1jr1(dP(w)P(w))=wjP(w)h^{r-1}_{m+1-j}(\frac{dP(w)}{P(w)})=\frac{w_{j}}{P(w)}, for wWw\in W. Let Π:WW,w(hmr1(w),,h1r1(w))\Pi:W^{*}\to W,w\mapsto(h^{r-1}_{m}(w),\cdots,h^{r-1}_{1}(w)). Identifying EE with W\mathbb{P}W^{*}, the rational map π¯\overline{\pi} and Π\Pi are same under the projectivization. Consider the morphisms

Φ1:\displaystyle\Phi_{1}: WW,wdP(w),\displaystyle\;W\to W^{*},w\mapsto dP(w),
Φ1:\displaystyle\Phi_{1}^{*}: WW,wdP(w).\displaystyle\;W^{*}\to W,w\mapsto dP_{*}(w).

By Proposition 2.7 and the definition of EKP-homaloidal polynomials, we know that

Φ1Φ1(w)=Pr2(w)w,Φ1Φ1(w)=Pr2(w)w.\Phi_{1}\circ\Phi_{1}^{*}(w)=P_{*}^{r-2}(w)w,\Phi_{1}^{*}\circ\Phi_{1}(w)=P^{r-2}(w)w.

On the other hand since hjr1h^{r-1}_{j} are homogeneous polynomials of degree r1r-1, we also have

ΠΦ1(w)=Pr2(w)w,\Pi\circ\Phi_{1}(w)=P^{r-2}(w)w,

therefore ΠΦ1=Φ1Φ1\Pi\circ\Phi_{1}=\Phi_{1}^{*}\circ\Phi_{1}. Composed on the right by the morphsim Φ1\Phi_{1}^{*}, we have

Φ1Φ1Φ1(w)\displaystyle\Phi_{1}^{*}\circ\Phi_{1}\circ\Phi_{1}^{*}(w) =P(r2)(r1)(w)Φ1(w),\displaystyle=P^{(r-2)(r-1)}_{*}(w)\Phi_{1}^{*}(w),
ΠΦ1Φ1(w)\displaystyle\Pi\circ\Phi_{1}\circ\Phi_{1}^{*}(w) =P(r2)(r1)(w)Π(w).\displaystyle=P^{(r-2)(r-1)}_{*}(w)\Pi(w).

Therefore, we know that Π=Φ1\Pi=\Phi_{1}^{*}. ∎

3. Białynicki-Birula decomposition and the boundary divisor

Lemma 3.1.

Let 𝔾mGL(VP)\mathbb{G}_{m}\subset GL(V_{P}) be the multiplicative subgroup corresponding to the \mathbb{C}^{*}-action on MPM_{P} in Remark 1.7. Let VFPiVP\mathbb{P}V_{F_{P}^{i}}\subset\mathbb{P}V_{P} be the (mi1)(m_{i}-1)-plane, Mi=MPVFPiM_{i}=M_{P}\cap\mathbb{P}V_{F^{i}_{P}}. Then the set of fixed points (MP)𝔾m=i=0rMi(M_{P})^{\mathbb{G}_{m}}=\coprod_{i=0}^{r}M_{i}, which is a disjoint union.

Proof.

By the definition of 𝔾m\mathbb{G}_{m}-action on the projective space VP\mathbb{P}V_{P}, the set of fixed points is the disjoint union i=0rVFPi\coprod_{i=0}^{r}\mathbb{P}V_{F_{P}^{i}}. Hence, (MP)𝔾m=i=0rMi(M_{P})^{\mathbb{G}_{m}}=\coprod_{i=0}^{r}M_{i}. ∎

Proposition 3.2.

Fix a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a reduced homogeneous polynomial of degree rr. If the Euler-symmetric variety MPVPM_{P}\subset\mathbb{P}V_{P} is smooth, then for a general point xMPx\in M_{P}, there exists a line in VP\mathbb{P}V_{P} which lies on MPM_{P}.

Proof.

