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

Vector Bundles on Rational Homogeneous Spaces

Rong Du , Xinyi Fang and Yun Gao School of Mathematical Sciences Shanghai Key Laboratory of PMMP, East China Normal University, Rm. 312, Math. Bldg, No. 500, Dongchuan Road, Shanghai, 200241, P. R. China, [email protected]. The Research is Sponsored by National Natural Science Foundation of China (Grant No. 11531007), Natural Science Foundation of China and the Israel Science Foundation (Grant No. 11761141005) and Science and Technology Commission of Shanghai Municipality (Grant No. 18dz2271000).School of Mathematical Sciences Shanghai Key Laboratory of PMMP, East China Normal University, No. 500, Dongchuan Road, Shanghai, 200241, P. R. China, [email protected]. The Research is Sponsored by National Natural Science Foundation of China (Grant No. 11531007) and Science and Technology Commission of Shanghai Municipality (Grant No. 18dz2271000).School of Mathematical Sciences, Shanghai Jiao Tong University, Shanghai 200240, P. R. of China, [email protected]. The Research is Sponsored by National Natural Science Foundation of China (Grant No. 11531007).
Abstract

We consider a uniform rr-bundle EE on a complex rational homogeneous space XX and show that if EE is poly-uniform with respect to all the special families of lines and the rank rr is less than or equal to some number that depends only on XX, then EE is either a direct sum of line bundles or δi\delta_{i}-unstable for some δi\delta_{i}. So we partially answer a problem posted by Muñoz-Occhetta-Solá Conde(19). In particular, if XX is a generalized Grassmannian 𝒢\mathcal{G} and the rank rr is less than or equal to some number that depends only on XX, then EE splits as a direct sum of line bundles. We improve the main theorem of Muñoz-Occhetta-Solá Conde (18 Theorem 3.1) when XX is a generalized Grassmannian by considering the Chow ring. Moreover, by calculating the relative tangent bundles between two rational homogeneous spaces, we give explicit bounds for the generalized Grauert-Mülich-Barth theorem on rational homogeneous spaces.

Dedicate to people who devote their lives to fight against coronaviruses

Key words: vector bundle, generalized Grassmannian, rational homogeneous space

1 Introduction

Algebraic vector bundles on a projective variety XX over complex number field \mathbb{C} are fundamental research objects in algebraic geometry. However, up to now, algebraic vector bundles are still mysterious for general projective varieties. According to Serre (GAGA), the classification of algebraic vector bundles over \mathbb{C} is equivalent to the classification of holomorphic vector bundles. So there are not only algebraic ways but also analytic ways to handle the problems of vector bundles. For simplicity, we just call vector bundles of rank rr or rr-bundles in the context.

If XX is 1\mathbb{P}^{1}, the structure of a vector bundle on XX is quite clear because Grothendieck tells us that it splits as a direct sum of line bundles. However, if XX is a projective space and the dimension of it is bigger than or equal to two then the structures of vector bundles on XX are not so easy to be determined. Since projective spaces are covered by lines, it is a natural way to consider the restriction of vector bundles to lines. From Grothendieck’s result, they split after being restricted to 1\mathbb{P}^{1}. By semicontinuity theorem, for “almost all” lines, rr-bundle EE has the constant splitting type. This means that the lines to which the vector bundle restricts having different splitting type consist a closed subset of a Grassmannian. If the closed subset is empty, such bundles are called uniform vector bundles. Uniform bundles are widely studied not only on projective spaces (25 27 24 9 10 11 3) but also on special Fano manifolds of Picard number one (2 15 12 6 18). Please see Introduction in 6 for the details.

Instead of considering vector bundles, Occhetta-Solá Conde-Wiśniewski (22) studied flag bundles which are constructed upon the action of the defining group GG on the flag manifold G/BG/B. Recently, Muñoz-Occhetta-Solá Conde (17) studied uniform principle GG-bundles with GG semisimple over Fano manifolds. They present a number of theorems that are flag bundle’s versions of some of the central results in the theory of uniform vector bundles. More precisely, they paid special attention to homogeneous filtrations of relative tangent bundles between flag manifolds and generalized the different standard decomposability notions of vector bundles. They used an interesting concept of "tag" of a G/BG/B-bundle to describe diagonalizability of any uniform flag bundle of low rank. But they mainly focus on XX with Picard number one. In 19, the authors proposed a problem as follows.

Problem 1.1.

Classify low rank uniform principle GG-bundles (GG semisimple algebraic group) on rational homogeneous spaces.

In this paper, we consider uniform bundles on generalized flag varieties or even rational homogeneous spaces with arbitrary Picard numbers and give partial answers to this problem.

Let

X=G/PG1/PI1×G2/PI2××Gm/PIm,X=G/P\simeq G_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}},

where GiG_{i} is a simple Lie group with Dynkin diagram 𝒟i\mathcal{D}_{i} whose set of nodes is DiD_{i} and PIiP_{I_{i}} is a parabolic subgroup of GiG_{i} corresponding to IiDiI_{i}\subset D_{i}. We set F(Ii):=Gi/PIiF(I_{i}):=G_{i}/P_{I_{i}} by marking on the Dynkin diagram 𝒟i\mathcal{D}_{i} of GiG_{i} the nodes corresponding to IiI_{i}. Let δi\delta_{i} be a node in 𝒟i\mathcal{D}_{i} and N(δi)N(\delta_{i}) be the set of nodes in 𝒟i\mathcal{D}_{i} that are connected to δi\delta_{i}.

If δiIi\delta_{i}\in I_{i}, we call

iδic:=Gi/Piδic×Gi/PIi^(1im),\mathcal{M}_{i}^{\delta_{i}^{c}}:=G_{i}/P_{i}^{\delta_{i}^{c}}\times\widehat{G_{i}/P_{I_{i}}}~{}(1\leq i\leq m),

the ii-th special family of lines of class δˇi\check{\delta}_{i}, where Piδic:=PIi\δiN(δi)P_{i}^{\delta_{i}^{c}}:=P_{I_{i}\backslash\delta_{i}\cup N(\delta_{i})} and Gi/PIi^\widehat{G_{i}/P_{I_{i}}} is G1/PI1×G2/PI2××Gm/PImG_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}} by deleting ii-th term Gi/PIiG_{i}/P_{I_{i}}. Denote by

𝒰iδic:=Gi/PIiN(δi)×Gi/PIi^\mathcal{U}_{i}^{\delta_{i}^{c}}:=G_{i}/P_{I_{i}\cup N(\delta_{i})}\times\widehat{G_{i}/P_{I_{i}}}

the ii-th universal family of class δˇi\check{\delta}_{i}, which has a natural 1\mathbb{P}^{1}-bundle structure over iδic\mathcal{M}_{i}^{\delta_{i}^{c}}.

We separate our discussion into two cases:

Case I: N(δi)IiN(\delta_{i})\subseteq I_{i}, then 𝒰iδic=X\mathcal{U}_{i}^{\delta_{i}^{c}}=X and we have the natural projection XiδicX\rightarrow\mathcal{M}_{i}^{\delta_{i}^{c}};

Case II: N(δi)IiN(\delta_{i})\nsubseteq I_{i}, then we have the standard diagram

(1.5)

For the definition of "poly-uniform", "δi\delta_{i}-stable", "δi\delta_{i}-unstable", "ς(𝒢)\varsigma(\mathcal{G})" and "ν(X)\nu(X)", please see Section 3.

Theorem 1.2.

On XX, if an rr-bundle EE is poly-uniform with respect to all the special families of lines and rν(X)2r\leq\nu(X)-2, then EE is δi\delta_{i}-unstable for some δi\delta_{i} (1im1\leq i\leq m) or EE splits as a direct sum of line bundles.

In particular, if the Picard number of XX is one, we improve an interesting theorem of Muñoz-Occhetta-Solá Conde (18 Theorem 3.1) a little bit (see Table 2).

Theorem 1.3.

Suppose that EE is a uniform rr-bundle on a generalized Grassmann 𝒢\mathcal{G}. If rς(𝒢)r\leq\varsigma(\mathcal{G}), then EE splits as a direct sum of line bundles.

By calculating the relative tangent bundles and using Descent Lemma, we can have explicit bounds for the generalized Grauert-Mülich-Barth theorem on rational homogeneous spaces.

Theorem 1.4.

Fix δiIi\delta_{i}\in I_{i} and assume that αδi\alpha_{\delta_{i}} is not an exposed short root. Let EE be a holomorphic rr-bundle over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),~{}a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. If for some t<rt<r,

at(δi)at+1(δi){1,andN(δi)fits Case I2,andN(δi)fits Case II,a_{t}^{(\delta_{i})}-a_{t+1}^{(\delta_{i})}\geq\left\{\begin{array}[]{ll}1,&and~{}N(\delta_{i})~{}\text{fits Case I}\\ 2,&and~{}N(\delta_{i})~{}\text{fits Case II},\end{array}\right.

then there is a normal subsheaf KEK\subset E of rank tt with the following properties: over the open set VE=q1(q21(UE(δi)))XV_{E}=q_{1}(q_{2}^{-1}(U_{E}^{(\delta_{i})}))\subset X, where UE(δi)U_{E}^{(\delta_{i})} is an open set in δic\mathcal{M}^{\delta_{i}^{c}}, the sheaf KK is a subbundle of EE, which on the line LXL\subset X given by lUE(δi)l\in U_{E}^{(\delta_{i})} has the form

K|Ls=1t𝒪L(as(δi)).K|_{L}\cong\oplus_{s=1}^{t}\mathcal{O}_{L}(a_{s}^{(\delta_{i})}).
Corollary 1.5.

With the same assumption as Theorem 1.4. For a δi\delta_{i}-semistable rr-bundle EE over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}, we have

as(δi)as+1(δi)1for alls=1,,r1.a_{s}^{(\delta_{i})}-a_{s+1}^{(\delta_{i})}\leq 1~{}~{}\text{for all}~{}s=1,\ldots,r-1.

In particular, if N(δi)N(\delta_{i}) fits Case I, then we have as(δi)a_{s}^{(\delta_{i})}’s are constant for all 1sr1\leq s\leq r.

Theorem 1.6.

Fix δiIi\delta_{i}\in I_{i} and assume that αδi\alpha_{\delta_{i}} is an exposed short root. Let EE be a holomorphic rr-bundle over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),~{}a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. If for some t<rt<r,

at(δi)at+1(δi){1,andN(δi)fits Case I4,andN(δi)fits Case II,a_{t}^{(\delta_{i})}-a_{t+1}^{(\delta_{i})}\geq\left\{\begin{array}[]{ll}1,&and~{}N(\delta_{i})~{}\text{fits Case I}\\ 4,&and~{}N(\delta_{i})~{}\text{fits Case II},\end{array}\right.

then there is a normal subsheaf KEK\subset E of rank tt with the following properties: over the open set VE=q1(q21(UE(δi)))XV_{E}=q_{1}(q_{2}^{-1}(U_{E}^{(\delta_{i})}))\subset X, where UE(δi)U_{E}^{(\delta_{i})} is an open set in δic\mathcal{M}^{\delta_{i}^{c}}, the sheaf KK is a subbundle of EE, which on the line LXL\subset X given by lUE(δi)l\in U_{E}^{(\delta_{i})} has the form

K|Ls=1t𝒪L(as(δi)).K|_{L}\cong\oplus_{s=1}^{t}\mathcal{O}_{L}(a_{s}^{(\delta_{i})}).
Corollary 1.7.

With the same assumption as Theorem 1.6. For a δi\delta_{i}-semistable rr-bundle EE over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}, we have

as(δi)as+1(δi)3for alls=1,,r1.a_{s}^{(\delta_{i})}-a_{s+1}^{(\delta_{i})}\leq 3~{}~{}\text{for all}~{}s=1,\ldots,r-1.

In particular, if N(δi)N(\delta_{i}) fits Case I, then we have as(δi)a_{s}^{(\delta_{i})}’s are constant for all 1sr1\leq s\leq r.

If X=G/BX=G/B, where GG is a semi-simple Lie group and BB is a Borel subgroup of GG, then we have the following result.

Corollary 1.8.

If an rr-bundle EE on X is δi\delta_{i}-semistable for all ii and δi(1im)\delta_{i}~{}(1\leq i\leq m), then EE splits as a direct sum of line bundles.

2 Preliminaries

Throughout this paper, all algebraic varieties and morphisms will be defined over complex number field \mathbb{C}.

2.1 Semisimple Lie groups and algebras

In mathematics, Lie group–Lie algebra correspondence allows one to study Lie groups, which are geometric objects, in terms of Lie algebras, which are linear objects. Let GG be a semi-simple Lie group. Assume VV to be a nonzero finite dimensional complex vector space, fix a maximal torus HGH\subset G, and let φ:GGL(V)\varphi:G\rightarrow GL(V) be a representation of GG. It is well known that VV decomposes into a direct sum of simultaneous eigenspaces

V=Vλ,V=\bigoplus V_{\lambda},

where the direct sum run over λ\lambda in the character group of HH, which is the set of all holomorphic homomorphisms λ\lambda from HH to \mathbb{C}^{*}, and

Vλ={vV|φ(h)=λ(h)v,for allhH}.V_{\lambda}=\{v\in V|\varphi(h)=\lambda(h)v,~{}\text{for all}~{}h\in H\}.

Since VV is a finite dimensional vector space, we have Vλ=0V_{\lambda}=0 for all but finitely many values of λ\lambda. Those values of λ\lambda for which Vλ0V_{\lambda}\neq 0 are called the weights of VV, and VλV_{\lambda} is called the weight space.

In the Lie algebra side, let 𝔤\mathfrak{g} be the associated semi-simple Lie algebra of GG. The maximal torus corresponds to 𝔥𝔤\mathfrak{h}\subset\mathfrak{g} which is an abelian subalgebra of maximal dimension, i.e. Cartan subalgebra. A holomorphic representation φ:GGL(V)\varphi:G\rightarrow GL(V) of a complex Lie group GG gives rise to a complex linear representation φ𝔤:𝔤𝔤𝔩(V)\varphi_{\mathfrak{g}}:\mathfrak{g}\rightarrow\mathfrak{gl}(V) of the Lie algebra 𝔤\mathfrak{g} of GG. Similarly, every finite dimensional representation of 𝔤\mathfrak{g} admits a decomposition

V=λ{λ𝔥|Vλ0}Vλ,V=\bigoplus_{\lambda\in\{\lambda\in\mathfrak{h}^{\vee}|V_{\lambda}\neq 0\}}V_{\lambda},

where

Vλ={vV|[h,v]=λ(h)v,for allh𝔥}.V_{\lambda}=\{v\in V|[h,v]=\lambda(h)v,~{}\text{for all}~{}h\in\mathfrak{h}\}.

Those λ\lambda are still called the weights of VV, and VλV_{\lambda} is called the weight space. If we apply the above decomposition to V=𝔤V=\mathfrak{g} and φ𝔤\varphi_{\mathfrak{g}} the adjoint representation, we have Cartan decomposition of 𝔤\mathfrak{g}:

𝔤=𝔥α𝔥\{0}𝔤α,\mathfrak{g}=\mathfrak{h}\oplus\bigoplus_{\alpha\in\mathfrak{h}^{\vee}\backslash\{0\}}\mathfrak{g}_{\alpha},

where

𝔤α:={g𝔤|[h,g]=α(h)g,for allh𝔥}.\mathfrak{g}_{\alpha}:=\{g\in\mathfrak{g}|[h,g]=\alpha(h)g,~{}\text{for all}~{}h\in\mathfrak{h}\}.

The elements α𝔥\{0}\alpha\in\mathfrak{h}^{\vee}\backslash\{0\} for which 𝔤α0\mathfrak{g}_{\alpha}\neq 0 are called roots of 𝔤\mathfrak{g}, and the set of these elements will be denoted by Φ\Phi and be called root system. For every αΦ\alpha\in\Phi, 𝔤α\mathfrak{g}_{\alpha} is called root space which is one dimensional.

Fix a linear functional

f:spanΦf:\text{span}_{\mathbb{R}}\Phi\rightarrow\mathbb{R}

whose kernel does not intersect Φ\Phi. Let

Φ+:={αΦ|f(α)>0}andΦ:={αΦ|f(α)<0}.\Phi^{+}:=\{\alpha\in\Phi|f(\alpha)>0\}~{}\text{and}~{}\Phi^{-}:=\{\alpha\in\Phi|f(\alpha)<0\}.

Then Φ+\Phi^{+} is called positive system of roots and Φ\Phi^{-} is called negative system of roots. Given such a positive system Φ+\Phi^{+}, we define the fundamental system ΠΦ+\Pi\subset\Phi^{+} as follows: αΠ\alpha\in\Pi if and only if αΦ+\alpha\in\Phi^{+} and α\alpha cannot be expressed as the sum of two elements of Φ+\Phi^{+}. A non-zero representation VV of 𝔤\mathfrak{g} is called a highest weight representation if it is generated by a vector vVλv\in V_{\lambda} such that gv=0gv=0 for all gαΦ+𝔤αg\in\bigoplus_{\alpha\in\Phi^{+}}\mathfrak{g}_{\alpha}. In this case, vv is called the highest weight vector, and λ\lambda is the highest weight of VV.

For each αΦ\alpha\in\Phi there is a unique element hα𝔥h_{\alpha}\in\mathfrak{h} such that

α(h)=<hα,h>for allh𝔥.\alpha(h)=<h_{\alpha},h>~{}\text{for all}~{}h\in\mathfrak{h}.

The vectors hαh_{\alpha} for αΦ\alpha\in\Phi span 𝔥\mathfrak{h}. We denote by 𝔥\mathfrak{h}_{\mathbb{R}} the set of all elements of form i=1laihα\sum_{i=1}^{l}a_{i}h_{\alpha} for aia_{i}\in\mathbb{R}.