From Theorem 1.6 we know MPM_{P} is an equivariant compactification of the vector group WW, it suffices to prove that at the original point oMPo\in M_{P}, there exists a line in VP\mathbb{P}V_{P} which lies on MPM_{P}. Also from Theorem 1.6, this can be reduced to prove that V(FP2)V(F^{2}_{P})\neq\emptyset.

Note that from Proposition 2.10, Fzr1=FPr1=dPF^{r-1}_{z}=F^{r-1}_{P_{*}}=\langle dP_{*}\rangle. Let ZZ be the image of V(P)V(P_{*}) of the following rational map:

ΦFzr1:TzMP(V(P))VFzr1(VP),[w][dP(w)].\Phi_{F^{r-1}_{z}}:\mathbb{P}T_{z}M_{P}(\supset V(P_{*}))\dashrightarrow\mathbb{P}V_{F_{z}^{r-1}}(\subset\mathbb{P}V_{P}),[w]\mapsto[dP_{*}(w)].

Since PP is a reduced homogeneous polynomial, PP_{*} is also a reduced homogeneous polynomial. Clearly ZZ is the dual variety of V(P)V(P_{*}), which is non-empty. From the proof of Proposition 2.5, ZZ is fixed by 𝔾m\mathbb{G}_{m} and hence ZM1Z\subset M_{1}.

Write the coordinates of VP\mathbb{P}V_{P} as [z0:z1::zm:w1(2)::wm2(2)::w1(r)][z_{0}:z_{1}:\cdots:z_{m}:w^{(2)}_{1}:\cdots:w^{(2)}_{m_{2}}:\cdots:w^{(r)}_{1}], and let FP2=Q1(2),,Qm2(2)F^{2}_{P}=\langle Q^{(2)}_{1},\cdots,Q^{(2)}_{m_{2}}\rangle generated by Q1(2),,Qm2(2)Q^{(2)}_{1},\cdots,Q^{(2)}_{m_{2}}. Then the following homogeneous polynomials of degree two lie in the vanishing ideal of MPVPM_{P}\subset\mathbb{P}V_{P}:

f1(2)\displaystyle f_{1}^{(2)} =z0w1(2)Q1(2)(z1,,zm);\displaystyle=z_{0}w_{1}^{(2)}-Q_{1}^{(2)}(z_{1},\cdots,z_{m});
\displaystyle\vdots
fm2(2)\displaystyle f_{m_{2}}^{(2)} =z0wm2(2)Qm2(2)(z1,,zm).\displaystyle=z_{0}w_{m_{2}}^{(2)}-Q_{m_{2}}^{(2)}(z_{1},\cdots,z_{m}).

Therefore, ZM1V(FP2)Z\subset M_{1}\subset V(F^{2}_{P}). In particular, V(FP2)V(F^{2}_{P})\neq\emptyset. ∎

Remark 3.3.
  • The existence of a line is equivalent to ord(FP)=1\operatorname{ord}(\textbf{F}_{P})=1 where the order is defined to be the largest natural number kk such that V(FPk)=V(F_{P}^{k})=\emptyset;

  • Furthermore, we have M1=V(FP2)M_{1}=V(F^{2}_{P}). Since every point in V(FP2)V(F^{2}_{P}) corresponds to a line through oo on MPM_{P}, whose terminal point lies in M1M_{1}.

Next we recall the Białynicki-Birula decomposition theorem (over \mathbb{C}) as follows.

Theorem 3.4 ([BB73], Theorem 4.1, 4.2, 4.3).

Let MM be a complete smooth complex manifold with an algebraic 𝔾m\mathbb{G}_{m}-action. Let M𝔾mM^{\mathbb{G}_{m}} be the set of fixed points under the 𝔾m\mathbb{G}_{m}-action and M𝔾m=i𝐈YiM^{\mathbb{G}_{m}}=\bigsqcup_{i\in\mathbf{I}}Y_{i} be the decomposition of M𝔾mM^{\mathbb{G}_{m}} into connected components. Then there exists a unique locally closed 𝔾m\mathbb{G}_{m}-invariant decomposition of MM,

M=i𝐈M+(Yi)=i𝐈M(Yi)M=\bigsqcup_{i\in\mathbf{I}}M^{+}(Y_{i})=\bigsqcup_{i\in\mathbf{I}}M^{-}(Y_{i})

and morphisms γi±:M±(Yi)Yi\gamma^{\pm}_{i}:M^{\pm}(Y_{i})\rightarrow Y_{i}, i𝐈i\in\mathbf{I}, such that the following holds.

  1. (1)

    M±(Yi)M^{\pm}(Y_{i}) is a smooth \mathbb{C}^{*}-invariant complex submanifold of MM. YiY_{i} is a closed complex submanifold of M±(Yi)M^{\pm}(Y_{i}). M+(Yi)M(Yi)=YiM^{+}(Y_{i})\cap M^{-}(Y_{i})=Y_{i}.

  2. (2)

    γi±\gamma^{\pm}_{i} is algebraic and is a \mathbb{C}^{*}-fibration over YiY_{i} (i.e., each fiber is a \mathbb{C}^{*}-module) such that γi±|Yi\gamma^{\pm}_{i}|_{Y_{i}} is the identity, for i𝐈\forall i\in\mathbf{I}.

  3. (3)

    Let Tx(M)+,Tx(M)0,Tx(M)T_{x}(M)^{+},T_{x}(M)^{0},T_{x}(M)^{-} be the weight spaces of the isotropy action on the tangent space Tx(M)T_{x}(M) with positive, zero, negative weights respectively. Then for any xYix\in Y_{i}, the tangent space Tx(M±(Yi))=Tx(M)0Tx(M)±T_{x}(M^{\pm}(Y_{i}))=T_{x}(M)^{0}\oplus T_{x}(M)^{\pm} and the dimension of the fibration given by γi±\gamma_{i}^{\pm} equals dimTx(M)±\dim T_{x}(M)^{\pm}.

We know the Białynicki-Birula decomposition of VP\mathbb{P}V_{P} with respect to the \mathbb{C}^{*}-action given by 𝔾mGL(VP)\mathbb{G}_{m}\subset GL(V_{P}) in Lemma 3.1 can be written as follows:

(VP)𝔾m\displaystyle(\mathbb{P}V_{P})^{\mathbb{G}_{m}} =k=0rVFPk,\displaystyle=\bigsqcup^{r}_{k=0}\mathbb{P}V_{F^{k}_{P}},
(VP)i+\displaystyle(\mathbb{P}V_{P})_{i}^{+} =(k=irVFPk)\(k=i+1rVFPk),\displaystyle=\mathbb{P}(\oplus^{r}_{k=i}V_{F^{k}_{P}})\backslash\mathbb{P}(\oplus^{r}_{k=i+1}V_{F^{k}_{P}}),
(VP)i\displaystyle(\mathbb{P}V_{P})_{i}^{-} =(k=0iVFPk)\(k=0i1VFPk).\displaystyle=\mathbb{P}(\oplus^{i}_{k=0}V_{F^{k}_{P}})\backslash\mathbb{P}(\oplus^{i-1}_{k=0}V_{F^{k}_{P}}).

Then we let Mi+=(VP)i+MPM_{i}^{+}=(\mathbb{P}V_{P})_{i}^{+}\cap M_{P} and Mi=(VP)iMPM_{i}^{-}=(\mathbb{P}V_{P})_{i}^{-}\cap M_{P} respectively. From Iversen’s fixed point theorem [Ive72, Proposition 1.3] and the smoothness of MPM_{P}, we know that (MP)𝔾m=i=1rMi(M_{P})^{\mathbb{G}_{m}}=\bigcup^{r}_{i=1}M_{i} is a smooth subvariety. Hence MiM_{i} are also smooth for any 0ir0\leq i\leq r.

Proposition 3.5.

Fix a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a reduced homogeneous polynomial of degree rr. If the Euler-symmetric variety MPM_{P} is smooth, then M1=ZM_{1}=Z and moreover M1M_{1} is not dual defective, i.e., its dual variety is a hypersurface.

Proof.