The Killing form <α,β>:=tr(adαadβ)<\alpha,\beta>:=tr(\text{ad}_{\alpha}\circ\text{ad}_{\beta}) defines a nondegenerated bilinear form on 𝔥\mathfrak{h}, where α,β𝔤\alpha,\beta\in\mathfrak{g} and ad is the adjoint representation. It can be shown that 𝔥\mathfrak{h}_{\mathbb{R}}^{\vee} is a Euclidean space with respect to <,><,>. Set n:=dim(𝔥)n:=dim_{\mathbb{C}}(\mathfrak{h}) and D:={1,2,,n}D:=\{1,2,\ldots,n\}. We identify DD with the set of fundamental roots with respect to a choice of maximal torus TT and fixed Borel subgroup BB. It is known that every fundamental system ΠΦ\Pi\subset\Phi can form a basis of 𝔥\mathfrak{h}_{\mathbb{R}}^{\vee}. Let Π={α1,,αn}\Pi=\{\alpha_{1},\ldots,\alpha_{n}\} be a fundamental system. Then we define AijA_{ij} by

Aij=2<αi,αj><αi,αi>,i,j=1,n.A_{ij}=2\frac{<\alpha_{i},\alpha_{j}>}{<\alpha_{i},\alpha_{i}>}\in\mathbb{Z},\quad i,j=1,\ldots n.

The matrix A=(Aij)A=(A_{ij}) is called the Cartan matrix of 𝔤\mathfrak{g}.

The Dynkin diagram of GG, which we denoted by 𝒟:=𝒟(G)\mathcal{D}:=\mathcal{D}(G), is determined by the Cartan matrix. It consists of a graph whose set of nodes is DD and where the nodes ii and jj are joined by AijAjiA_{ij}A_{ji} edges. When two nodes ii and jj are joined by a double or triple edge, we add to it an arrow pointing to ii if |Aij|>|Aji||A_{ij}|>|A_{ji}|. We call αi\alpha_{i} a short root of 𝒟\mathcal{D} and αj\alpha_{j} a non-short (or long) root of 𝒟\mathcal{D}. (Sometimes, for the sake of narrative convenience, we freely interchange the terminology "node" and "root".) One may prove that there is a one to one correspondence between isomorphism classes of semisimple Lie algebras and Dynkin diagrams of reduced root systems. Moreover, every reduced root system is a disjoint union of mutually orthogonal irreducible root subsystems, each of them corresponding to one of the connected finite Dynkin diagrams AnA_{n}, BnB_{n}, CnC_{n}, DnD_{n} (n>0)(n\in\mathbb{Z}_{>0}), E6E_{6}, E7E_{7}, E8E_{8}, F4F_{4}, G2G_{2}:

An:A_{n}:12n-2n-1nBn:B_{n}:12n-2n-1n
Dn:D_{n}:12n-3n-2n-1nCn:C_{n}:12n-2n-1n
E6:E_{6}:134562F4:F_{4}:1234
E7:E_{7}:1345672G2:G_{2}:12
E8:E_{8}:13456782

The connected components of the Dynkin diagram 𝒟\mathcal{D} determine the simple Lie groups that are factors of the semisimple Lie group GG, each of them corresponding to one of the Dynkin diagrams above.

2.2 Parabolic Subgroup and Subalgebra

A closed subgroup PP of GG is called parabolic if the quotient space G/PG/P is complete, hence projective. A maximal connected solvable subgroup BB of GG is called a Borel subgroup. We fix a Cartan subalgebra 𝔥\mathfrak{h}. Let

𝔟=𝔥αΦ𝔤α\mathfrak{b}=\mathfrak{h}\oplus\bigoplus_{\alpha\in\Phi^{-}}\mathfrak{g}_{\alpha}

be a fixed Borel subalgebra. It is easy to determine the parabolic subalgebras containing 𝔟\mathfrak{b}. They are all of the form

𝔭=𝔟αΦP+𝔤α,\mathfrak{p}=\mathfrak{b}\oplus\bigoplus_{\alpha\in\Phi^{+}_{P}}\mathfrak{g}_{\alpha},

where ΦP+\Phi^{+}_{P} is a subset of Φ+\Phi^{+} that is closed under the addition of roots. Hence the parabolic subalgebras containing 𝔟\mathfrak{b} lie in bijection with the subsets of DD. For some subset II of DD, we write 𝔭I\mathfrak{p}_{I} the parabolic subalgebra corresponding to II. We define PIP_{I} by the parabolic subgroup of GG such that its Lie algebra is 𝔭I\mathfrak{p}_{I}. Therefore the parabolic subgroup PIP_{I} corresponds to the subset II (so a maximal parabolic subgroup is defined by a single root). Then G/PIG/P_{I} has a minimal homogeneous embedding in projective space of the highest weigh module VλV_{\lambda} of GG corresponding to the highest weight λ=iIωi\lambda=\sum_{i\in I}\omega_{i}, where ωi\omega_{i} is the ii-th fundamental weight dual to the roots αiΠ\alpha_{i}\in\Pi, by a very ample line bundle LλL_{\lambda}.

2.3 Rational homogeneous spaces

It is well known that G/PG/P carries a transitive GG-action, it is a smooth projective variety. Borel and Remmert’s classical theorem (4) states that a projective complex manifold which admits a transitive action of its automorphism group is a direct product of an abelian variety and a rational homogeneous space G/PG/P, where GG is a semi-simple algebraic group and PP is a parabolic subgroup.

Every rational homogeneous space G/PG/P can be decomposed into a product

G/PG1/PI1×G2/PI2××Gm/PImG/P\simeq G_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}}

of rational homogeneous spaces with simple algebraic group G1,,GmG_{1},\cdots,G_{m}. Each rational homogeneous space Gi/PIiG_{i}/P_{I_{i}}, called the generalized flag manifold, only depends on the Lie algebra 𝔤i\mathfrak{g}_{i} of GiG_{i}, which is classically determined by the marked Dynkin diagram (16). In the most common notation, we set FI:=G/PIF_{I}:=G/P_{I} by marking on the Dynkin diagram 𝒟\mathcal{D} of GG the nodes corresponding to II. For instance, numbering the nodes of AnA_{n}, the usual flag manifold F(d1,,ds;n+1)F(d_{1},\ldots,d_{s};n+1) corresponds to the marking of I={d1,,ds}I=\{d_{1},\ldots,d_{s}\} (sometimes we omit the braces and just write as Pd1,,dsP_{d_{1},\ldots,d_{s}}).

The two extremal cases correspond to the generalized complete flag manifolds (all nodes marked), and the generalized Grassmannian (only one node marked).

In 16, the authors explicitly describe the lines through a point of a rational homogeneous space G/PIG/P_{I}, where GG is a simple Lie group. Let jIj\in I and N(j)N(j) be the set of nodes in 𝒟\mathcal{D} that are connected to jj.

Definition 2.1.

We call αj(jI)\alpha_{j}(j\in I) an exposed short root if the connected component of jj in D\(I\j)D\backslash(I\backslash j) contains root longer than αj\alpha_{j}, i.e., if an arrow in D\(I\j)D\backslash(I\backslash j) points towards jj.

Remark 2.2.

Obviously, long roots of 𝒟\mathcal{D} in II are not exposed short roots. If II is a set of single point, i.e. X=G/PIX=G/P_{I} is the generalized Grassmannian, then the exposed short root is just the usual short root. It’s worth mentioning that if II contains all long roots of 𝒟\mathcal{D}, then short roots of DD in II are not exposed short roots.

Theorem 2.3.

(16  Theorem 4.3) Let ID={1,,n}I\subseteq D=\{1,\ldots,n\}. Suppose GG to be a simple Lie group. Consider X=G/PIX=G/P_{I} in its minimal homogeneous embedding. Denote by F1(X)F_{1}(X) the space of 1,\mathbb{P}^{1,}s in XX. Then

  1. 1.

    F1(X)=jIF1j(X)F_{1}(X)=\coprod_{j\in I}F_{1}^{j}(X), where F1j(X)F_{1}^{j}(X) is the space of lines of class αˇjH2(X,)\check{\alpha}_{j}\in H_{2}(X,\mathbb{Z}).

  2. 2.

    If αj\alpha_{j} is not an exposed short root, then F1j(X)=G/PI\jN(j)F_{1}^{j}(X)=G/P_{I\backslash j\cup N(j)}.

  3. 3.

    If αj\alpha_{j} is an exposed short root, then F1j(X)F_{1}^{j}(X) is the union of two GG-orbits, an open orbit and its boundary G/PI\jN(j)G/P_{I\backslash j\cup N(j)}.

Remark 2.4.

If I={j}I=\{j\} and αj\alpha_{j} is a long root, then F1(X)F_{1}(X) is just the variety of lines on XX.

Example 2.5.

Let’s consider the Dynkin diagram AnA_{n}, i.e. X=SLn+1/PIX=SL_{n+1}/P_{I} is the generalized flag manifold.

1) For I={k}I=\{k\}, XX is the usual Grassmannian and F1(X)F_{1}(X) is just the variety of lines on XX.

X:X:12k-1kk+1n-2n-1nF1(X):F_{1}(X):12k-1kk+1n-2n-1n

2) For I={d1,d2}I=\{d_{1},d_{2}\}, XX is the usual flag manifold F(d1,d2;n+1)F(d_{1},d_{2};n+1) and F1(X)F_{1}(X) is the disjoint union of F1d1(X)F_{1}^{d_{1}}(X) and F1d2(X)F_{1}^{d_{2}}(X).

X:X:12d11{}_{d_{1}-1}d1{}_{d_{1}}d1+1{}_{d_{1}+1}d21{}_{d_{2}-1}d2{}_{d_{2}}d2+1{}_{d_{2}+1}n-2n-1nF1d1(X):F_{1}^{d_{1}}(X):12d11{}_{d_{1}-1}d1{}_{d_{1}}d1+1{}_{d_{1}+1}d21{}_{d_{2}-1}d2{}_{d_{2}}d2+1{}_{d_{2}+1}n-2n-1n
X:X:12d11{}_{d_{1}-1}d1{}_{d_{1}}d1+1{}_{d_{1}+1}d21{}_{d_{2}-1}d2{}_{d_{2}}d2+1{}_{d_{2}+1}n-2n-1nF1d2(X):F_{1}^{d_{2}}(X):12d11{}_{d_{1}-1}d1{}_{d_{1}}d1+1{}_{d_{1}+1}d21{}_{d_{2}-1}d2{}_{d_{2}}d2+1{}_{d_{2}+1}n-2n-1n

There is a similar statement for CxTxXC_{x}\subseteq\mathbb{P}T_{x}X, the set of tangent directions to lines on XX passing through a fixed point xx. It is a disjoint union of spaces of lines of class αˇ\check{\alpha} through xx.

Theorem 2.6.

(16  Theorem 4.8) Let ID={1,,n}I\subseteq D=\{1,\ldots,n\} and jIj\in I. Suppose GG to be a simple Lie group. Consider X=G/PIX=G/P_{I} in its minimal homogeneous embedding. Let HH be the semisimple part of PIP_{I} and D(H)D(H) be the components of (D\I¯)\j(\overline{D\backslash I})\backslash j containing an element of N(j)N(j), where D\I¯\overline{D\backslash I} means D\ID\backslash I plus any nodes of II attached to a node of D\ID\backslash I. Denote by CxjC_{x}^{j} the space of lines of class αˇj\check{\alpha}_{j} through xx. Then

  1. 1.

    If αj\alpha_{j} is not an exposed short root, then Cxj=H/PN(j)C_{x}^{j}=H/P_{N(j)}.

  2. 2.

    If αj\alpha_{j} is an exposed short root, then CxjC_{x}^{j} is a union of an open PIP_{I}-orbit and its boundary H/PN(j)H/P_{N(j)}.

Remark 2.7.

If I={j}I=\{j\}, then the set of nodes of the Dynkin diagram HH is D(H)=D\jD(H)=D\backslash{j}. PN(j)P_{N(j)} is a parabolic subgroup of HH by marking in D(H)D(H) the nodes in DD that are connected to jj and Cxj=H/PN(j)C_{x}^{j}=H/P_{N(j)}. Moreover, if αj\alpha_{j} is a long root then CxjC_{x}^{j} is just the variety of lines through fixed points, i.e. so-called VMRTs. We refer to 14 for a complete account on VMRTs.

Example 2.8.

Let’s consider the Dynkin diagram AnA_{n}, i.e. X=SLn+1/PIX=SL_{n+1}/P_{I} is the generalized flag manifold.

1) For I={k}I=\{k\}, XX is the usual Grassmannian and Cx=k1×nkC_{x}=\mathbb{P}^{k-1}\times\mathbb{P}^{n-k} is just the variety of lines through xx.

X:X:12k-1kk+1n-2n-1n×\timesCx:C_{x}:12k-1k+1n-2n-1n

2) For I={d1,d2}I=\{d_{1},d_{2}\}, XX is the usual flag manifold F(d1,d2;n+1)F(d_{1},d_{2};n+1) and CxC_{x} is the disjoint union of Cxd1C_{x}^{d_{1}} and Cxd2C_{x}^{d_{2}}.

Cxd1=d11×d2d11,Cxd2=d2d11×nd2.C_{x}^{d_{1}}=\mathbb{P}^{d_{1}-1}\times\mathbb{P}^{d_{2}-d_{1}-1},~{}~{}C_{x}^{d_{2}}=\mathbb{P}^{d_{2}-d_{1}-1}\times\mathbb{P}^{n-d_{2}}.
X:X:12d11{}_{d_{1}-1}d1{}_{d_{1}}d1+1{}_{d_{1}+1}d21{}_{d_{2}-1}d2{}_{d_{2}}d2+1{}_{d_{2}+1}n-2n-1n×\timesCxd1:C_{x}^{d_{1}}:12d11{}_{d_{1}-1}d1+1{}_{d_{1}+1}d22{}_{d_{2}-2}d21{}_{d_{2}-1}
X:X:12d11{}_{d_{1}-1}d1{}_{d_{1}}d1+1{}_{d_{1}+1}d21{}_{d_{2}-1}d2{}_{d_{2}}d2+1{}_{d_{2}+1}n-2n-1n×\timesCxd2:C_{x}^{d_{2}}:d1+1{}_{d_{1}+1}d1+2{}_{d_{1}+2}d21{}_{d_{2}-1}d2+1{}_{d_{2}+1}n-1n

Not only 1\mathbb{P}^{1} but also all linear spaces can be read from the marked Dynkin diagrams.

Theorem 2.9.

(16  Theorem 4.9, 4.14) Let GG be a simple group and X=G/PSX=G/P_{S} is a rational homogeneous space. Let Fkα(X)F_{k}^{\alpha}(X) denote the variety parameterizing the α\alpha-class k\mathbb{P}^{k}’s on XX.

  • If αS\alpha\in S is not an exposed short root, then for all kk, Fkα(X)F_{k}^{\alpha}(X) is the disjoint union of homogeneous spaces G/PβjG/P_{\sum\beta_{j}}, where {βj}\{\beta_{j}\} is a set of positive roots such that the component of 𝒟\{βj}\mathcal{D}\backslash\{\beta_{j}\} containing α\alpha is isomorphic to 𝒟(Ak)\mathcal{D}(A_{k}), intersects SS only in α\alpha, and α\alpha is an extremal node of this component.

  • If αS\alpha\in S is an exposed short root, then for all kk, Fkα(X)F_{k}^{\alpha}(X) consists of a finite number of GG-orbits.

3 Uniform vector bundles

Given a smooth projective variety XX and a vector bundle EE on XX, we denote \mathcal{M} to be an unsplit family of rational curves on XX. \mathcal{M} is called unsplit if \mathcal{M} is a proper \mathbb{C}-scheme. We say that EE is uniform with respect to \mathcal{M} if the restriction of EE to the normalization of every curve in \mathcal{M} splits as a direct sum of line bundles with the same splitting type. If XX is a generalized Grassmannian G/PkG/P_{k}, then we just call EE uniform without mention the unspilt family \mathcal{M}.

3.1 Uniform vector bundles on generalized Grassmannians

Along this section we will work on uniform vector bundles on rational homogeneous spaces of Picard number one, i.e. generalized Grassmannians. Let GG be a simple Lie group and D={1,2,,n}D=\{1,2,\ldots,n\} be the set of nodes of the Dynkin diagram 𝒟\mathcal{D} of GG. Denote by PkP_{k} the parabolic subgroup of GG corresponds to the node kk. Consider the generalized Grassmannian 𝒢=G/Pk\mathcal{G}=G/P_{k} or, for brevity, 𝒟/Pk\mathcal{D}/P_{k}. Denote by :=G/PN(k)\mathcal{M}:=G/P_{N(k)} the generalized flag manifold defined by the marked Dynkin diagram (𝒟,N(k))(\mathcal{D},N(k)) and by 𝒰:=G/Pk,N(k)=G/(PN(k)Pk)\mathcal{U}:=G/P_{k,N(k)}=G/(P_{N(k)}\cap P_{k}) the universal family, which has a natural 1\mathbb{P}^{1}-bundle structure over \mathcal{M}, i.e. we have the natural diagram

(3.5)

Remarkably, \mathcal{M} defined above is indeed an unsplit family of rational curves on 𝒢\mathcal{G}. Given x𝒢x\in\mathcal{G}, x=q(p1(x))\mathcal{M}_{x}=q(p^{-1}(x)), which we call the special family of lines of class αkˇ\check{\alpha_{k}} through xx, coincides with H/PN(k)H/P_{N(k)} by Remark 2.7, where the set of nodes of the Dynkin diagram HH is D(H)=D\kD(H)=D\backslash{k}.

When kk is an extremal node, that is, the subdiagram 𝒟(H)\mathcal{D}(H) is connected. Remarkably, in the case 𝒟=Dn\mathcal{D}=D_{n}, i.e. G=SO(2n)G=SO(2n), since G/Pn1G/PnG/P_{n-1}\cong G/P_{n}, we only need to think about the extremal node nn. Similarly, since E6/P1E6/P6E_{6}/P_{1}\cong E_{6}/P_{6}, we just consider the the extremal node 11 in E6E_{6}. According to Theorem 2.6, x\mathcal{M}_{x} has the following possibilities:

  • Projective spaces or smooth quadrics,

  • Grassmannians,

  • Spinor varities,

  • E6/P6E_{6}/P_{6}, E7/P7E_{7}/P_{7}, C3/P3C_{3}/P_{3}.