By Proposition 2.10, we know that ZVFP2Z\subset\mathbb{P}V_{F_{P}^{2}} is the dual variety of V(P)TzMPV(P_{*})\subset\mathbb{P}T_{z}M_{P}. Therefore, from the proof of Proposition 3.2 and Remark 3.3, we only need to show that M1=ZM_{1}=Z.

Let LzrL_{z}^{r} be the union of all 𝔾m\mathbb{G}_{m}-stable rational curves through zz whose tangent direction at zz lies in Sm(V(P))=V(P)(dP0)TzMP\operatorname{Sm}(V(P_{*}))=V(P_{*})\cap(dP_{*}\neq 0)\subset\mathbb{P}T_{z}M_{P}. Then we have

dimLzr=dim(V(P))+1=n1.\operatorname{dim}L_{z}^{r}=\operatorname{dim}(V(P_{*}))+1=n-1.

From the definition of LzrL_{z}^{r}, we know that Lzr\{z}M1+L_{z}^{r}\backslash\{z\}\subset M_{1}^{+}, hence ZZ is a union of some irreducible components of M1M_{1}.

Since M1M_{1} is smooth, the irreducible components of M1M_{1} do not intersect. Let M1=ZZ1M_{1}=Z\cup Z_{1}. Suppose that Z1Z_{1}\neq\emptyset, then Z1+¯\overline{Z_{1}^{+}} is also a divisor. By the definition of ZZ (see the proof of Proposition 3.2), ZZ is actually the set of source points of the 𝔾m\mathbb{G}_{m}-stable rational curves on MPM_{P} through zz with degree r1r-1. Since Z1Z=Z_{1}\cap Z=\emptyset, we know zZ1+¯z\notin\overline{Z_{1}^{+}}. Write the coordinates of VP\mathbb{P}V_{P} as [z0:z1::zm:w1(2)::wm2(2)::w1(r)][z_{0}:z_{1}:\cdots:z_{m}:w^{(2)}_{1}:\cdots:w^{(2)}_{m_{2}}:\cdots:w^{(r)}_{1}], then Z1+¯\overline{Z_{1}^{+}} is contained in the hyperplane {w1(r)=0}\{w_{1}^{(r)}=0\}. Also we know Z1+¯{z0=0}\overline{Z_{1}^{+}}\subset\{z_{0}=0\}. This implies that dim(Z1+¯)m2\dim(\overline{Z^{+}_{1}})\leq m-2, which contradicts with dim(Z1+¯)=m1\operatorname{dim}(\overline{Z_{1}^{+}})=m-1. Therefore, M1=ZM_{1}=Z. ∎

Proposition 3.6.

Let the boundary divisor D=jJDjD=\cup_{j\in J}D_{j} and DjD_{j} be irreducible components of DD, for all jJj\in J. Then the terminal point zDjz\in D_{j}, for any jJj\in J.

Proof.

From [HT99, Proposition 2.3, Theorem 2.5], we know that the action WW on boundary divisor DD stabilizes all its irreducible components. Let D1D_{1} be an irreducible component of DD, and let p=[0::0:fj::fr]p=[0:\cdots:0:f^{j}:\cdots:f^{r}] be a point of D1D_{1} where each fi(jir)f^{i}(j\leq i\leq r) is viewed as an element in (FPi)(F^{i}_{P})^{*} and fj0f^{j}\neq 0. Then j1j\geq 1.

For vWv\in W, define

vp={gtvpt}¯.\mathcal{L}_{v}^{p}=\overline{\{g_{tv}\cdot p\mid t\in\mathbb{C}\}}.

Therefore, vp\mathcal{L}_{v}^{p} is a rational curve and vpD1\mathcal{L}_{v}^{p}\subset D_{1} for any vector vWv\in W.

From the explicit expression of the action of WW on VP\mathbb{P}V_{P} (see Remark 1.7), we know that

gtvp=[0::0:gtvpfj::gtvpfr],g_{tv}\cdot p=[0:\cdots:0:g_{tv}^{p}\cdot f^{j}:\cdots:g_{tv}^{p}\cdot f^{r}],

where for each jirj\leq i\leq r,

gtvpfi==jiCifιtvi==jitiCifιvi.g_{tv}^{p}\cdot f^{i}=\sum_{\ell=j}^{i}C_{i}^{\ell}f^{\ell}\circ\iota_{tv}^{i-\ell}=\sum_{\ell=j}^{i}t^{i-\ell}C_{i}^{\ell}f^{\ell}\circ\iota_{v}^{i-\ell}.

Since FPjF^{j}_{P} consists of all (rj)(r-j)-th derivatives of PP and j1j\geq 1, for general vWv\in W, fjιvrj0f^{j}\circ\iota_{v}^{r-j}\neq 0. Hence gtvpfr0g_{tv}^{p}\cdot f^{r}\neq 0 and the rational curve vp\mathcal{L}_{v}^{p} contains the terminal point zz, then we have zvpD1z\in\mathcal{L}_{v}^{p}\subset D_{1}. ∎

Remark 3.7.
  • For a non-reduced polynomial PP, this proposition also holds.

  • For general homogeneous polynomial PP, the smoothness of the Euler-symmetric variety MPM_{P} only depends on the point zz. That is to say, the point zz is ’the most singular’ point of MPM_{P}, and if the point zz is smooth then MPM_{P} is smooth variety.

4. Proof of Theorem 1.10

Note that Theorem 1.4 is only applicable for general points, then a prior we do not know if Fz\textbf{F}_{z} is also a symbol system. [EN23, Theorem 3.3] is essential to obtain that Fz\textbf{F}_{z} is actually a symbol system. For the reader’s convenience we restate [EN23, Theorem 3.3] as follows.

Let XNX\subset\mathbb{P}^{N} be a projective variety and L=𝒪X(1)L=\mathcal{O}_{X}(1) be the restriction of hyperplane line bundle of N\mathbb{P}^{N} on XX. We denote Jp(L)J^{p}(L) the sheaf of pp-jets of LL. We have a natural short exact sequence

0SympTXLJp(L)πp1pJp1(L)0,0\rightarrow\operatorname{Sym}^{p}T^{*}X\otimes L\rightarrow J^{p}(L)\stackrel{{\scriptstyle\pi_{p-1}^{p}}}{{\rightarrow}}J^{p-1}(L)\rightarrow 0,

where TXT^{*}X is the cotangent sheaf of XX and πp1p\pi_{p-1}^{p} is the truncation map. Let

αi:H0(N,𝒪N(1))𝒪XJi(L)\alpha_{i}:H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1))\otimes\mathcal{O}_{X}\rightarrow J^{i}(L)

be the Taylor series map by taking terms with order i\leq i in the Taylor series of a global section of LL. Define 𝐔k\mathbf{U}_{k} to be the maximal open subset of XX contained in the smooth locus of XX such that the quotient sheaf Ji(L)/Im(αi)J^{i}(L)/\operatorname{Im}(\alpha_{i}) is locally free of constant rank over 𝐔k\mathbf{U}_{k} for all iki\leq k. And the kernel of the restriction of the truncation map πp1p:Im(αp)Im(αp1)\pi_{p-1}^{p}:\operatorname{Im}(\alpha_{p})\to\operatorname{Im}(\alpha_{p-1}) is the pp-th fundamental form. [EN23, Theorem 3.3] can be interpreted as

Theorem 4.1.

The system of fundamental forms 𝐅xk:=i=0kFxi\mathbf{F}_{x}^{\leq k}:=\oplus_{i=0}^{k}F_{x}^{i} is a symbol system for any x𝐔kx\in\mathbf{U}_{k}.

From this we obtain

Lemma 4.2.

Fix a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a reduced homogeneous polynomial of degree rr. If the Euler-symmetric variety MPM_{P} is smooth, then the system of fundamental forms 𝐅z\mathbf{F}_{z} is a symbol system.

Proof.