The possibilities are list in Table 1 below.

Table 1: x\mathcal{M}_{x} corresponding to an extremal node
nodenode 𝒟\mathcal{D} x\mathcal{M}_{x} AnA_{n} BnB_{n} CnC_{n} DnD_{n} En(n=6,7,8)E_{n}(n=6,7,8) Fn(n=4)F_{n}(n=4) Gn(n=2)G_{n}(n=2)
11 n1\mathbb{P}_{n-1} Q2n3Q_{2n-3} 2n3\mathbb{P}_{2n-3} Q2n4Q_{2n-4} 𝒮n2\mathcal{S}_{n-2} C3/P3C_{3}/P_{3} 1\mathbb{P}_{1}
nn n1\mathbb{P}_{n-1} n1\mathbb{P}_{n-1} n1\mathbb{P}_{n-1} G(2,n)G(2,n) En1/Pn1(n6)E_{n-1}/P_{n-1}(n\neq 6) 𝒮3\mathcal{S}_{3} 1\mathbb{P}_{1}
22 G(3,n)G(3,n)

We observe that for x𝒢x\in\mathcal{G}, the morphism from x\mathcal{M}_{x} to Grassmannian plays a critical role in determining whether a uniform vector bundle can split as a direct sum of line bundles. Let ς\varsigma be a positive integer smaller than or equal to dimx\text{dim}~{}\mathcal{M}_{x}. As long as we show that the morphism xG(t,ς)\mathcal{M}_{x}\rightarrow G(t,\varsigma) can only be constant for any integer 1t[ς2]1\leq t\leq[\frac{\varsigma}{2}] and every x𝒢x\in\mathcal{G}, then every uniform rr-bundle on 𝒢\mathcal{G} splits for rςr\leq\varsigma. We suggest that interested readers refer to Theorem 3.1 in paper 18 for details. Now, let’s analyze the morphism xG(t,ς)\mathcal{M}_{x}\rightarrow G(t,\varsigma) one by one according to the probabilities of x\mathcal{M}_{x}.

Case I. When x\mathcal{M}_{x} is a projective space N\mathbb{P}^{N} or a smooth quadric QN(N=2m+1)Q^{N}~{}(N=2m+1), then their Chow rings have the form

[]/(N+1),\mathbb{Z}[\mathcal{H}]/(\mathcal{H}^{N+1}),

where \mathcal{H} is a hyperplane section. In particular, dimH2t(x,)=1\text{dim}H^{2t}(\mathcal{M}_{x},\mathbb{C})=1 for every t[N2]t\leq[\frac{N}{2}]. By the proof of Lemma 3.4 in paper 18, the only morphisms xG(t,N)\mathcal{M}_{x}\rightarrow G(t,N) are constant for any integer 1t[N2]1\leq t\leq[\frac{N}{2}].

When x\mathcal{M}_{x} is a smooth quadric QN(N=2m)Q^{N}~{}(N=2m), since

A(Q2m)=[,𝒰]/(2m+1,2𝒰m+1,m𝒰𝒰2),A(Q^{2m})=\mathbb{Z}[\mathcal{H},\mathcal{U}]/(\mathcal{H}^{2m+1},2\mathcal{H}\mathcal{U}-\mathcal{H}^{m+1},\mathcal{H}^{m}\mathcal{U}-\mathcal{U}^{2}),

where \mathcal{H} is a hyperplane section and 𝒰\mathcal{U} is a subvariety of codimension mm, then we get that xG(t,N1)\mathcal{M}_{x}\rightarrow G(t,N-1) can only be constant map similarly.

Case II. x\mathcal{M}_{x} is Grassmannian G(d,n)(2dnd)G(d,n)~{}(2\leq d\leq n-d). We claim that the only morphisms G(d,n)G(t,nd+1)G(d,n)\rightarrow G(t,n-d+1) are constant for any integer 1t[nd+12]1\leq t\leq[\frac{n-d+1}{2}].

Lemma 3.1.

There are no nonconstant maps from G(d,n)(2dnd)G(d,n)~{}(2\leq d\leq n-d) to G(t,nd+1)G(t,n-d+1) for any integer 1t[nd+12]1\leq t\leq[\frac{n-d+1}{2}].

Proof.

Assume that we have a nonconstant morphism ϕ:=G(d,n)G(t,nd+1)\phi:=G(d,n)\rightarrow G(t,n-d+1). Then there exists a maximal linear subspace nd\mathbb{P}_{n-d} such that ϕ\phi restricts to it is also nonconstant. Denote by ψ\psi to be the restriction map. Let’s consider ψHt,ψQnd+1t\psi^{\ast}H_{t},\psi^{\ast}Q_{n-d+1-t} and ϕHt,ϕQnd+1t\phi^{\ast}H_{t},\phi^{\ast}Q_{n-d+1-t}(the pull back of universal bundle HtH_{t} and universal quotient bundle Qnd+1tQ_{n-d+1-t} under ψ,ϕ\psi,\phi). Denote by c1,,ctc_{1},\ldots,c_{t} and d1,,dnd+1td_{1},\ldots,d_{n-d+1-t} the Chern classes of ψHt\psi^{\ast}H_{t} and ψQnd+1t\psi^{\ast}Q_{n-d+1-t}, respectively, and by C1,,CtC_{1},\ldots,C_{t} and D1,,Dnd+1tD_{1},\ldots,D_{n-d+1-t} the Chern classes of ϕHt\phi^{\ast}H_{t} and ϕQnd+1t\phi^{\ast}Q_{n-d+1-t}, respectively.

On G(d,n)(2dnd)G(d,n)~{}(2\leq d\leq n-d), we have an exact sequence

0ϕHt𝒪G(d,n)nd+1ϕQnd+1t00\rightarrow\phi^{\ast}H_{t}\rightarrow\mathcal{O}^{\oplus n-d+1}_{G(d,n)}\rightarrow\phi^{\ast}Q_{n-d+1-t}\rightarrow 0

which is the pull back of the universal exact sequence

0Ht𝒪G(t,nd+1t)nd+1Qnd+1t00\rightarrow H_{t}\rightarrow\mathcal{O}^{\oplus n-d+1}_{G(t,n-d+1-t)}\rightarrow Q_{n-d+1-t}\rightarrow 0

on G(t,nd+1)G(t,n-d+1). Then

c(ϕHt)c(ϕQnd+1t)=1,c(\phi^{\ast}H_{t})\cdot c(\phi^{\ast}Q_{n-d+1-t})=1,

i.e.

(1+C1++Ct)(1+D1++Dnd+1t)=1.(1+C_{1}+\cdots+C_{t})\cdot(1+D_{1}+\cdots+D_{n-d+1-t})=1.

Since nd+1<d(nd)=dimG(d,n)n-d+1<d(n-d)=\text{dim}G(d,n), obviously we have CtDnd+1t=0C_{t}\cdot D_{n-d+1-t}=0.

On the other hand, on ndG(d,n)\mathbb{P}_{n-d}\subset G(d,n), we also have an exact sequence

0ψHt𝒪ndnd+1ψQnd+1t0.0\rightarrow\psi^{\ast}H_{t}\rightarrow\mathcal{O}^{\oplus n-d+1}_{\mathbb{P}_{n-d}}\rightarrow\psi^{\ast}Q_{n-d+1-t}\rightarrow 0.

Then

c(ψHt)c(ψQnd+1t)=1,c(\psi^{\ast}H_{t})\cdot c(\psi^{\ast}Q_{n-d+1-t})=1,

i.e.

(1+c1++ct)(1+d1++dnd+1t)=1.(1+c_{1}+\cdots+c_{t})\cdot(1+d_{1}+\cdots+d_{n-d+1-t})=1.

Combining the above equation and the ψ\psi nonconstant assumption, i.e.

c1=degψ𝒪G(t,nd+1)(1)0,c_{1}=\text{deg}\psi^{\ast}\mathcal{O}_{G(t,n-d+1)}(1)\neq 0,

we get ct0c_{t}\neq 0 and dnd+1t0d_{n-d+1-t}\neq 0.

In order to show that this contradicts that CtDnd+1t=0C_{t}\cdot D_{n-d+1-t}=0, we need to use some Schubert classes in G(d,n)G(d,n) (see 8 Chapter 4). Let’s review some basic facts and fix some notation first. Choose a complete flag 𝒱\mathcal{V} in n\mathbb{C}^{n}, that is, a nested sequence of subspaces

0V1VndVnd=n0\subset V_{1}\subset\cdots\subset V_{n-d}\subset V_{n-d}=\mathbb{C}^{n}

with dim Vi=iV_{i}=i. For a sequence a=(a1,,ad)a=(a_{1},\ldots,a_{d}) with nda1ad0n-d\geq a_{1}\geq\ldots\geq a_{d}\geq 0, we define the Schubert cycle Σa(𝒱)G(d,n)\Sigma_{a}(\mathcal{V})\subset G(d,n) to be the closed subset

Σa(𝒱)={ΛG(d,n)|dim(Vnd+iaiΛi)for alli}.\Sigma_{a}(\mathcal{V})=\{\Lambda\in G(d,n)|\text{dim}(V_{n-d+i-a_{i}}\cap\Lambda\geq i)~{}\text{for all}~{}i\}.

The class σa:=[Σa(𝒱)]\sigma_{a}:=[\Sigma_{a}(\mathcal{V})] ia called Schubert class. By Theorem 4.1 in 8, σa\sigma_{a} is of codimension a1++ada_{1}+\cdots+a_{d}. To simplify notation, we generally suppress trailing zeros in the indices and write σa1,,as\sigma_{a_{1},\ldots,a_{s}} in place of σa1,,as,0,,0\sigma_{a_{1},\ldots,a_{s},0,\ldots,0}. With this notation, nd\mathbb{P}_{n-d} can be represented as σnd,,nd\sigma_{n-d,\ldots,n-d} and we may write

(1)tCt=a1++ad=txa1,,adσa1,,adandDnd+1t=a1++ad=nd+1tya1,,adσa1,,ad,(-1)^{t}C_{t}=\sum_{a_{1}+\cdots+a_{d}=t}x_{a_{1},\ldots,a_{d}}\sigma_{a_{1},\ldots,a_{d}}~{}\text{and}~{}D_{n-d+1-t}=\sum_{a_{1}+\cdots+a_{d}=n-d+1-t}y_{a_{1},\ldots,a_{d}}\sigma_{a_{1},\ldots,a_{d}},

where the xa1,,ad{x_{a_{1},\ldots,a_{d}}}^{\prime}s and ya1,,ad{y_{a_{1},\ldots,a_{d}}}^{\prime}s are non negative integers by the nefness of HtH_{t}^{\vee} (the dual of HtH_{t}) and Qnd+1tQ_{n-d+1-t}.

Pieri’s formula (see Proposition 4.9 in 8) tells us that for any integers b1,,bdb_{1},\ldots,b_{d} with b1++bd=lndb_{1}+\cdots+b_{d}=l\leq n-d,

σb1σbdnd=σb1σbdσnd,,nd,0=σnd,,nd,l.\sigma_{b_{1}}\cdots\sigma_{b_{d}}\cap\mathbb{P}_{n-d}=\sigma_{b_{1}}\cdots\sigma_{b_{d}}\cdot\sigma_{n-d,\ldots,n-d,0}=\sigma_{n-d,\ldots,n-d,l}.

Using Giambelli’s formula (see Proposition 4.16 in 8) for σa1,,ad\sigma_{a_{1},\ldots,a_{d}}’s, expanding the determinants and then intersecting with σnd,,nd,0\sigma_{n-d,\ldots,n-d,0}, we can immediately get the following identities:

ct\displaystyle c_{t} =Ctnd\displaystyle=C_{t}\cap\mathbb{P}_{n-d}
=(1)txt,0,,0σtσnd,,nd,0\displaystyle=(-1)^{t}x_{t,0,\ldots,0}\sigma_{t}\cdot\sigma_{n-d,\ldots,n-d,0}
=(1)txt,0,,0σnd,,nd,t;\displaystyle=(-1)^{t}x_{t,0,\ldots,0}\sigma_{n-d,\ldots,n-d,t};
dnd+1t\displaystyle d_{n-d+1-t} =Dnd+1tnd\displaystyle=D_{n-d+1-t}\cap\mathbb{P}_{n-d}
=ynd+1t,0,,0σnd+1tσnd,,nd,0\displaystyle=y_{n-d+1-t,0,\ldots,0}\sigma_{n-d+1-t}\cdot\sigma_{n-d,\ldots,n-d,0}
=ynd+1t,0,,0σnd,,nd,nd+1t.\displaystyle=y_{n-d+1-t,0,\ldots,0}\sigma_{n-d,\ldots,n-d,n-d+1-t}.

Hence xt,0,,0ynd+1t,0,,00x_{t,0,\ldots,0}y_{n-d+1-t,0,\ldots,0}\neq 0. But

0=CtDnd+1t=xt,0,,0ynd+1t,0,,0σtσnd+1t+,0=C_{t}\cdot D_{n-d+1-t}=x_{t,0,\ldots,0}y_{n-d+1-t,0,\ldots,0}\sigma_{t}\cdot\sigma_{n-d+1-t}+\ldots,

where the summation is linear combination of Schubert cycles with non-negative coefficients by the Littlewood-Richardson formula. Therefore

xt,0,,0ynd+1t,0,,0=0,x_{t,0,\ldots,0}y_{n-d+1-t,0,\ldots,0}=0,

a contradiction. ∎

When 𝒢=Dn/Pn\mathcal{G}=D_{n}/P_{n}, Mx=G(2,n)M_{x}=G(2,n). By the above Lemma, we obtain the only morphisms G(2,n)G(t,n1)G(2,n)\rightarrow G(t,n-1) are constant for any integer 1t[n12]1\leq t\leq[\frac{n-1}{2}].

When 𝒢=En/P2(n=6,7,8)\mathcal{G}=E_{n}/P_{2}~{}(n=6,7,8), Mx=G(3,n)(n=6,7,8)M_{x}=G(3,n)~{}(n=6,7,8). By the above Lemma, we obtain the only morphisms G(3,n)G(t,n2)(n=6,7,8)G(3,n)\rightarrow G(t,n-2)~{}(n=6,7,8) are constant for any integer 1t[n22]1\leq t\leq[\frac{n-2}{2}]. Remarkably, in these cases, the value of ς\varsigma can be appropriately enlarged. Since G(3,n)(n=6,7,8)G(3,n)~{}(n=6,7,8) is 3(n3)3(n-3) dimensional, we can easily know

dim(G(3,6))=9>dimG(t,5),\text{dim}(G(3,6))=9>\text{dim}G(t,5),
dim(G(3,7))=12>dimG(t,6)\text{dim}(G(3,7))=12>\text{dim}G(t,6)

and

dim(G(3,8))=15>dimG(t,7)for allt1.\text{dim}(G(3,8))=15>\text{dim}G(t,7)~{}\text{for all}~{}t\geq 1.

Since the Picard number of G(3,n)(n=6,7,8)G(3,n)~{}(n=6,7,8) is one, the only morphisms

G(3,6)G(t,5),G(3,6)\rightarrow G(t,5),
G(3,7)G(t,6)G(3,7)\rightarrow G(t,6)

and

G(3,8)G(t,7)G(3,8)\rightarrow G(t,7)

are all constant for any integer t1t\geq 1.

Case III. x\mathcal{M}_{x} is spinor variety 𝒮n(n=3,4,5,6)\mathcal{S}_{n}~{}(n=3,4,5,6). The Chow ring of 𝒮n\mathcal{S}_{n} is presented as a quotient of [X1,,Xn]\mathbb{Z}[X_{1},\ldots,X_{n}] module the relations

Xs2+2i=1s1(1)iXs+iXsi+(1)sX2s=0X_{s}^{2}+2\sum_{i=1}^{s-1}(-1)^{i}X_{s+i}X_{s-i}+(-1)^{s}X_{2s}=0

for 1sn1\leq s\leq n, where Xj,sX_{j}^{,}s are the Schubert classes of codimension jj, X0=1X_{0}=1 and Xj=0X_{j}=0 for j<0j<0 or j>nj>n (see Section 3.2 in 26). In particular, dimH2t(x,)=1\text{dim}H^{2t}(\mathcal{M}_{x},\mathbb{C})=1 for every t[52]t\leq[\frac{5}{2}]. Hence, the only morphisms xG(t,5)\mathcal{M}_{x}\rightarrow G(t,5) are constant for any integer 1t[52]1\leq t\leq[\frac{5}{2}]. Remarkably, for n=4,5,6n=4,5,6, the value of ς\varsigma can be appropriately enlarged. Due to the dimension of 𝒮n\mathcal{S}_{n} is n(n+1)2\frac{n(n+1)}{2}, one can check that

dim(𝒮4)=10>dimG(t,6),\text{dim}(\mathcal{S}_{4})=10>\text{dim}G(t,6),
dim(𝒮5)=15>dimG(t,7)\text{dim}(\mathcal{S}_{5})=15>\text{dim}G(t,7)

and

dim(𝒮6)=21>dimG(t,9)for allt1.\text{dim}(\mathcal{S}_{6})=21>\text{dim}G(t,9)~{}\text{for all}~{}t\geq 1.

Since the Picard number of 𝒮n\mathcal{S}_{n} is one, the only morphisms

𝒮4G(t,6),\mathcal{S}_{4}\rightarrow G(t,6),
𝒮5G(t,7)\mathcal{S}_{5}\rightarrow G(t,7)

and

𝒮6G(t,9)\mathcal{S}_{6}\rightarrow G(t,9)

are all constant for any integer t1t\geq 1.