From above theorem, we only need to prove that the terminal point z𝐔rz\in\mathbf{U}_{r}. By upper semi-continuity of the rank function of a coherent sheaf (cf.[Har79, Chapter III 12.7.2]), for any point qMP\𝐔rq\in M_{P}\backslash\mathbf{U}_{r}, there exists 2ir2\leq i\leq r such that

rankq(Ji(L)/Im(αi))>rankx(Ji(L)/Im(αi)),\operatorname{rank}_{q}(J^{i}(L)/\operatorname{Im}(\alpha_{i}))>\operatorname{rank}_{x}(J^{i}(L)/\operatorname{Im}(\alpha_{i})),

for x𝐔rx\in\mathbf{U}_{r}. Suppose that z𝐔rz\notin\mathbf{U}_{r}, then z𝐔j\𝐔j+1z\in\mathbf{U}_{j}\backslash\mathbf{U}_{j+1} for some 1jr11\leq j\leq r-1. Therefore, we have

rankz(Jj+1(L)/Im(αj+1))\displaystyle\operatorname{rank}_{z}(J^{j+1}(L)/\operatorname{Im}(\alpha_{j+1})) >rankx(Jj+1(L)/Im(αj+1));\displaystyle>\operatorname{rank}_{x}(J^{j+1}(L)/\operatorname{Im}(\alpha_{j+1}));
rankz(Jk(L)/Im(αk))\displaystyle\operatorname{rank}_{z}(J^{k}(L)/\operatorname{Im}(\alpha_{k})) =rankx(Jk(L)/Im(αk)),kj,\displaystyle=\operatorname{rank}_{x}(J^{k}(L)/\operatorname{Im}(\alpha_{k})),\forall\;k\leq j,

for x𝐔jx\in\mathbf{U}_{j}. Since we have the following exact sequence at point zz

0Fzj+1Im(αj+1)zIm(αj)z0,0\to F^{j+1}_{z}\to\operatorname{Im}(\alpha_{j+1})_{z}\to\operatorname{Im}(\alpha_{j})_{z}\to 0,

we have dim(Fzj+1)<dim(Fxj+1)\operatorname{dim}(F^{j+1}_{z})<\operatorname{dim}(F^{j+1}_{x}), which contradicts with Proposition 2.5. ∎

Then we immediately have

Proposition 4.3.

Fix a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a reduced homogeneous polynomial of degree rr. If the Euler-symmetric variety MPM_{P} is smooth, then MPM_{P} is homogeneous.

Proof.

Consider the rational map ΨFz\Psi_{\textbf{F}_{z}} defined by the system of fundamental forms of MPM_{P} at zz:

ΨFz:(W)((Fzr)(Fz2)W),\Psi_{\textbf{F}_{z}}:\mathbb{P}(\mathbb{C}\oplus W^{*})\dashrightarrow\mathbb{P}((F_{z}^{r})^{*}\oplus\cdots\oplus(F_{z}^{2})^{*}\oplus W^{*}\oplus\mathbb{C}),
[t:w][ιwr:tιwr1::tr2ιw2:tr1w:tr].[t:w]\mapsto[\iota^{r}_{w}:t\iota^{r-1}_{w}:\cdots:t^{r-2}\iota^{2}_{w}:t^{r-1}w:t^{r}].

In terms of the natural identification of ((Fzr)(Fz2)W)\mathbb{P}((F_{z}^{r})^{*}\oplus\cdots\oplus(F_{z}^{2})^{*}\oplus W^{*}\oplus\mathbb{C}) and (VP)\mathbb{P}(V_{P}), MP=Im(ΨFz)M_{P}=\operatorname{Im}(\Psi_{\textbf{F}_{z}}). Since Fz\textbf{F}_{z} is a symbol system, the WW^{*}-action on (W)\mathbb{P}(\mathbb{C}\oplus W^{*}) can be extended to MPM_{P}, making the orbit of zz is biregular to WW^{*}. Let UzU_{z} denote the open orbit of zz. By Proposition 3.6, we know that for any point pDp\in D in the boundary divisor, there exists a vector vWv\in W such that gvpUzg_{v}\cdot p\in U_{z}.

Since UzUMPU_{z}\cap U\subset M_{P} is also an open subset of MPM_{P}, there exists a vector wWw\in W^{*} such that gw(gvp)Ug_{w}\cdot(g_{v}\cdot p)\in U. Hence MPM_{P} is homogeneous under the algebraic group generated by WW and WW^{*} in GL(VP)GL(V_{P}). ∎

Proposition 4.4.

Fix a vector space WW. Let PSymrWP\in\operatorname{Sym}^{r}W^{*} be a reduced homogeneous polynomial of degree rr. If the Euler-symmetric variety MPM_{P} is smooth, then the Picard number ρ(MP)\rho(M_{P}) equals the number of irreducible components of M1M_{1}, namely,

ρ(MP)=#irrM1,\rho(M_{P})=\#_{irr}M_{1},

where #irrY\#_{irr}Y is the number of irreducible components of YY.

Proof.

From the proof of Proposition 3.5 and [HT99, Theorem 2.5], we have

ρ(MP)#irrM1.\rho(M_{P})\geq\#_{irr}M_{1}.

If ρ(MP)#irrM1\rho(M_{P})\neq\#_{irr}M_{1}, then M1+¯D\overline{M_{1}^{+}}\subsetneq D, and there exist an irreducible component D1D_{1} of DD and an irreducible component KK of MjM_{j}, for some 2jr12\leq j\leq r-1, such that D1=K+¯D_{1}=\overline{K^{+}}.

Let LzkL_{z}^{k} be the closure of the union of all 𝔾m\mathbb{G}_{m}-stable rational curves through zz such that the tangent direction at zz lies in V(Fzk)TzMPV(F_{z}^{k})\subset\mathbb{P}T_{z}M_{P}. Therefore, D1Lzrj+1D_{1}\subset L_{z}^{r-j+1} by Proposition 3.6. But we have

dim(Lzrj+1)dim(V(Fzrj+1))+1dim(V(Fzr1))+1n2,\operatorname{dim}(L_{z}^{r-j+1})\leq\operatorname{dim}(V(F_{z}^{r-j+1}))+1\leq\operatorname{dim}(V(F_{z}^{r-1}))+1\leq n-2,

which contradicts with the fact that D1D_{1} is a divisor. ∎

Remark 4.5.

Consider the negative Białynicki-Birula decomposition for the 𝔾m\mathbb{G}_{m}-action MP=i=0rMiM_{P}=\coprod_{i=0}^{r}M_{i}^{-}. we have the similar results as follows:

  1. (1)

    Mr=zM_{r}=z, the terminal point;

  2. (2)

    MrM_{r}^{-} is biregular to an affine space of dimension nn, and the boundary divisor Dz=i=0r1Mi=Mr1¯D_{z}=\coprod_{i=0}^{r-1}M_{i}^{-}=\overline{M^{-}_{r-1}};

  3. (3)

    Mr1VFr1M_{r-1}\subset\mathbb{P}V_{F^{r-1}} is also the dual variety of V(P)ToMPV(P)\subset\mathbb{P}T_{o}M_{P};

  4. (4)

    all the irreducible component of DzD_{z} contain the original point oo;

  5. (5)

    ρ(MP)=#irrMr1\rho(M_{P})=\#_{irr}M_{r-1}.

Corollary 4.6.

Assumption as above, then we have

ρ(MP)=#irrV(P)=#irrV(P).\rho(M_{P})=\#_{irr}V(P)=\#_{irr}V(P_{*}).
Proof.

Since M0+M_{0}^{+} and MrM_{r}^{-} are affine open subvariety which is biregular to affine space, then

#irrD=ρ(MP)=#irrDz.\#_{irr}D=\rho(M_{P})=\#_{irr}D_{z}.

From Theorem 4.4 and Remark 4.5, we know that Lor=DzL_{o}^{r}=D_{z}, Lzr=DL_{z}^{r}=D. From definition, LorL_{o}^{r} is also the closure of the 𝔾m\mathbb{G}_{m}-fibration Lor~\widetilde{L_{o}^{r}}:

Lor~\textstyle{\widetilde{L_{o}^{r}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M0+\o\textstyle{M_{0}^{+}\backslash o\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sm(V(P))\textstyle{\operatorname{Sm}(V(P))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ToMP.\textstyle{\mathbb{P}T_{o}M_{P}.}

Therefore, #irrLor=#irrLor~=#irrSm(V(P))=#irrV(P)\#_{irr}L^{r}_{o}=\#_{irr}\widetilde{L_{o}^{r}}=\#_{irr}\operatorname{Sm}(V(P))=\#_{irr}V(P). Similarly, #irrLzr=#irrV(P)\#_{irr}L^{r}_{z}=\#_{irr}V(P_{*}). ∎

Now we can finish the proof of Theorem 1.10.