Case IV. x=E6/P6\mathcal{M}_{x}=E_{6}/P_{6}. The Chow ring of E6/P6E_{6}/P_{6} have the following form (see 7 Theorem 5). Let y1,y4y_{1},y_{4} be the Schubert classes on E6/P6E_{6}/P_{6}. Then

A(E6/P6)=[y1,y4]/(r9,r12),A(E_{6}/P_{6})=\mathbb{Z}[y_{1},y_{4}]/(r_{9},r_{12}),

where

r9=2y19+3y1y426y15y4;r_{9}=2y_{1}^{9}+3y_{1}y_{4}^{2}-6y_{1}^{5}y_{4};
r12=y436y14y42+y112.r_{12}=y_{4}^{3}-6y_{1}^{4}y_{4}^{2}+y_{1}^{12}.
Lemma 3.2.

There are no nonconstant maps from E6/P6E_{6}/P_{6} to G(t,10)G(t,10) for any integer 1t51\leq t\leq 5.

Proof.

(i). 1t31\leq t\leq 3. Since dimH2t(x,)=1\text{dim}H^{2t}(\mathcal{M}_{x},\mathbb{C})=1, by the proof of Lemma 3.4 in paper 18, the only morphisms E6/P6G(t,10)E_{6}/P_{6}\rightarrow G(t,10) are constant.

(ii). t=4t=4. Let ϕ\phi be a morphism from E6/P6E_{6}/P_{6} to G(4,10)G(4,10). On E6/P6E_{6}/P_{6}, we have an exact sequence

0ϕH4𝒪G(d,n)10ϕQ60.0\rightarrow\phi^{\ast}H_{4}\rightarrow\mathcal{O}^{\oplus 10}_{G(d,n)}\rightarrow\phi^{\ast}Q_{6}\rightarrow 0.

Then c(ϕH4)c(ϕQ6)=1c(\phi^{\ast}H_{4})\cdot c(\phi^{\ast}Q_{6})=1. According to the Chow ring of E6/P6E_{6}/P_{6}, we can expand the equation into the following form:

(1+a1y1+a2y12+a3y13+\displaystyle(1+a_{1}y_{1}+a_{2}y_{1}^{2}+a_{3}y_{1}^{3}+ a4y14+a4~y4)(1+b1y1+b2y12+\displaystyle a_{4}y_{1}^{4}+\widetilde{a_{4}}y_{4})\cdot(1+b_{1}y_{1}+b_{2}y_{1}^{2}+
b3y13+b4y14+b4~y4+b5y15+b5~y1y4+b6y16+b6~y12y4)=1.\displaystyle b_{3}y_{1}^{3}+b_{4}y_{1}^{4}+\widetilde{b_{4}}y_{4}+b_{5}y_{1}^{5}+\widetilde{b_{5}}y_{1}y_{4}+b_{6}y_{1}^{6}+\widetilde{b_{6}}y_{1}^{2}y_{4})=1.

Since A4(E6/P6)A^{4}(E_{6}/P_{6}) is freely generated by the classes y14,y4y_{1}^{4},y_{4}, the above equation implies that the coefficient of y4y_{4} is 0, i.e. a4~+b4~=0\widetilde{a_{4}}+\widetilde{b_{4}}=0. On the other hand, A8(E6/P6)A^{8}(E_{6}/P_{6}) is freely generated by the classes y18,y14y4,y42y_{1}^{8},y_{1}^{4}y_{4},y_{4}^{2}, the above equation implies that the coefficient of y42y_{4}^{2} is also zero, i.e. a4~b4~=0\widetilde{a_{4}}\cdot\widetilde{b_{4}}=0. Hence a4~=b4~=0\widetilde{a_{4}}=\widetilde{b_{4}}=0. Therefore, this case can boil down to case (i).

(iii). t=5t=5. By iterating the previous process, we get c(ϕH5)c(ϕQ5)=1c(\phi^{\ast}H_{5})\cdot c(\phi^{\ast}Q_{5})=1, i.e.

(1+a1y1+a2y12+a3y13\displaystyle(1+a_{1}y_{1}+a_{2}y_{1}^{2}+a_{3}y_{1}^{3} +a4y14+a4~y4+a5y15+a5~y1y4)\displaystyle+a_{4}y_{1}^{4}+\widetilde{a_{4}}y_{4}+a_{5}y_{1}^{5}+\widetilde{a_{5}}y_{1}y_{4})
\displaystyle\cdot (1+b1y1+b2y12+b3y13+b4y14+b4~y4+b5y15+b5~y1y4)=1.\displaystyle(1+b_{1}y_{1}+b_{2}y_{1}^{2}+b_{3}y_{1}^{3}+b_{4}y_{1}^{4}+\widetilde{b_{4}}y_{4}+b_{5}y_{1}^{5}+\widetilde{b_{5}}y_{1}y_{4})=1.

In a similar way, we can prove a4~=b4~=0\widetilde{a_{4}}=\widetilde{b_{4}}=0. Next, let’s consider the vanishing of a5~\widetilde{a_{5}}. Since A5(E6/P6)A^{5}(E_{6}/P_{6}) is freely generated by the classes y15,y1y4y_{1}^{5},y_{1}y_{4}, the above equation implies that the coefficient of y1y4y_{1}y_{4} is 0, i.e. a5~+b5~=0\widetilde{a_{5}}+\widetilde{b_{5}}=0. On the other hand, A6(E6/P6)A^{6}(E_{6}/P_{6}) is freely generated by the classes y16,y12y4y_{1}^{6},y_{1}^{2}y_{4}, the above equation implies that the coefficient of y12y4y_{1}^{2}y_{4} is also zero, i.e. a1b5~+b1a5~=0a_{1}\widetilde{b_{5}}+b_{1}\widetilde{a_{5}}=0. Combining these equations with a1=b1a_{1}=-b_{1}, we obtain that a5~=b5~=0\widetilde{a_{5}}=\widetilde{b_{5}}=0. Therefore, this case can boil down to case (i). ∎

Case V. x=E7/P7\mathcal{M}_{x}=E_{7}/P_{7}. The Chow ring of E7/P7E_{7}/P_{7} have the following form (see 7 Theorem 6). Let y1,y5,y9y_{1},y_{5},y_{9} be the Schubert classes on E7/P7E_{7}/P_{7}. Then

A(E7/P7)=[y1,y5,y9]/(r10,r14,r18),A(E_{7}/P_{7})=\mathbb{Z}[y_{1},y_{5},y_{9}]/(r_{10},r_{14},r_{18}),

where

r10=y522y1y9;r_{10}=y_{5}^{2}-2y_{1}y_{9};
r14=2y5y99y14y52+6y19y5y114;r_{14}=2y_{5}y_{9}-9y_{1}^{4}y_{5}^{2}+6y_{1}^{9}y_{5}-y_{1}^{14};
r18=y92+10y13y539y18y52+2y113y5.r_{1}8=y_{9}^{2}+10y_{1}^{3}y_{5}^{3}-9y_{1}^{8}y_{5}^{2}+2y_{1}^{13}y_{5}.
Lemma 3.3.

There are no nonconstant maps from E7/P7E_{7}/P_{7} to G(t,13)G(t,13) for any integer 1t61\leq t\leq 6.

Proof.

(i). 1t41\leq t\leq 4. Since dimH2t(x,)=1\text{dim}H^{2t}(\mathcal{M}_{x},\mathbb{C})=1, by the proof of Lemma 3.4 in paper 18, the only morphisms E7/P7G(t,13)E_{7}/P_{7}\rightarrow G(t,13) are constant.

(ii). t=5t=5. Let ϕ\phi be a morphism from E7/P7E_{7}/P_{7} to G(5,13)G(5,13). On E7/P7E_{7}/P_{7}, we have an exact sequence

0ϕH5𝒪G(d,n)13ϕQ80.0\rightarrow\phi^{\ast}H_{5}\rightarrow\mathcal{O}^{\oplus 13}_{G(d,n)}\rightarrow\phi^{\ast}Q_{8}\rightarrow 0.

Then c(ϕH5)c(ϕQ8)=1c(\phi^{\ast}H_{5})\cdot c(\phi^{\ast}Q_{8})=1. According to the Chow ring of E7/P7E_{7}/P_{7}, we can expand the equation into the following form:

(1+a1y1+\displaystyle(1+a_{1}y_{1}+ a2y12+a3y13+a4y14+a5y15+a5~y5)(1+b1y1+b2y12+b3y13\displaystyle a_{2}y_{1}^{2}+a_{3}y_{1}^{3}+a_{4}y_{1}^{4}+a_{5}y_{1}^{5}+\widetilde{a_{5}}y_{5})\cdot(1+b_{1}y_{1}+b_{2}y_{1}^{2}+b_{3}y_{1}^{3}
+b4y14+b5y15+b5~y5+b6y16+b6~y1y5+b7y17+b7~y12y5+b8y18+b8~y13y5)=1.\displaystyle+b_{4}y_{1}^{4}+b_{5}y_{1}^{5}+\widetilde{b_{5}}y_{5}+b_{6}y_{1}^{6}+\widetilde{b_{6}}y_{1}y_{5}+b_{7}y_{1}^{7}+\widetilde{b_{7}}y_{1}^{2}y_{5}+b_{8}y_{1}^{8}+\widetilde{b_{8}}y_{1}^{3}y_{5})=1.

Since A5(E7/P7)A^{5}(E_{7}/P_{7}) is freely generated by the classes y15,y5y_{1}^{5},y_{5}, the above equation implies that the coefficient of y5y_{5} is 0, i.e. a5~+b5~=0\widetilde{a_{5}}+\widetilde{b_{5}}=0. On the other hand, A10(E7/P7)A^{10}(E_{7}/P_{7}) is freely generated by the classes y110,y15y5,y52y_{1}^{10},y_{1}^{5}y_{5},y_{5}^{2}, the above equation implies that the coefficient of y52y_{5}^{2} is also zero, i.e. a5~b5~=0\widetilde{a_{5}}\widetilde{b_{5}}=0. Hence a5~=b5~=0\widetilde{a_{5}}=\widetilde{b_{5}}=0. Therefore, this case can boil down to case (i).

(iii). t=6t=6. By iterating the previous process, we get c(ϕH6)c(ϕQ7)=1c(\phi^{\ast}H_{6})\cdot c(\phi^{\ast}Q_{7})=1, i.e.

(1+a1y1+a2y12\displaystyle(1+a_{1}y_{1}+a_{2}y_{1}^{2} +a3y13+a4y14+a5y15+a5~y5+a6y16+a6~y1y5)(1+b1y1+\displaystyle+a_{3}y_{1}^{3}+a_{4}y_{1}^{4}+a_{5}y_{1}^{5}+\widetilde{a_{5}}y_{5}+a_{6}y_{1}^{6}+\widetilde{a_{6}}y_{1}y_{5})\cdot(1+b_{1}y_{1}+
b2y12+b3y13+b4y14+b5y15+b5~y5+b6y16+b6~y1y5+b7y17+b7~y12y5)=1.\displaystyle b_{2}y_{1}^{2}+b_{3}y_{1}^{3}+b_{4}y_{1}^{4}+b_{5}y_{1}^{5}+\widetilde{b_{5}}y_{5}+b_{6}y_{1}^{6}+\widetilde{b_{6}}y_{1}y_{5}+b_{7}y_{1}^{7}+\widetilde{b_{7}}y_{1}^{2}y_{5})=1.

In a similar way, we can prove a5~=b5~=0\widetilde{a_{5}}=\widetilde{b_{5}}=0. Next, let’s consider the vanishing of a6~\widetilde{a_{6}}. Since A6(E7/P7)A^{6}(E_{7}/P_{7}) is freely generated by the classes y16,y1y5y_{1}^{6},y_{1}y_{5}, the above equation implies that the coefficient of y1y5y_{1}y_{5} is 0, i.e. a6~+b6~=0\widetilde{a_{6}}+\widetilde{b_{6}}=0. On the other hand, A12(E7/P7)A^{12}(E_{7}/P_{7}) is freely generated by the classes y112,y12y52,y17y5y_{1}^{12},y_{1}^{2}y_{5}^{2},y_{1}^{7}y_{5}, the above equation implies that the coefficient of y12y52y_{1}^{2}y_{5}^{2} is also zero, i.e. a6~b6~=0\widetilde{a_{6}}\widetilde{b_{6}}=0. Hence, a6~=b6~=0\widetilde{a_{6}}=\widetilde{b_{6}}=0. Therefore, this case can boil down to case (i). ∎

Case VI. x=C3/P3\mathcal{M}_{x}=C_{3}/P_{3}. The Chow ring of C3/P3C_{3}/P_{3} is

A(C3/P3)=[y1,y3]/(y148y1y3,x32),A(C_{3}/P_{3})=\mathbb{Z}[y_{1},y_{3}]/(y_{1}^{4}-8y_{1}y_{3},x_{3}^{2}),

where y1,y3y_{1},y_{3} are the Schubert classes on C3/P3C_{3}/P_{3} (see Section 3.1 in 26). Since dimH2t(C3/P3,)=1\text{dim}H^{2t}(C_{3}/P_{3},\mathbb{C})=1 for 1t21\leq t\leq 2, the only morphisms C3/P3G(t,5)C_{3}/P_{3}\rightarrow G(t,5) are constant for any integer 1t21\leq t\leq 2.

Summing up, we have obtained the following Table 2.

Table 2: ς(𝒢)\varsigma(\mathcal{G})
𝒢\mathcal{G}
An/P1A_{n}/P_{1}
An/PnA_{n}/P_{n}
Bn/P1B_{n}/P_{1}
Bn/PnB_{n}/P_{n}
Cn/P1C_{n}/P_{1}
Cn/PnC_{n}/P_{n}
Dn/P1D_{n}/P_{1}
Dn/Pn1D_{n}/P_{n-1}
Dn/PnD_{n}/P_{n}
E6/P1E_{6}/P_{1}
E7/P1E_{7}/P_{1}
E8/P1E_{8}/P_{1}
E6/P2E_{6}/P_{2}
E7/P2E_{7}/P_{2}
E8/P2E_{8}/P_{2}
E6/P6E_{6}/P_{6}
E7/P7E_{7}/P_{7}
E8/P8E_{8}/P_{8}
F4/P1F_{4}/P_{1}
F4/P4F_{4}/P_{4}
G2/P1G_{2}/P_{1}
G2/P2G_{2}/P_{2}
ς(𝒢)\varsigma(\mathcal{G})
n1n-1
n1n-1
2n32n-3
n1n-1
2n22n-2
n1n-1
2n52n-5
n1n-1
n1n-1
6
7
9
5
6
7
6
10
13
5
5
3
1
Theorem 3.4.

Let EE be a uniform rr-bundle on a generalized Grassmannian 𝒢\mathcal{G} with extremal node marked. If rς(𝒢)r\leq\varsigma(\mathcal{G}), then EE splits as a direct sum of line bundles.

Remark 3.5.

One can find the value of ς(𝒢)\varsigma(\mathcal{G}) is equal to or bigger than the value of ς\varsigma that we analyzed such that the morphism xG(t,ς)\mathcal{M}_{x}\rightarrow G(t,\varsigma) can only be constant. ς(𝒢)\varsigma(\mathcal{G}) is equal to ς\varsigma except 𝒢=Cn/P1\mathcal{G}=C_{n}/P_{1} and G2/P1G_{2}/P_{1}. The reason is that uniform 2n22n-2 bundles split on Cn/P12n1C_{n}/P_{1}\cong\mathbb{P}_{2n-1} and uniform 33 bundles split on G2/P1Q5G_{2}/P_{1}\cong Q^{5} by the previous arguments. In all cases, however, the morphism xG(t,ς(𝒢)2)\mathcal{M}_{x}\rightarrow G(t,\varsigma(\mathcal{G})-2) can only be constant.

Corollary 3.6.

If 𝒢\mathcal{G} is a generalized Grassmannian with marked point kk, kk is not extremal, then x\mathcal{M}_{x} is a product of rational homogeneous spaces. If

x=A1××At(2t3)\mathcal{M}_{x}=A_{1}\times\cdots\times A_{t}~{}(2\leq t\leq 3)

and

rς(𝒢):=min{ς(A1),,ς(At)},r\leq\varsigma(\mathcal{G}):=\text{min}\{\varsigma(A_{1}^{\prime}),\ldots,\varsigma(A_{t}^{\prime})\},

where Ai(1it)A_{i}~{}(1\leq i\leq t) is the special family of lines of class αkˇ\check{\alpha_{k}} through xx on the generalized Grassmannian AiA_{i}^{\prime} with extremal node marked. Then uniform rr-bundle EE on 𝒢\mathcal{G} splits as a direct sum of line bundles.

Proof.

In order to prove EE splits, we just need to prove that the only morphisms xG(t,r)\mathcal{M}_{x}\rightarrow G(t,r) are constant. Let ϕ\phi be a morphism from x\mathcal{M}_{x} to G(t,r)G(t,r). Because every Ai(1it)A_{i}~{}(1\leq i\leq t) can be regarded as the subspace of x\mathcal{M}_{x} corresponding to the special family of lines of class αkˇ\check{\alpha_{k}} through xx on AiA_{i}^{\prime}, we can consider the restriction of ϕ\phi to Ai(1it)A_{i}~{}(1\leq i\leq t). By assumption rmin{ς(A1),,ς(At)}r\leq\text{min}\{\varsigma(A_{1}^{\prime}),\ldots,\varsigma(A_{t}^{\prime})\}, all the restriction maps are constant. Hence, ϕ\phi is also constant. ∎

Remark 3.7.

  • For the case where x=1\mathcal{M}_{x}=\mathbb{P}^{1} or x=1×\mathcal{M}_{x}=\mathbb{P}^{1}\times\cdots, we can say nothing about the splitting result according to our theorem.

  • In some cases, the value of ς(𝒢)\varsigma(\mathcal{G}) cannot be expanded anymore, which means that there exist uniform but nonsplitting ς(𝒢)+1\varsigma(\mathcal{G})+1-bundles. For instance, Grassmannian An1/Pd=G(d,n)(dnd)A_{n-1}/P_{d}=G(d,n)~{}(d\leq n-d) has uniform but nonsplitting dd-bundle HdH_{d} (the universal bundle of G(d,n)G(d,n)); Spinor variety Dn+1/Pn+1=𝒮nD_{n+1}/P_{n+1}=\mathcal{S}_{n} has uniform but nonsplitting n+1n+1-bundle Qn+1Q_{n+1} (the universal quotient bundle of 𝒮n\mathcal{S}_{n}).