Proof.

From Proposition 4.3, MPM_{P} is homogeneous. Since Euler-symmetric varieties are rational, hence MPM_{P} is rational homogeneous and it is actually a Hermitian symmetric space as it is an equivariant compactification of a vector group (see [Arz11]). As PP is reduced, V(P)V(P) is reduced. MPM_{P} is a product of irreducible compact Hermitian symmetric spaces, then M1M_{1} is the disjoint union of the VMRTs of each factors. Since M1M_{1} is not dual defect, we know each factor of MPM_{P} must be isomorphic to an irreducible compact Hermitian symmetric spaces listed in Theorem 1.10. The embedding is obvious.

For the converse part, it can be deduced directly from the irreducible case, which is well-known (see for example [LM02]). ∎

Remark 4.7.

Even if the polynomial PP is non-reduced, the Proposition 4.3 still holds. Consequently, we can relax the requirement that PP is reduced in Theorem 1.10. However, this only allows us to conclude that MPM_{P} can be factored as a product of those irreducible Hermitian symmetric spaces listed in Theorem 1.10, without specifying the exact embedding.

Acknowledgement

The authors would like to thank Jun-Muk Hwang and Baohua Fu for some useful suggestions and comments. The second author would also like to thank Wenbo Niu for introducing his work and helpful discussions. The first author was supported by a start-up funding of Shenzhen University (000001032064). The second author was supported by the Institute for Basic Science (IBS-R032-D1-2024-a00).

References

  • [Arz11] Ivan V. Arzhantsev. Flag varieties as equivariant compactifications of 𝔾an\mathbb{G}^{n}_{a}. Proc. Amer. Math. Soc., 139(3):783–786, 2011.
  • [BB73] A. Białynicki-Birula. Some theorems on actions of algebraic groups. Ann. of Math. (2), 98:480–497, 1973.
  • [EKP02] Pavel Etingof, David Kazhdan, and Alexander Polishchuk. When is the fourier transform of an elementary function elementary? Selecta Math. (N.S.), 8(1):27, 2002.
  • [EN23] Lawrence Ein and Wenbo Niu. On vanishing of fundamental forms of algebraic varieties. arXiv preprint arXiv:2304.08430, 2023.
  • [FH20] Baohua Fu and Jun-Muk Hwang. Euler-symmetric projective varieties. Algebr. Geom., 7(3), 2020.
  • [GL19] Fulvio Gesmundo and Joseph M. Landsberg. Explicit polynomial sequences with maximal spaces of partial derivatives and a question of K. Mulmuley. Theory Comput., 15(3):1–24, 2019.
  • [Har79] Robin Hartshorne. Algebraic geometry, volume 52. Springer Science & Business Media, 1979.
  • [HT99] Brendan Hassett and Yuri Tschinkel. Geometry of equivariant compactifications of 𝔾an\mathbb{G}_{a}^{n}. Int. Math. Res. Not., 22:1211–1230, 1999.
  • [Ive72] Birger Iversen. A fixed point formula for action of tori on algebraic varieties. Invent. Math., 16:229–236, 1972.
  • [LM02] Joseph M. Landsberg and Laurent Manivel. Construction and classification of complex simple Lie algebras via projective geometry. Selecta Math. (N.S.), 8:137–159, 2002.
  • [Luo23] Zhijun Luo. Euler-symmetric complete intersections in projective space. J. Algebra, 624:41–62, 2023.
  • [Sha23] Anton Shafarevich. Euler-symmetric projective toric varieties and additive actions. Indag. Math. (N.S.), 34(1):42–53, 2023.