  • Compare to the main theorem (Theorem 3.1) in 18, we improve their results. In particular, we enlarge the splitting threshold for uniform bundles on Hermitian symmetric spaces E6/P6E_{6}/P_{6} from 5\geq 5 to 6\geq 6 and E7/P7E_{7}/P_{7} from 7\geq 7 to 10\geq 10 (see Table 1 in 18).

Corollary 3.8.

Let 𝒢\mathcal{G} be a generalized Grassmannian covered by linear projective subspaces of dimension 22 and EE be an rr-bundle on 𝒢\mathcal{G}. If EE splits as a direct sum of line bundles when it restricts to every 2𝒢\mathbb{P}^{2}\subseteq\mathcal{G}, then EE splits as a direct sum of line bundles on 𝒢\mathcal{G}.

Proof.

If 𝒢\mathcal{G} is a projective space, then the result holds (23 Theorem 2.3.2). Suppose 𝒢=G/Pk\mathcal{G}=G/P_{k} which is not a projective space and kk is the unique black node, where GG is a simple Lie group and PkP_{k} is a maximal parabolic subgroup of GG. The condition implies that EE is uniform. The reason for this is the fact every line LL is contained in two different 2\mathbb{P}^{2} by Theorem 2.9.

We prove the corollary by induction on rr. If we have the exact sequence of vector bundles

0MEN0\displaystyle 0\rightarrow M\rightarrow E\rightarrow N\rightarrow 0 (3.6)

on 𝒢\mathcal{G}, where the rank of MM and NN are smaller than rr, such that

M|Z=i=1rt𝒪Z(at+i),N|Z=𝒪Zt,M|_{Z}=\bigoplus\limits_{i=1}^{r-t}\mathcal{O}_{Z}(a_{t+i}),N|_{Z}=\mathcal{O}_{Z}^{\oplus t},

for every Z2Z\simeq\mathbb{P}^{2}, then by the induction hypothesis, MM and NN split. Since H1(𝒢,NM)=0H^{1}(\mathcal{G},N^{\vee}\otimes M)=0, the above exact sequence splits and hence also EE.

Similar to the proof of Theorem 3.3 in 6, on 𝒰=G/(PkPN(k))\mathcal{U}=G/(P_{k}\cap P_{N(k)}), we can obtain an exact sequence

0M~q1EN~0.0\rightarrow\widetilde{M}\rightarrow{q_{1}}^{\ast}E\rightarrow\widetilde{N}\rightarrow 0.

If we prove that the morphism φ\varphi is constant for every x𝒢x\in\mathcal{G}, then there exist two bundles MM, NN over GG with M~=q1M,N~=q1N\widetilde{M}={q_{1}}^{\ast}M,\widetilde{N}={q_{1}}^{\ast}N. By projecting the bundle sequence

0q1Mq1Eq1N00\rightarrow{q_{1}}^{\ast}M\rightarrow{q_{1}}^{\ast}E\rightarrow{q_{1}}^{\ast}N\rightarrow 0

onto 𝒢\mathcal{G}, we can get the desired exact sequence (3.6). Thus, to prove the existence of the above exact sequence, it suffices to show that the map

φ:x𝔾k(t1,k(Ex))\varphi:\mathcal{M}_{x}\rightarrow\mathbb{G}_{k}(t-1,\mathbb{P}_{k}(E^{\vee}_{x}))

is constant for every x𝒢x\in\mathcal{G} . Given a projective subspace ZZ of dimension 22 and a line LZL\subseteq Z, we take any point xLx\in L and denote by ZZ^{\prime} the subspace of x\mathcal{M}_{x} corresponding to the tangent directions to ZZ at xx. By the hypothesis, E|ZE|Z is a direct sum of line bundles, so

φ|Z:Z𝔾k(t1,k(Ex))\varphi|_{Z^{\prime}}:Z^{\prime}\rightarrow\mathbb{G}_{k}(t-1,\mathbb{P}_{k}(E^{\vee}_{x}))

is constant. Since 𝒢\mathcal{G} covered by linear projective subspaces of dimension 22 and MxM_{x} is chain-connected by 1\mathbb{P}^{1}, φ\varphi is constant for every x𝒢x\in\mathcal{G}. ∎

3.2 Uniform vector bundles on rational homogeneous spaces

Let

X=G/PG1/PI1×G2/PI2××Gm/PIm,X=G/P\simeq G_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}},

where GiG_{i} is a simple Lie group with Dynkin diagram 𝒟i\mathcal{D}_{i} whose set of nodes is DiD_{i} and PIiP_{I_{i}} is a parabolic subgroup of GiG_{i} corresponding to IiDiI_{i}\subset D_{i}. We set F(Ii):=Gi/PIiF(I_{i}):=G_{i}/P_{I_{i}} by marking on the Dynkin diagram 𝒟i\mathcal{D}_{i} of GiG_{i} the nodes corresponding to IiI_{i}. Let δi\delta_{i} be a node in 𝒟i\mathcal{D}_{i} and N(δi)N(\delta_{i}) be the set of nodes in 𝒟i\mathcal{D}_{i} that are connected to δi\delta_{i}.

If δiIi\delta_{i}\in I_{i}, we call

iδic:=Gi/Piδic×Gi/PIi^(1im),\mathcal{M}_{i}^{\delta_{i}^{c}}:=G_{i}/P_{i}^{\delta_{i}^{c}}\times\widehat{G_{i}/P_{I_{i}}}~{}(1\leq i\leq m),

the ii-th special family of lines of class δˇi\check{\delta}_{i} by Theorem 2.3, where Piδic:=PIi\δiN(δi)P_{i}^{\delta_{i}^{c}}:=P_{I_{i}\backslash\delta_{i}\cup N(\delta_{i})} and Gi/PIi^\widehat{G_{i}/P_{I_{i}}} is G1/PI1×G2/PI2××Gm/PImG_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}} by deleting ii-th term Gi/PIiG_{i}/P_{I_{i}}. For i=1i=1 and δI\delta\in I, we will use the notation δ\mathcal{M}^{\delta} to denote the special family of lines of class δˇ\check{\delta}.

For xXx\in X, we call

SPVMRTx(δi)={Liδic|xL}\text{SPVMRT}_{x}^{(\delta_{i})}=\{L\in\mathcal{M}_{i}^{\delta_{i}^{c}}|x\in L\}

the δi\delta_{i}-th special part of variety of minimal rational tangents at xx (sometimes we just write SPVMRTx\text{SPVMRT}_{x} if there is no confusion).

Fix δiIi\delta_{i}\in I_{i}. SPVMRTx(δi)\text{SPVMRT}_{x}^{(\delta_{i})} is just the special family of lines of class δˇi\check{\delta}_{i} through xx on the generalized Grassmannian 𝒢δi\mathcal{G}^{\delta_{i}} whose Dynkin diagram 𝒟δi\mathcal{D}^{\delta_{i}} is the maximal sub-diagram of (𝒟i,Ii)(\mathcal{D}_{i},I_{i}) with the only marked point δi\delta_{i}. Denote ν(X,δi):=ς(𝒢δi)\nu(X,\delta_{i}):=\varsigma(\mathcal{G}^{\delta_{i}}). Let

ν(X):=mini{minδiIi{ν(X,δi)}}.\nu(X):=\text{min}_{i}\{\text{min}_{\delta_{i}\in I_{i}}\{\nu(X,\delta_{i})\}\}.
Definition 3.9.

A vector bundle EE on XX is called poly-uniform with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}} for every i(1im)i~{}(1\leq i\leq m) and every δiIi\delta_{i}\in I_{i} if the restriction of EE to every line in iδic\mathcal{M}_{i}^{\delta_{i}^{c}} has the same splitting type. We also call that EE poly-uniform with respect to all the special families of lines.

Let \mathscr{F} be a torsion free coherent sheaf of rank rr over XX. Fix integer i(1im)i~{}(1\leq i\leq m) and δiIi\delta_{i}\in I_{i}. Since the singularity set S()S(\mathscr{F}) of \mathscr{F} has codimension at least 22, there are lines LiδicL\in\mathcal{M}_{i}^{\delta_{i}^{c}} which do not meet S()S(\mathscr{F}). If

|L𝒪L(a1(δi))𝒪L(ar(δi)).\mathscr{F}|L\cong\mathcal{O}_{L}(a_{1}^{(\delta_{i})})\oplus\cdots\oplus\mathcal{O}_{L}(a_{r}^{(\delta_{i})}).

Let

c1(δi)()=a1(δi)++ar(δi),c_{1}^{(\delta_{i})}(\mathscr{F})=a_{1}^{(\delta_{i})}+\cdots+a_{r}^{(\delta_{i})},

which is independent of the choice of LL. We set

μ(δi)()=c1(δi)()rk().\mu^{(\delta_{i})}(\mathscr{F})=\frac{c_{1}^{(\delta_{i})}(\mathscr{F})}{\text{rk}(\mathscr{F})}.
Definition 3.10.

A torsion free coherent sheaf \mathscr{E} over XX is δi\delta_{i}-semistable (δi\delta_{i}-stable) if for every coherent subsheaf \mathscr{F}\subseteq\mathcal{E} with 0<rk()<rk()0<\text{rk}(\mathscr{F})<\text{rk}(\mathcal{E}), we have

μ(δi)()(<)μ(δi)().\mu^{(\delta_{i})}(\mathscr{F})\leq~{}(<)~{}\mu^{(\delta_{i})}(\mathcal{E}).

If EE is not δi\delta_{i}-semistable, then we call EE is δi\delta_{i}-unstable.

Proposition 3.11.

Fix integer i(1im)i~{}(1\leq i\leq m) and δiIi\delta_{i}\in I_{i}. Let EE be a uniform rr-bundle on XX of type (a1(δi),,ar(δi)),a1(δi)ar(δi)(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),~{}a_{1}^{(\delta_{i})}\leq\ldots\leq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. If rν(X,δi)2r\leq\nu(X,\delta_{i})-2 and these aj(δi)a_{j}^{(\delta_{i})}’s are not all same, then EE can be expressed as an extension of uniform bundles with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. In particular, EE is δi\delta_{i}-unstable.

Proof.

After twisting with an appropriate line bundle, we can assume that EE has the splitting type

a¯E(δi)=(0,,0,at+1(δi),,ar(δi)),at+i(δi)>0,fori=1,,rt.\underline{a}_{E}^{(\delta_{i})}=(0,\ldots,0,a_{t+1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),~{}a_{t+i}^{(\delta_{i})}>0,~{}\text{for}~{}i=1,\ldots,r-t.

with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}.

Let’s consider the standard diagram

(3.11)

For LiδiL\in\mathcal{M}_{i}^{\delta_{i}}, the q2q_{2}-fiber

L~=q21(L)={(x,L)X×iδic|xL},\widetilde{L}={q_{2}}^{-1}(L)=\{(x,L)\in X\times\mathcal{M}_{i}^{\delta_{i}^{c}}|x\in L\},

is mapped under q1q_{1} to the line LL identically in G/PG/P and we have

q1E|L~E|L.{q_{1}}^{\ast}E|_{\widetilde{L}}\cong E|_{L}.

For xG/Px\in G/P, the q1q_{1}-fiber q11(x){q_{1}}^{-1}(x) is mapped isomorphically under q2q_{2} to the subvariety

SPVMRTx={Liδic|xL}.\text{SPVMRT}_{x}=\{L\in\mathcal{M}_{i}^{\delta_{i}^{c}}|x\in L\}.

Because

E|L𝒪Lti=1rt𝒪L(at+i(δi)),at+i(δi)>0,E|_{L}\cong\mathcal{O}_{L}^{\oplus t}\oplus\bigoplus\limits_{i=1}^{r-t}\mathcal{O}_{L}(a_{t+i}^{(\delta_{i})}),~{}a_{t+i}^{(\delta_{i})}>0,
h0(q21(L),q1(E)|q21(L))=th^{0}\left({q_{2}}^{-1}(L),{q_{1}}^{\ast}(E^{\vee})|_{q_{2}^{-1}(L)}\right)=t

for all LiδicL\in\mathcal{M}_{i}^{\delta_{i}^{c}}. Thus the direct image q2q1(E){q_{2}}_{\ast}{q_{1}}^{\ast}(E^{\vee}) is a vector bundle of rank tt over iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. The canonical homomorphism of sheaves

q2q2q1(E)q1(E){q_{2}}^{\ast}{q_{2}}_{\ast}{q_{1}}^{\ast}(E^{\vee})\rightarrow{q_{1}}^{\ast}(E^{\vee})

makes N~:=q2q2q1(E)\widetilde{N}^{\vee}:={q_{2}}^{\ast}{q_{2}}_{\ast}{q_{1}}^{\ast}(E^{\vee}) to be a subbundle of q1(E){q_{1}}^{\ast}(E^{\vee}). Because over each q2{q_{2}}-fiber L~\widetilde{L}, the evaluation map

N~|L~=H0(L~,q1(E)|L~)k𝒪L~q1(E)|L~\widetilde{N}^{\vee}|_{\widetilde{L}}=H^{0}(\widetilde{L},{q_{1}}^{\ast}(E^{\vee})|_{\widetilde{L}})\otimes_{k}\mathcal{O}_{\widetilde{L}}\rightarrow{q_{1}}^{\ast}(E^{\vee})|_{\widetilde{L}}

identifies N~|L~\widetilde{N}^{\vee}|_{\widetilde{L}} with 𝒪Lt𝒪Lti=1rt𝒪L(at+i(δi))=E|L.\mathcal{O}_{L}^{\oplus t}\subseteq\mathcal{O}_{L}^{\oplus t}\oplus\bigoplus_{i=1}^{r-t}\mathcal{O}_{L}(-a_{t+i}^{(\delta_{i})})=E^{\vee}|_{L}. Over 𝒰iδic\mathcal{U}_{i}^{\delta_{i}^{c}} we thus obtain an exact sequence

0M~q1EN~00\rightarrow\widetilde{M}\rightarrow{q_{1}}^{\ast}E\rightarrow\widetilde{N}\rightarrow 0

of vector bundles, whose restriction to q2{q_{2}}-fibers L~\widetilde{L} looks as follows:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M~|L~\textstyle{\widetilde{M}|_{\widetilde{L}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}q1E|L~\textstyle{{q_{1}}^{\ast}E|_{\widetilde{L}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}N~|L~\textstyle{\widetilde{N}|_{\widetilde{L}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i=1rt𝒪L(at+i(δi))\textstyle{\bigoplus_{i=1}^{r-t}\mathcal{O}_{L}(a_{t+i}^{(\delta_{i})})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪Lti=1rt𝒪L(at+i(δi))\textstyle{\mathcal{O}_{L}^{\oplus t}\oplus\bigoplus_{i=1}^{r-t}\mathcal{O}_{L}(a_{t+i}^{(\delta_{i})})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪Lt\textstyle{\mathcal{O}_{L}^{\oplus t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

Because N~\widetilde{N}^{\vee} is a subbundle of q1(E){q_{1}}^{\ast}(E^{\vee}) of rank tt, for every point xGx\in G, it provides a morphism

φ:SPVMRTx𝔾k(t1,k(Ex)).\varphi:\text{SPVMRT}_{x}\rightarrow\mathbb{G}_{k}(t-1,\mathbb{P}_{k}(E^{\vee}_{x})).

Since rν(X,δi)2r\leq\nu(X,\delta_{i})-2, morphism φ\varphi is constant by Remark 3.5 and the definition of ν(X,δi)\nu(X,\delta_{i}). It follows that M~\widetilde{M} and N~\widetilde{N} are trivial on all q1{q_{1}}-fibers. So the canonical morphisms q1q1M~M~{q_{1}}^{\ast}{q_{1}}_{\ast}\widetilde{M}\rightarrow\widetilde{M} and q1q1N~N~{q_{1}}^{\ast}{q_{1}}_{\ast}\widetilde{N}\rightarrow\widetilde{N} are isomorphisms. Hence there are uniform bundles M=q1M~M={q_{1}}_{\ast}\widetilde{M}, N=q1N~N={q_{1}}_{\ast}\widetilde{N} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}} over XX with

M~=q1M,N~=q1N.\widetilde{M}={q_{1}}^{\ast}M,\widetilde{N}={q_{1}}^{\ast}N.

By projection formula and the rationality of the q1q_{1} fiber SPVMRTx, we project the bundle sequence

0q1Mq1Eq1N00\rightarrow{q_{1}}^{\ast}M\rightarrow{q_{1}}^{\ast}E\rightarrow{q_{1}}^{\ast}N\rightarrow 0

onto XX to get the exact sequence

0MEN0.\displaystyle 0\rightarrow M\rightarrow E\rightarrow N\rightarrow 0. (3.12)

So μ(δi)(N)>μ(δi)(E)\mu^{(\delta_{i})}(N)>\mu^{(\delta_{i})}(E) and thus EE is δi\delta_{i}-unstable ∎

Proposition 3.12.

On XX, if an rr-bundle EE is poly-uniform with respect to all the special families of lines on XX such that the splitting type with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}} is (a(δi),,a(δi))(a^{(\delta_{i})},\ldots,a^{(\delta_{i})}) for each ii and δi\delta_{i}, then EE splits as a direct sum of line bundles.

Proof.

After twisting with an appropriate line bundle, we can assume that EE is trivial on all the special families of lines on XX. We are going to show that EE is trivial.

Let’s first consider the case that XX is a generalized flag manifold, which corresponds to a connected marked Dynkin diagram 𝒟(X)\mathcal{D}(X) with ll black nodes. We prove the lemma by induction on ll. For l=1l=1, XX is just a generalized Grassmannian. Then the result holds by 1 Proposition 1.2. Suppose the assertion is true for all generalized flag manifolds with connected marked Dynkin diagram and ll^{\prime} black nodes (1l<l1\leq l^{\prime}<l). Let’s consider the natural projection

π:XX,\pi:X\rightarrow X^{\prime},

where XX^{\prime} is corresponding to the marked Dynkin diagram 𝒟(X)\mathcal{D}(X) by changing the first black node δ\delta to white. It’s not hard to see that every π\pi-fiber π1(x)\pi^{-1}(x) is isomorphic to the generalized Grassmannian 𝒢δ\mathcal{G}^{\delta} with the only marked point δ\delta. Since the restriction of EE to every line in δ\mathcal{M}^{\delta} is trivial, so is to every line in π1(x)\pi^{-1}(x). Thus EE is trivial on all π\pi-fibers by 1 Proposition 1.2. It follows that E=πEE^{\prime}=\pi_{*}E is an algebraic vector bundle of rank rr over XX^{\prime} and EπEE\cong\pi^{*}E^{\prime}.

Claim. EE^{\prime} is trivial on all the special families of lines on XX^{\prime}.

In fact, let γ\gamma be a black node in 𝒟(X)\mathcal{D}(X) that is different from δ\delta and LL be a line in γ\mathcal{M}^{\gamma}. Then π(L)\pi(L) is a line in the family of lines of XX^{\prime}. When γ\gamma runs through all black nodes except δ\delta and LL runs through all lines in γ\mathcal{M}^{\gamma} in XX, π(L)\pi(L) also runs through all lines in all the special families of lines of XX^{\prime}. The projection π\pi induces an isomorphism

E|π(L)πE|LE|L.E^{\prime}|_{\pi(L)}\cong\pi^{*}E^{\prime}|_{L}\cong E|_{L}.

We identify LL with π(L)\pi(L). Since E|LE|_{L} is trivial for every line LL in γ\mathcal{M}^{\gamma} by assumption, E|LE^{\prime}|_{L} is trivial for every line LL in all the special families of lines of XX^{\prime}. By the induction hypothesis, EE^{\prime} is trivial. Thus EqEE\cong q^{*}E^{\prime} is trivial.

Now let’s think about the general case, where the marked Dynkin diagram of XX is not connected. Assume XX can be decomposed into a product

X=G/PG1/PI1×G2/PI2××Gm/PIm,X=G/P\simeq G_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}},

where m2m\geq 2 and Gi(1im)G_{i}~{}(1\leq i\leq m) is a simple Lie group with connected Dynkin diagram. We prove EE is trivial by induction on mm. For m=1m=1, the result holds from the previous analysis. Consider the natural projection

f:XX:=G2/PI2××Gm/PIm,f:X\rightarrow X^{\prime}:=G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}},

it’s easy to see that every ff-fiber f1(x)f^{-1}(x) is isomorphic to G1/PI1G_{1}/P_{I_{1}}. By assumption, EE is trivial on all the special families of lines on f1(x)G1/PI1f^{-1}(x)\cong G_{1}/P_{I_{1}}. Thus EE is trivial on all ff-fibers by the previous analysis. It follows that E=fEE^{\prime}=f_{*}E is an algebraic vector bundle of rank rr over XX^{\prime} and EfEE\cong f^{*}E^{\prime}. Similarly, we can prove that EE^{\prime} is trivial and thus EE is trivial. ∎

Theorem 3.13.

On XX, if rr-bundle EE is poly-uniform with respect to all the special families of lines and rν(X)2r\leq\nu(X)-2, then EE is δi\delta_{i}-unstable for some δi\delta_{i} (1im1\leq i\leq m) or EE splits as a direct sum of line bundles.

Proof.

It is obviously from Proposition 3.11 and 3.12. ∎

4 Semistable vector bundles on rational homogeneous spaces

Let GG be a simple Lie group and D={1,2,,n}D=\{1,2,\ldots,n\} be the set of nodes of the Dynkin diagram 𝒟\mathcal{D} of GG. Denote by PkP_{k} the parabolic subgroup of GG corresponding to the node kk. Consider X=G/PkX=G/P_{k} in its minimal homogeneous embedding. Denote by :=G/PN(k)\mathcal{M}:=G/P_{N(k)} the generalized flag manifold defined by the marked Dynkin diagram (𝒟,N(k))(\mathcal{D},N(k)) and by 𝒰:=G/Pk,N(k)=G/(PN(k)Pk)\mathcal{U}:=G/P_{k,N(k)}=G/(P_{N(k)}\cap P_{k}) the universal family, which has a natural 1\mathbb{P}^{1}-bundle structure over \mathcal{M}, i.e. we have the natural diagram:

(4.5)

Given xXx\in X, x=q(p1(x))\mathcal{M}_{x}=q(p^{-1}(x)) coincides with H/PN(k)H/P_{N(k)} by Remark 2.7 where the set of nodes of the Dynkin diagram HH is D(H)=D\kD(H)=D\backslash{k} .

In the above setting, we will show that the splitting type of T𝒰/X|q1(l)T_{\mathcal{U}/X}|_{q^{-1}(l)} take the form (1,,1)(-1,\ldots,-1), (1,,1,2,,2)(-1,\ldots,-1,-2,\ldots,-2) or (3)(-3). Building upon this assert, we will generalize the Grauert-Mu¨\ddot{\text{u}}lich-Barth theorem to any rational homogeneous spaces.

4.1 The classical simple Lie algebras

We say GG is of classical type when its Dynkin diagram is of type AnA_{n}, BnB_{n}, CnC_{n} or DnD_{n}. Because X=G/PkX=G/P_{k} has clear geometric explanations, we can even write down the specific form of the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}. In the case of type AnA_{n}, the relative tangent bundle T𝒰/XT_{\mathcal{U}/X} is known (see 6 Lemma 5.6) and the splitting type of T𝒰/X|q1(l)T_{\mathcal{U}/X}|_{q^{-1}(l)} is (1,,1)(-1,\ldots,-1). So let’s just consider the remaining three types.

4.1.1 The marked Dynkin diagram (Bn,k)(B_{n},k)

In this section, we consider the Dynkin diagram BnB_{n}, which corresponds to the classical Lie group SO2n+1SO_{2n+1}. Denote by Bn/PI:=SO2n+1/PIB_{n}/P_{I}:=SO_{2n+1}/P_{I} the generalized flag manifold with I={i1,,is}DI=\{i_{1},\ldots,i_{s}\}\subseteq D. Let V=2n+1V=\mathbb{C}^{2n+1} be a vector space equipped with a nondegenerate symmetric bilinear form 𝒬\mathcal{Q}. Then Bn/PIB_{n}/P_{I} is actually the odd Orthogonal flag manifold OG(i1,,is;2n+1)OG(i_{1},\ldots,i_{s};2n+1), which parametrizes flags

Vi1Vi2VisV,V_{i_{1}}\subset V_{i_{2}}\subset\cdots\subset V_{i_{s}}\subset V,

where each Vit(1ts)V_{i_{t}}(1\leq t\leq s) is an iti_{t}-dimensional isotropic subspace in VV.

There is a universal flag of subbundles

0=H0Hi1HisHisHi1𝒪Bn/PI×V0=H_{0}\subset H_{i_{1}}\subset\cdots\subset H_{i_{s}}\subset H_{i_{s}}^{\bot}\subset\cdots\subset H_{i_{1}}^{\bot}\subset\mathcal{O}_{B_{n}/P_{I}}\times V

on Bn/PIB_{n}/P_{I}, where HitH_{i_{t}}^{\bot} is the 𝒬\mathcal{Q}-orthogonal complement of HitH_{i_{t}}, rank Hit=itH_{i_{t}}=i_{t} and rank Hit=2n+1it(1ts)H_{i_{t}}^{\bot}=2n+1-i_{t}~{}(1\leq t\leq s).

(1) For k=1k=1, the odd Orthogonal Grassmannian Bn/P1:=OG(1,2n+1)B_{n}/P_{1}:=OG(1,2n+1) is just the quadric Q2n1Q^{2n-1}. In this case, =Bn/P2\mathcal{M}=B_{n}/P_{2} and 𝒰=Bn/P1,2\mathcal{U}=B_{n}/P_{1,2}. We have the natural diagram

(4.10)

and x\mathcal{M}_{x} is Bn1/P1B_{n-1}/P_{1}, i.e. the quadric Q2n3Q^{2n-3}.

Lemma 4.1.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)2n3.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-3}.
Proof.

For xXx\in X, the pp-fiber p1(x)={(x,l)|xL}p^{-1}(x)=\{(x,l)|x\in L\} is isomorphic to x=Q2n3\mathcal{M}_{x}=Q^{2n-3}. Over p1(x)Q2n3p^{-1}(x)\cong Q^{2n-3}, we have the universal bundle sequence

0𝒪Q2n3(1)𝒪Q2n3(1)𝒪Q2n3(1)/𝒪Q2n3(1)0,\displaystyle 0\rightarrow\mathcal{O}_{Q^{2n-3}}(-1)\rightarrow\mathcal{O}_{Q^{2n-3}}(-1)^{\bot}\rightarrow\mathcal{O}_{Q^{2n-3}}(-1)^{\bot}/\mathcal{O}_{Q^{2n-3}}(-1)\rightarrow 0, (4.11)

where 𝒪Q2n3(1)\mathcal{O}_{Q^{2n-3}}(-1) is the rank 1 tautological bundle over Q2n3Q^{2n-3}, which is the pull back of 𝒪2n2(1)\mathcal{O}_{\mathbb{P}^{2n-2}}(-1) under the embedding Q2n32n2Q^{2n-3}\hookrightarrow\mathbb{P}^{2n-2} and 𝒪Q2n3(1)\mathcal{O}_{Q^{2n-3}}(-1)^{\bot} denotes its 𝒬\mathcal{Q}-orthogonal complement. Notice that the tangent bundle of Q2n3Q^{2n-3} can be represented as

TQ2n3=𝒪Q2n3(1)𝒪Q2n3(1).T_{Q^{2n-3}}=\mathcal{O}_{Q^{2n-3}}(1)\otimes\mathcal{O}_{Q^{2n-3}}(-1)^{\bot}.

Let’s consider the exact sequence

0H2/H1H2/H1q(H2/H2)0,0\rightarrow H_{2}/H_{1}\rightarrow H_{2}^{\bot}/H_{1}\rightarrow q^{*}(H_{2}^{\bot}/H_{2})\rightarrow 0,

of vector bundles on 𝒰=Bn/P1,2\mathcal{U}=B_{n}/P_{1,2}. By restricting it to the the pp-fiber p1(x)p^{-1}(x), we obtain the universal bundle sequence (4.11) on p1(x)p^{-1}(x). Therefore, T𝒰/X=(H2/H1)q(H2/H2)T_{\mathcal{U}/X}=(H_{2}/H_{1})^{\vee}\otimes q^{*}(H_{2}^{\bot}/H_{2}) and

T𝒰/X|L~=𝒪L~(1)2n3.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-3}.

(2) For k(2kn1)k~{}(2\leq k\leq n-1), the odd Orthogonal Grassmannian Bn/Pk:=OG(k,2n+1)B_{n}/P_{k}:=OG(k,2n+1) parametrizes the kk-dimensional isotropic subspaces in VV. In this case, =Bn/Pk1,k+1\mathcal{M}=B_{n}/P_{k-1,k+1} and 𝒰=Bn/Pk1,k,k+1\mathcal{U}=B_{n}/P_{k-1,k,k+1}. We have the natural diagram:

(4.16)

and x\mathcal{M}_{x} is k1×Bnk/P1=k1×Q2(nk)1\mathbb{P}^{k-1}\times B_{n-k}/P_{1}=\mathbb{P}^{k-1}\times Q^{2(n-k)-1}.

Lemma 4.2.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)2nk2.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-k-2}.
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following two exact sequences of vector bundles:

0(Hk/Hk1)pHkqHk10,\displaystyle 0\rightarrow(H_{k}/H_{k-1})^{\vee}\rightarrow p^{\ast}H_{k}^{\vee}\rightarrow q^{\ast}H_{k-1}^{\vee}\rightarrow 0,
0Hk+1/HkHk+1/Hkq(Hk+1/Hk+1)0.\displaystyle 0\rightarrow H_{k+1}/H_{k}\rightarrow H_{k+1}^{\bot}/H_{k}\rightarrow q^{*}(H_{k+1}^{\bot}/H_{k+1})\rightarrow 0.

We will consider the following diagram

(4.23)

All the morphisms in the above diagram are projections. For any xXx\in X, the p1p_{1}-fiber p11(x)p_{1}^{-1}(x) is isomorphic to k1\mathbb{P}^{k-1} and the p2p_{2}-fiber p21(x)p_{2}^{-1}(x) is isomorphic to Q2(nk)1Q^{2(n-k)-1}. Note that for xXx\in X, the projection qq induces an isomorphism

q|p1(x):p1(x)x=k1×Q2(nk)1.q|p^{-1}(x):p^{-1}(x)\rightarrow\mathcal{M}_{x}=\mathbb{P}^{k-1}\times Q^{2(n-k)-1}.

Hence p1(x)=p11(x)×p21(x)p^{-1}(x)=p_{1}^{-1}(x)\times p_{2}^{-1}(x). So we get

T𝒰/X=π1(T𝒰1/X)π2(T𝒰2/X).T_{\mathcal{U}/X}=\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\oplus\pi_{2}^{*}(T_{\mathcal{U}_{2}/X}).

We mimic the proof of Lemma 5.6 in 6 to conclude that

π1(T𝒰1/X)Hk/Hk1qHk1,π2(T𝒰2/X)(Hk+1/Hk)q(Hk+1/Hk+1).\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\cong H_{k}/H_{k-1}\otimes q^{\ast}H_{k-1}^{\vee},\pi_{2}^{*}(T_{\mathcal{U}_{2}/X})\cong(H_{k+1}/H_{k})^{\vee}\otimes q^{*}(H_{k+1}^{\bot}/H_{k+1}).

Therefore, T𝒰/X(Hk/Hk1qHk1)((Hk+1/Hk)q(Hk+1/Hk+1))T_{\mathcal{U}/X}\cong\big{(}H_{k}/H_{k-1}\otimes q^{\ast}H_{k-1}^{\vee}\big{)}\bigoplus\big{(}(H_{k+1}/H_{k})^{\vee}\otimes q^{*}(H_{k+1}^{\bot}/H_{k+1})\big{)} and

T𝒰/X|L~=𝒪L~(1)2nk2.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-k-2}.

(3) For k=nk=n (corresponding to the short root αn\alpha_{n}), the odd Orthogonal Grassmannian Bn/Pn:=OG(n,2n+1)B_{n}/P_{n}:=OG(n,2n+1) parametrizes the nn-dimensional isotropic subspaces in VV. In this case, =Bn/Pn1\mathcal{M}=B_{n}/P_{n-1} and 𝒰=Bn/Pn1,n\mathcal{U}=B_{n}/P_{n-1,n}. We have the natural diagram

(4.28)

and x\mathcal{M}_{x} is n1\mathbb{P}^{n-1}.

Lemma 4.3.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(2)n1.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-2)^{\oplus n-1}.
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following exact sequence of vector bundles:

0(Hn/Hn1)pHnqHn10.0\rightarrow(H_{n}/H_{n-1})^{\vee}\rightarrow p^{\ast}H_{n}^{\vee}\rightarrow q^{\ast}H_{n-1}^{\vee}\rightarrow 0.

Restricting it to the pp-fiber p1(x)n1p^{-1}(x)\cong\mathbb{P}^{n-1} is just the Euler sequence on n1\mathbb{P}^{n-1}. In fact, the projection pp identifies 𝒰\mathcal{U} in a canonical fashion with the projective bundle (Hn)\mathbb{P}(H_{n}) of XX. Hence T𝒰/XHn/Hn1qHn1T_{\mathcal{U}/X}\cong H_{n}/H_{n-1}\otimes q^{\ast}H_{n-1}^{\vee} and

T𝒰/X|L~=𝒪L~(2)n1.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-2)^{\oplus n-1}.

4.1.2 The marked Dynkin diagram (Cn,k)(C_{n},k)

In this section, we consider the Dynkin diagram CnC_{n}, which corresponds to the classical Lie group Sp2nSp_{2n}. Denote by Cn/PI:=Sp2n/PIC_{n}/P_{I}:=Sp_{2n}/P_{I} the generalized flag manifold with I={i1,,is}DI=\{i_{1},\ldots,i_{s}\}\subseteq D. Let V=2nV=\mathbb{C}^{2n} be a vector space equipped with a nondegenerate skew-symmetric bilinear form Ω\Omega. Then Cn/PIC_{n}/P_{I} is actually the Lagrangian flag manifold LG(i1,,is;2n)LG(i_{1},\ldots,i_{s};2n), which parametrizing flags

Vi1Vi2VisV,V_{i_{1}}\subset V_{i_{2}}\subset\cdots\subset V_{i_{s}}\subset V,

where each Vit(1ts)V_{i_{t}}(1\leq t\leq s) is an iti_{t}-dimensional isotropic subspace in VV.

There is universal flag of subbundles

0=H0Hi1HisHisHi1𝒪Cn/PI×V0=H_{0}\subset H_{i_{1}}\subset\cdots\subset H_{i_{s}}\subset H_{i_{s}}^{\bot}\subset\cdots\subset H_{i_{1}}^{\bot}\subset\mathcal{O}_{C_{n}/P_{I}}\times V

on Cn/PIC_{n}/P_{I}, where HitH_{i_{t}}^{\bot} is the Ω\Omega-orthogonal complement of HitH_{i_{t}}, rank Hit=itH_{i_{t}}=i_{t} and rank Hit=2nit(1ts)H_{i_{t}}^{\bot}=2n-i_{t}~{}(1\leq t\leq s).

(1) For k=1k=1 (corresponding to the short root α1\alpha_{1}), the Lagrangian Grassmannian Cn/P1:=LG(1,2n)C_{n}/P_{1}:=LG(1,2n) is just the projective space 2n1\mathbb{P}^{2n-1}. In this case, =Cn/P2\mathcal{M}=C_{n}/P_{2} and 𝒰=Cn/P1,2\mathcal{U}=C_{n}/P_{1,2}. We have the natural diagram

(4.33)

and x\mathcal{M}_{x} is Cn1/P1C_{n-1}/P_{1}, i.e. the the projective space 2n3\mathbb{P}^{2n-3}.

Lemma 4.4.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)2n4𝒪L~(2).T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-4}\oplus\mathcal{O}_{\widetilde{L}}(-2).
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following exact sequence of vector bundles:

0H2/H1p(H1/H1)H1/H20.0\rightarrow H_{2}/H_{1}\rightarrow p^{*}(H_{1}^{\bot}/H_{1})\rightarrow H_{1}^{\bot}/H_{2}\rightarrow 0.

Restricting it to the pp-fiber p1(x)2n3p^{-1}(x)\cong\mathbb{P}^{2n-3} is just the Euler sequence on 2n3\mathbb{P}^{2n-3}. In fact, the projection pp identifies 𝒰\mathcal{U} in a canonical fashion with the projective bundle (H1/H1)\mathbb{P}(H_{1}^{\bot}/H_{1}) of XX. Hence T𝒰/X(H2/H1)H1/H2T_{\mathcal{U}/X}\cong(H_{2}/H_{1})^{\vee}\otimes H_{1}^{\bot}/H_{2} and

T𝒰/X|L~=𝒪L~(1)2n4𝒪L~(2).T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-4}\oplus\mathcal{O}_{\widetilde{L}}(-2).

(2) For k(2kn1)k~{}(2\leq k\leq n-1) (corresponding to the short root αk\alpha_{k}), the Lagrangian Grassmannian Cn/Pk:=LG(k,2n+1)C_{n}/P_{k}:=LG(k,2n+1) parametrizes the kk-dimensional isotropic subspaces in VV. In this case =Cn/Pk1,k+1\mathcal{M}=C_{n}/P_{k-1,k+1} and 𝒰=Cn/Pk1,k,k+1\mathcal{U}=C_{n}/P_{k-1,k,k+1}. We have the natural diagram

(4.38)

and x\mathcal{M}_{x} is k1×Cnk/P1=k1×2(nk)1\mathbb{P}^{k-1}\times C_{n-k}/P_{1}=\mathbb{P}^{k-1}\times\mathbb{P}^{2(n-k)-1}.

Lemma 4.5.

Let L~=q1(l)𝒰k\widetilde{L}=q^{-1}(l)\subset\mathcal{U}^{k} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)2nk3𝒪L~(2).T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-k-3}\oplus\mathcal{O}_{\widetilde{L}}(-2).
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following two exact sequences of vector bundles:

0(Hk/Hk1)pHkqHk10,\displaystyle 0\rightarrow(H_{k}/H_{k-1})^{\vee}\rightarrow p^{\ast}H_{k}^{\vee}\rightarrow q^{\ast}H_{k-1}^{\vee}\rightarrow 0,
0Hk+1/Hkp(Hk/Hk)Hk/Hk+10.\displaystyle 0\rightarrow H_{k+1}/H_{k}\rightarrow p^{*}(H_{k}^{\bot}/H_{k})\rightarrow H_{k}^{\bot}/H_{k+1}\rightarrow 0.

We will consider the following diagram

(4.45)

All the morphisms in the above diagram are projections. For any xXx\in X, the p1p_{1}-fiber p11(x)p_{1}^{-1}(x) is isomorphic to k1\mathbb{P}^{k-1} and the p2p_{2}-fiber p21(x)p_{2}^{-1}(x) is isomorphic to 2(nk)1\mathbb{P}^{2(n-k)-1}. Note that for xXx\in X, the projection qq induces an isomorphism

q|p1(x):p1(x)x=k1×2(nk)1.q|_{p^{-1}(x)}:p^{-1}(x)\rightarrow\mathcal{M}_{x}=\mathbb{P}^{k-1}\times\mathbb{P}^{2(n-k)-1}.

Hence p1(x)=p11(x)×p21(x)p^{-1}(x)=p_{1}^{-1}(x)\times p_{2}^{-1}(x). So we get

T𝒰/X=π1(T𝒰1/X)π2(T𝒰2/X).T_{\mathcal{U}/X}=\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\oplus\pi_{2}^{*}(T_{\mathcal{U}_{2}/X}).

We mimic the proof of Lemma 5.6 in 6 to conclude that

π1(T𝒰1/X)Hk/Hk1qHk1,π2(T𝒰2/X)(Hk+1/Hk)Hk/Hk+1.\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\cong H_{k}/H_{k-1}\otimes q^{\ast}H_{k-1}^{\vee},\pi_{2}^{*}(T_{\mathcal{U}_{2}/X})\cong(H_{k+1}/H_{k})^{\vee}\otimes H_{k}^{\bot}/H_{k+1}.

Therefore, T𝒰/X(Hk/Hk1qHk1)(Hk+1/Hk)Hk/Hk+1)T_{\mathcal{U}/X}\cong\big{(}H_{k}/H_{k-1}\otimes q^{\ast}H_{k-1}^{\vee}\big{)}\bigoplus\big{(}H_{k+1}/H_{k})^{\vee}\otimes H_{k}^{\bot}/H_{k+1}\big{)} and

T𝒰/X|L~=𝒪L~(1)2nk3𝒪L~(2).T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-k-3}\oplus\mathcal{O}_{\widetilde{L}}(-2).

(3) For k=nk=n (corresponding to the long root αn\alpha_{n}), the Lagrangian Grassmannian Cn/Pn:=LG(n,2n)C_{n}/P_{n}:=LG(n,2n) parametrizes the nn-dimensional isotropic subspaces in VV. In this case =Cn/Pn1\mathcal{M}=C_{n}/P_{n-1} and 𝒰=Cn/Pn1,n\mathcal{U}=C_{n}/P_{n-1,n}. We have the natural diagram

(4.50)

and x\mathcal{M}_{x} is n1\mathbb{P}^{n-1}.

Lemma 4.6.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)n1.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus n-1}.
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following exact sequence of vector bundles:

0(Hn/Hn1)pHnqHn10.0\rightarrow(H_{n}/H_{n-1})^{\vee}\rightarrow p^{\ast}H_{n}^{\vee}\rightarrow q^{\ast}H_{n-1}^{\vee}\rightarrow 0.

Restricting it to the pp-fiber p1(x)n1p^{-1}(x)\cong\mathbb{P}^{n-1} is just the Euler sequence on n1\mathbb{P}^{n-1}. In fact, the projection pp identifies 𝒰\mathcal{U} in a canonical fashion with the projective bundle (Hn)\mathbb{P}(H_{n}) of XX. Hence T𝒰/XHn/Hn1qHn1T_{\mathcal{U}/X}\cong H_{n}/H_{n-1}\otimes q^{\ast}H_{n-1}^{\vee} and

T𝒰/X|L~=𝒪L~(1)n1.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus n-1}.

4.1.3 The marked Dynkin diagram (Dn,k)(D_{n},k)

In this section, we consider the Dynkin diagram DnD_{n}, it corresponds to the classical Lie group SO2nSO_{2n}. Denote by Dn/PI:=SO2n/PID_{n}/P_{I}:=SO_{2n}/P_{I} the generalized flag manifold with I={i1,,is}DI=\{i_{1},\ldots,i_{s}\}\subseteq D. Let V=2nV=\mathbb{C}^{2n} be a vector space equipped with a nondegenerate symmetric bilinear form 𝒬\mathcal{Q}. Then

Case 1. n1,nIn-1,n\in I. Dn/PID_{n}/P_{I} is the even Orthogonal flag manifold OG(i1,,is2,n1,n;2n)OG(i_{1},\ldots,i_{s-2},n-1,n;2n) which parametrizes two families of flags

Vi1Vis2Vn1VandVi1Vis2Vn1VV_{i_{1}}\subset\cdots\subset V_{i_{s-2}}\subset V_{n-1}\subset V~{}\text{and}~{}V_{i_{1}}\subset\cdots\subset V_{i_{s-2}}\subset V_{n-1}^{\prime}\subset V

where each Vit(1ts2)V_{i_{t}}(1\leq t\leq s-2) is an iti_{t}-dimensional isotropic subspace in VV, Vn1V_{n-1} and Vn1V_{n-1}^{\prime} are the (n1)(n-1)-dimensional isotropic subspaces in VV.

There are universal flags of subbundles

0=H0Hi1Hn1𝒪Dn/PI×V,\displaystyle 0=H_{0}\subset H_{i_{1}}\subset\cdots\subset H_{n-1}\subset\mathcal{O}_{D_{n}/P_{I}}\times V,
0=H0Hi1Hn1𝒪Dn/PI×V\displaystyle 0=H_{0}\subset H_{i_{1}}\subset\cdots\subset H_{n-1}^{\prime}\subset\mathcal{O}_{D_{n}/P_{I}}\times V

on Dn/PID_{n}/P_{I}, where rank Hit=it(1ts2)H_{i_{t}}=i_{t}~{}(1\leq t\leq s-2) and rankHn1=rankHn1=n1\text{rank}~{}H_{n-1}=\text{rank}~{}H_{n-1}^{\prime}=n-1.

Case 2. n1I,nIn-1\in I,n\notin I or n1I,nIn-1\notin I,n\in I. Dn/PID_{n}/P_{I} is one of the two irreducible components of the even Orthogonal flag manifold OG(i1,,is1,n;2n)(is=n)OG(i_{1},\ldots,i_{s-1},n;2n)~{}(i_{s}=n), which parametrizes flag

Vi1Vi2VnV,V_{i_{1}}\subset V_{i_{2}}\subset\cdots\subset V_{n}\subset V,

where each Vit(1ts)V_{i_{t}}(1\leq t\leq s) is an iti_{t}-dimensional isotropic subspace in VV.

There is a universal flag of subbundles

0=H0Hi1Hn𝒪Dn/PI×V0=H_{0}\subset H_{i_{1}}\subset\cdots\subset H_{n}\subset\mathcal{O}_{D_{n}/P_{I}}\times V

on Dn/PID_{n}/P_{I}, where rank Hit=it(1ts)H_{i_{t}}=i_{t}~{}(1\leq t\leq s).

Case 3. n1,nIn-1,n\notin I. Dn/PID_{n}/P_{I} is actually the even Orthogonal flag manifold OG(i1,,is1,is;2n)OG(i_{1},\ldots,i_{s-1},i_{s};2n), which parametrizes flag

Vi1Vi2VisV,V_{i_{1}}\subset V_{i_{2}}\subset\cdots\subset V_{i_{s}}\subset V,

where each Vit(1ts)V_{i_{t}}(1\leq t\leq s) is an iti_{t}-dimensional isotropic subspace in VV.

There is a universal flag of subbundles

0=H0Hi1HisHisHi1𝒪Dn/PI×V0=H_{0}\subset H_{i_{1}}\subset\cdots\subset H_{i_{s}}\subset H_{i_{s}}^{\bot}\subset\cdots\subset H_{i_{1}}^{\bot}\subset\mathcal{O}_{D_{n}/P_{I}}\times V

on Dn/PID_{n}/P_{I}, where HitH_{i_{t}}^{\bot} is the 𝒬\mathcal{Q}-orthogonal complement of HitH_{i_{t}}, rank Hit=itH_{i_{t}}=i_{t} and rank Hit=2nit(1ts)H_{i_{t}}^{\bot}=2n-i_{t}~{}(1\leq t\leq s).

(1) For k(1kn3)k~{}(1\leq k\leq n-3), the even Orthogonal Grassmannian Dn/Pk:=OG(k,2n)D_{n}/P_{k}:=OG(k,2n) parametrizes the kk-dimensional isotropic subspaces in VV. In this case, =Dn/Pk1,k+1\mathcal{M}=D_{n}/P_{k-1,k+1} and 𝒰=Dn/Pk1,k,k+1\mathcal{U}=D_{n}/P_{k-1,k,k+1}, i.e. we have the natural diagram:

(4.55)

and x\mathcal{M}_{x} is k1×Dnk/P1=k1×Q2(nk1)\mathbb{P}^{k-1}\times D_{n-k}/P_{1}=\mathbb{P}^{k-1}\times Q^{2(n-k-1)}.

Lemma 4.7.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)2nk3.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-k-3}.
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following two exact sequences of vector bundles:

0(Hk/Hk1)pHkqHk10,\displaystyle 0\rightarrow(H_{k}/H_{k-1})^{\vee}\rightarrow p^{\ast}H_{k}^{\vee}\rightarrow q^{\ast}H_{k-1}^{\vee}\rightarrow 0,
0Hk+1/HkHk+1/Hkq(Hk+1/Hk+1)0.\displaystyle 0\rightarrow H_{k+1}/H_{k}\rightarrow H_{k+1}^{\bot}/H_{k}\rightarrow q^{*}(H_{k+1}^{\bot}/H_{k+1})\rightarrow 0.

We will consider the following diagram

(4.62)

All the morphisms in the above diagram are projections. For any xXx\in X, the p1p_{1}-fiber p11(x)p_{1}^{-1}(x) is isomorphic to k1\mathbb{P}^{k-1} and the p2p_{2}-fiber p21(x)p_{2}^{-1}(x) is isomorphic to Q2(nk1)Q^{2(n-k-1)}. Note that for xXx\in X, the projection qq induces an isomorphism

q|p1(x):p1(x)x=k1×Q2(nk1).q|p^{-1}(x):p^{-1}(x)\rightarrow\mathcal{M}_{x}=\mathbb{P}^{k-1}\times Q^{2(n-k-1)}.

Hence p1(x)=p11(x)×p21(x)p^{-1}(x)=p_{1}^{-1}(x)\times p_{2}^{-1}(x). So we get

T𝒰/X=π1(T𝒰1/X)π2(T𝒰2/X).T_{\mathcal{U}/X}=\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\oplus\pi_{2}^{*}(T_{\mathcal{U}_{2}/X}).

We mimic the proof of Lemma 5.6 in 6 to conclude that

π1(T𝒰1/X)Hk/Hk1qHk1,π2(T𝒰2/X)(Hk+1/Hk)q(Hk+1/Hk+1).\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\cong H_{k}/H_{k-1}\otimes q^{\ast}H_{k-1}^{\vee},\pi_{2}^{*}(T_{\mathcal{U}_{2}/X})\cong(H_{k+1}/H_{k})^{\vee}\otimes q^{*}(H_{k+1}^{\bot}/H_{k+1}).

Therefore, T𝒰/X(Hk/Hk1qHk1)((Hk+1/Hk)q(Hk+1/Hk+1))T_{\mathcal{U}/X}\cong\big{(}H_{k}/H_{k-1}\otimes q^{\ast}H_{k-1}^{\vee}\big{)}\bigoplus\big{(}(H_{k+1}/H_{k})^{\vee}\otimes q^{*}(H_{k+1}^{\bot}/H_{k+1})\big{)} and

T𝒰/X|L~=𝒪L~(1)2nk3.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2n-k-3}.

(2) For k=n2k=n-2, the even Orthogonal Grassmannian Dn/Pn2:=OG(n2,2n)D_{n}/P_{n-2}:=OG(n-2,2n) parametrizes the n2n-2-dimensional isotropic subspaces in VV. In this case =Dn/Pn3,n1,n\mathcal{M}=D_{n}/P_{n-3,n-1,n} and 𝒰=Dn/Pn3,n2,n1,n\mathcal{U}=D_{n}/P_{n-3,n-2,n-1,n}. We have the natural diagram

(4.67)

and x\mathcal{M}_{x} is n3×1×1\mathbb{P}^{n-3}\times\mathbb{P}^{1}\times\mathbb{P}^{1}.

Lemma 4.8.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)n1.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus n-1}.
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following three exact sequences of vector bundles:

0(Hn2/Hn3)pHn2qHn30,\displaystyle 0\rightarrow(H_{n-2}/H_{n-3})^{\vee}\rightarrow p^{\ast}H_{n-2}^{\vee}\rightarrow q^{\ast}H_{n-3}^{\vee}\rightarrow 0,
0Hn1/Hn2𝒪𝒰n/Hn2q(𝒪n/Hn1)0.\displaystyle 0\rightarrow H_{n-1}/H_{n-2}\rightarrow\mathcal{O}_{\mathcal{U}}^{\oplus n}/H_{n-2}\rightarrow q^{*}(\mathcal{O}_{\mathcal{M}}^{\oplus n}/H_{n-1})\rightarrow 0.
0Hn1/Hn2𝒪𝒰n/Hn2q(𝒪n/Hn1)0.\displaystyle 0\rightarrow H_{n-1}^{\prime}/H_{n-2}\rightarrow\mathcal{O}_{\mathcal{U}}^{\oplus n}/H_{n-2}\rightarrow q^{*}(\mathcal{O}_{\mathcal{M}}^{\oplus n}/H_{n-1}^{\prime})\rightarrow 0.

We will consider the following diagram

(4.74)

All the morphisms in the above diagram are projections. For any xXx\in X, the p1p_{1}-fiber p11(x)p_{1}^{-1}(x) is isomorphic to n3\mathbb{P}^{n-3} and the p2p_{2}-fiber p21(x)p_{2}^{-1}(x) is isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. Note that for xXx\in X, the projection qq induces an isomorphism

q|p1(x):p1(x)x=n3×1×1.q|_{p^{-1}(x)}:p^{-1}(x)\rightarrow\mathcal{M}_{x}=\mathbb{P}^{n-3}\times\mathbb{P}^{1}\times\mathbb{P}^{1}.

Hence p1(x)=p11(x)×p21(x)p^{-1}(x)=p_{1}^{-1}(x)\times p_{2}^{-1}(x). So we get

T𝒰/X=π1(T𝒰1/X)π2(T𝒰2/X).T_{\mathcal{U}/X}=\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\oplus\pi_{2}^{*}(T_{\mathcal{U}_{2}/X}).

We mimic the proof of Lemma 5.6 in 6 to conclude that

π1(T𝒰1/X)Hn2/Hn3qHn3\pi_{1}^{*}(T_{\mathcal{U}_{1}/X})\cong H_{n-2}/H_{n-3}\otimes q^{\ast}H_{n-3}^{\vee}

and

π2(T𝒰2/X)(Hn1/Hn2)q(𝒪n/Hn1)(Hn1/Hn2)q(𝒪n/Hn1).\pi_{2}^{*}(T_{\mathcal{U}_{2}/X})\cong(H_{n-1}/H_{n-2})^{\vee}\otimes q^{*}(\mathcal{O}_{\mathcal{M}}^{\oplus n}/H_{n-1})\oplus(H_{n-1}^{\prime}/H_{n-2})^{\vee}\otimes q^{*}(\mathcal{O}_{\mathcal{M}}^{\oplus n}/H_{n-1}^{\prime}).

Therefore,

T𝒰/X(Hn2/Hn3qHn3)\displaystyle T_{\mathcal{U}/X}\cong\big{(}H_{n-2}/H_{n-3}\otimes q^{\ast}H_{n-3}^{\vee}\big{)}\bigoplus
((Hn1/Hn2)q(𝒪n/Hn1))((Hn1/Hn2)q(𝒪n/Hn1))\displaystyle\big{(}(H_{n-1}/H_{n-2})^{\vee}\otimes q^{*}(\mathcal{O}_{\mathcal{M}}^{\oplus n}/H_{n-1})\big{)}\bigoplus\big{(}(H_{n-1}^{\prime}/H_{n-2})^{\vee}\otimes q^{*}(\mathcal{O}_{\mathcal{M}}^{\oplus n}/H_{n-1}^{\prime})\big{)}

and T𝒰/X|L~=𝒪L~(1)n1.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus n-1}.

(3) For k=n1k=n-1 or nn, Dn/PkD_{n}/P_{k} is one of the two irreducible components of the even Orthogonal Grassmannian OG(n,2n)OG(n,2n). In this case =Dn/Pn2\mathcal{M}=D_{n}/P_{n-2} and 𝒰=Dn/Pn2,k\mathcal{U}=D_{n}/P_{n-2,k}. We have the natural diagram

(4.79)

and x\mathcal{M}_{x} is Grassmannian G(2,n)G(2,n).

Lemma 4.9.

Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. For the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, we have

T𝒰/X|L~=𝒪L~(1)2(n2).T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2(n-2)}.
Proof.

It’s not hard to check that over 𝒰\mathcal{U}, we have the following exact sequence of vector bundles

0(Hn/Hn2)pHnqHn20.0\rightarrow(H_{n}/H_{n-2})^{\vee}\rightarrow p^{\ast}H_{n}^{\vee}\rightarrow q^{\ast}H_{n-2}^{\vee}\rightarrow 0.

Restricting it to the pp-fiber p1(x)G(2,n)p^{-1}(x)\cong G(2,n) is just the Euler sequence on G(2,n)G(2,n). In fact, the projection pp identifies 𝒰\mathcal{U} in a canonical fashion with the Grassmannian bundle G(2,Hn)G(2,H_{n}) of XX. Hence T𝒰/XHn/Hn2qHn2T_{\mathcal{U}/X}\cong H_{n}/H_{n-2}\otimes q^{\ast}H_{n-2}^{\vee} and

T𝒰/X|L~=𝒪L~(1)2(n2).T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 2(n-2)}.

4.2 The exceptional simple Lie algebra

We say GG is of exceptional type if its Dynkin diagram is of type En(n=6,7,8)E_{n}~{}(n=6,7,8), F4F_{4} or G2G_{2}. Due to the complexity of the geometry of X=G/PkX=G/P_{k}, it is difficult to write down the specific form of the relative tangent bundle T𝒰/XT_{\mathcal{U}/X}, so here we use the method in 17 to calculate the splitting type of T𝒰/X|q1(l)T_{\mathcal{U}/X}|_{q^{-1}(l)}.

When GG is simple, the fundamental root system will be written as in the standard reference 5. In the case of En(n=6,7,8)E_{n}~{}(n=6,7,8) type, the mark of the nodes of the Dynkin diagam in the reference is different from our paper, so is the expression of the fundamental root system. So we write down the fundamental root system corresponding to the diagram of En(n=6,7,8)E_{n}~{}(n=6,7,8)-type in our paper as follow:

Let VV be a real vector space with dim V=8V=8 with orthogonal basis ei,i=1,,8e_{i},i=1,\ldots,8. Then the vectors

α1=\displaystyle\alpha_{1}= 12(e1++e8),α2=e6e7,\displaystyle-\frac{1}{2}(e_{1}+\cdots+e_{8}),~{}\alpha_{2}=e_{6}-e_{7},
α3=\displaystyle\alpha_{3}= e6+e7,α4=e5e6,\displaystyle e_{6}+e_{7},~{}\alpha_{4}=e_{5}-e_{6},
α5=\displaystyle\alpha_{5}= e4e5,α6=e3e4,\displaystyle e_{4}-e_{5},~{}\alpha_{6}=e_{3}-e_{4},
α7=\displaystyle\alpha_{7}= e2e3,α8=e1e2\displaystyle e_{2}-e_{3},~{}\alpha_{8}=e_{1}-e_{2}

form a fundamental root system of type E8E_{8}. Since E6,E7E_{6},E_{7} can be identified canonically with subsystem of E8E_{8}, so α1,,αn\alpha_{1},\ldots,\alpha_{n} form a fundamental root system of type En(n=6,7)E_{n}(n=6,7).

It’s well known that homogeneous vector bundles especially the tangent bundle on a rational homogeneous space G/PG/P are determined by representations of the Lie algebra 𝔭\mathfrak{p}. Restricting the relative tangent bundle T𝒰/XT_{\mathcal{U}/X} to a line L~=q1(l)\widetilde{L}=q^{-1}(l), we will get a flag bundle H/PN(k)H/P_{N(k)} and D(H)=D\kD(H)=D\backslash{k} with tag, where the tag is the set of intersection numbers of a minimal section of the associated H/PN(k)H/P_{N(k)}-flag bundle with the relative canonical divisors (see 22 Proposition 3.17). We refer to 22, 20 for a complete account on the tag of a flag bundle. Assume H/PN(k):=G1/PI1×G2/PI2××Gm/PImH/P_{N(k)}:=G_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}} (m3m\leq 3) where GiG_{i} is simple Lie group. Calculating weights that give the tangent bundle of Gi/PIi(1im)G_{i}/P_{I_{i}}~{}(1\leq i\leq m), which are all the positive roots of the Lie algebra of GiG_{i} that contain the root adjacent to αk\alpha_{k}. Putting together the above tag and weights, we will get the splitting type of T𝒰/X|L~T_{\mathcal{U}/X}|_{\widetilde{L}}. We refer to (20, page 5-7) for a complete account. By calculation, we have the tables (Table 3, 4, 5, 6) in Appendix.

Professor L. E. Solá Conde told us the following example for calculating the relative tangent bundle restricting to an isotropic line by personal correspondance.

Example 4.10.

Let X=E7/P7X=E_{7}/P_{7} and 𝒰=E7/P6,7\mathcal{U}=E_{7}/P_{6,7}. Then H/PN(7)=E6/P6H/P_{N(7)}=E_{6}/P_{6}. Let E7/PI:=E7/P{1,2,3,4,5,6,7}E_{7}/P_{I}:=E_{7}/P_{\{1,2,3,4,5,6,7\}}. The isotropic lines in E7/PiE_{7}/P_{i} are the image of the fibers of the 1\mathbb{P}^{1}-bundle

pi:=E7/PIE7/PI\i,(i=1,2,7)p^{i}:=E_{7}/P_{I}\rightarrow E_{7}/P_{I\backslash i},~{}(i=1,2,\ldots 7)

into E7/P7E_{7}/P_{7} via the natural map

πi:E7/PIE7/Pi.\pi^{i}:E_{7}/P_{I}\rightarrow E_{7}/P_{i}.

Let Ki(i=1,2,7)-K_{i}~{}(i=1,2,\ldots 7) be the relative anticanonical bundles of pip^{i} and CiC_{i} be a fiber of pip^{i}. When mapped to the varieties of Picard number one E7/PiE_{7}/P_{i}’s, CiC_{i}’s are the isotropic lines, where 1i71\leq i\leq 7. The C7C_{7} can be regarded as a minimal section of the fibration morphism E7/P6,7E7/P7E_{7}/P_{6,7}\rightarrow E_{7}/P_{7} when it restricts to π7(C7)\pi^{7}(C_{7}). By 21 Proposition 2.13, the matrix (KiCj)(-K_{i}C_{j}) is the Cartan matrix of E7E_{7}. The tag is exactly

(K1C7,,K6C7)=(0,0,0,0,0,1),(K_{1}C_{7},...,K_{6}C_{7})=(0,0,0,0,0,1),

which is the 77-th column of the Cartan matrix without the last term and with the signs changed. The weights are coefficients of all the positive roots of the Lie algebra of E6E_{6} in term of the linear combinations of the fundamental roots that contain the 66-th root as a summand; that is the roots of the form m1α1++m6α6m_{1}\alpha_{1}+\cdots+m_{6}\alpha_{6} with mi0,m6>0m_{i}\geq 0,~{}m_{6}>0, where αi\alpha_{i}’s are the fundamental roots (i=1,,6i=1,\ldots,6) (See Table 4 in Appendix for the values of mim_{i}). So

T𝒰/X|L~=𝒪L~(1)16.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 16}.

Combining Table 3, 4, 5, 6 we can get the following Proposition.

Proposition 4.11.

Suppose GG to be of exceptional type and X=G/PkX=G/P_{k}. Let L~=q1(l)𝒰\widetilde{L}=q^{-1}(l)\subset\mathcal{U} for ll\in\mathcal{M}. When X=F4/P3,F4/P4,G2/P1X=F_{4}/P_{3},F_{4}/P_{4},G_{2}/P_{1}, T𝒰/X|L~T_{\mathcal{U}/X}|_{\widetilde{L}} has the following form respectively:

𝒪L~(2)2𝒪L~(1),𝒪L~(2)3𝒪L~(1)3,𝒪L~(3).\mathcal{O}_{\widetilde{L}}(-2)^{\oplus 2}\oplus\mathcal{O}_{\widetilde{L}}(-1),\mathcal{O}_{\widetilde{L}}(-2)^{\oplus 3}\oplus\mathcal{O}_{\widetilde{L}}(-1)^{\oplus 3},\mathcal{O}_{\widetilde{L}}(-3).

Otherwise

T𝒰/X|L~=𝒪L~(1)N,whereN=dim𝒰dimX.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{N},~{}\text{where}~{}N=dim~{}\mathcal{U}-dimX.

Summing up, we have obtained the following

Proposition 4.12.

Let GG be a simple Lie group with the Dynkin diagram 𝒟\mathcal{D} and αk\alpha_{k} is a long root of 𝒟\mathcal{D}. Denote by XX the generalized Grassmannian by marking on 𝒟\mathcal{D} the node kk. Then for the relative tangent bundle T𝒰/XT_{\mathcal{U}/X} (Notations as Section 4), we have

T𝒰/X|L~=𝒪L~(1)N,whereN=dim𝒰dimX.T_{\mathcal{U}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{N},~{}\text{where}~{}N=dim~{}\mathcal{U}-dimX.
Proof.

It is obviously from Lemma 4.1,4.2,4.6,4.7,4.8,4.9 and Proposition 4.11. ∎

4.3 The generalization of the Grauert-Mu¨\ddot{\text{u}}lich-Barth theorem

The construction of subsheaves in holomorphic vector bundles plays an important role in the proof of the generalized Grauert-Mu¨\ddot{\text{u}}lich-Barth theorem. A generalization of the Grauert-Mu¨\ddot{\text{u}}lich-Barth theorem to normal projective varieties in characteristic zero is proved (see 13 Theorem 3.1.2). Since the theorem in the book is for any normal projective variety, the bound is pretty coarse. In this section, we will find the explicit bound for rational homogeneous spaces.

The following Descent Lemma provides a way for us to prove the existence of subsheaves.

Lemma 4.13.

(Descent Lemma 23) Let XX, YY be nonsingular varieties over kk, f:XYf:X\rightarrow Y be a surjective submersion with connected fibers and EE be an algebraic rr-bundle over YY. Let K~fE\widetilde{K}\subset f^{\ast}E be a subbundle of rank tt in fEf^{\ast}E and Q~=fE/K~\widetilde{Q}=f^{\ast}E/\widetilde{K} be its quotient. If

Hom(TX/Y,om(K~,Q~))=0,Hom(T_{X/Y},\mathcal{H}om(\widetilde{K},\widetilde{Q}))=0,

then K~\widetilde{K} is the form K~=fK\widetilde{K}=f^{\ast}K for some algebraic subbundle KEK\subset E of rank tt.

Follow the previous notations. Let GG be a simple Lie group with the Dynkin diagram 𝒟\mathcal{D} and αk\alpha_{k} is a root of 𝒟\mathcal{D}. Let’s consider the standard diagram associated to XX

(4.84)
Theorem 4.14.

Let XX be a generalized Grassmannian by marking on 𝒟\mathcal{D} the node kk and αk\alpha_{k} be a long root of 𝒟\mathcal{D}. Let EE be a holomorphic rr-bundle over XX of type a¯E=(a1,,ar),a1ar\underline{a}_{E}=(a_{1},\ldots,a_{r}),a_{1}\geq\cdots\geq a_{r}. If for some t<rt<r,

atat+12forsomet<r,a_{t}-a_{t+1}\geq 2~{}~{}for~{}some~{}t<r,

then there is a normal subsheaf KEK\subset E of rank tt with the following properties: over the open set VE=p(q1(UE))XV_{E}=p(q^{-1}(U_{E}))\subset X, where UEU_{E} is an open set in \mathcal{M}, the sheaf KK is a subbundle of EE, which on the line LXL\subset X given by lUEl\in U_{E} has the form

K|Lj=1t𝒪L(aj).K|_{L}\cong\oplus_{j=1}^{t}\mathcal{O}_{L}(a_{j}).
Proof.

The Proposition 4.12 and the Lemma 4.13 play important roles in our proof and our proof applies almost verbatim the proof of Theorem 2.1.4 given in 23. ∎

This theorem has far reaching consequences. We give first a series of immediate deductions.

Corollary 4.15.

Let XX be a generalized Grassmannian with long root αk\alpha_{k}. For a semistable rr-bundle EE over XX of type a¯E=(a1,,ar),a1ar\underline{a}_{E}=(a_{1},\ldots,a_{r}),a_{1}\geq\cdots\geq a_{r}. we have

aiai+11fori=1,,r1.a_{i}-a_{i+1}\leq 1~{}~{}for~{}i=1,\ldots,r-1.
Proof.

If for some t<rt<r, we had atat+12a_{t}-a_{t+1}\geq 2, then we could find a normal subsheaf KEK\subset E which is of the form

K|Lj=1t𝒪L(aj)K|_{L}\cong\oplus_{j=1}^{t}\mathcal{O}_{L}(a_{j})

over the general line LXL\subset X. Then we would have μ(E)<μ(K)\mu(E)<\mu(K) contrary to hypothesis. ∎

In particular, we get the generalization theorem of Grauert-Mu¨\ddot{\text{u}}lich-Barth :

Corollary 4.16.

Let XX be a generalized Grassmannian with long root αk\alpha_{k}. The splitting type of a semistable normalized 2-bundle EE over generalized Grassmannian XX is

aE={(0,0)ifc1(E)=0(0,1)ifc1(E)=1.a_{E}=\left\{\begin{array}[]{ll}(0,0)&if~{}c_{1}(E)=0\\ (0,-1)&if~{}c_{1}(E)=-1.\end{array}\right.
Corollary 4.17.

Let XX be a generalized Grassmannian with long root αk\alpha_{k}. For a uniform rr-bundle E(r=ς(X)+1)E~{}(r=\varsigma(X)+1) over XX of type (see Section 3 for the notation ς(X)\varsigma(X))

a¯E=(a1,,ar),a1ar,\underline{a}_{E}=(a_{1},\ldots,a_{r}),~{}a_{1}\geq\cdots\geq a_{r},

which does not split, we have

aiai+11fori=1,,r1.a_{i}-a_{i+1}\leq 1~{}~{}for~{}i=1,\ldots,r-1.
Proof.

If for some t<rt<r, we had atat+12a_{t}-a_{t+1}\geq 2, then we could find a uniform subbundle KEK\subset E of type which is of a¯K=(a1,,at)\underline{a}_{K}=(a_{1},\ldots,a_{t}) (because VE=XV_{E}=X). Then quotient bundle Q=E/KQ=E/K would be uniform of type (as+1,,ar)(a_{s+1},\ldots,a_{r}). According to Theorem 3.4 the bundle KK and QQ must be direct sums of line bundles. The exact sequence

0KEQ00\rightarrow K\rightarrow E\rightarrow Q\rightarrow 0

would therefore split and hence EE would be a direct sum of line bundles contrary to hypothesis. ∎

When αk\alpha_{k} is a short root of 𝒟\mathcal{D}, \mathcal{M} is not the variety of lines on XX, but only a closed GG-orbit. Therefore we’re just going to think about the splitting type and semistability of vector bundles with respect to \mathcal{M}.

Theorem 4.18.

Let XX be a generalized Grassmannian by marking on 𝒟\mathcal{D} the node kk and αk\alpha_{k} be a short root of 𝒟\mathcal{D}. Let EE be a holomorphic rr-bundle over XX of type a¯E=(a1,,ar),a1ar\underline{a}_{E}=(a_{1},\ldots,a_{r}),a_{1}\geq\cdots\geq a_{r} with respect to \mathcal{M}. If for some t<rt<r,

atat+1{3,if𝒟G24,if𝒟=G2,a_{t}-a_{t+1}\geq\left\{\begin{array}[]{ll}3,&if~{}\mathcal{D}\neq G_{2}\\ 4,&if~{}\mathcal{D}=G_{2},\end{array}\right.

then there is a normal subsheaf KEK\subset E of rank tt with the following properties: over the open set VE=p(q1(UE))XV_{E}=p(q^{-1}(U_{E}))\subset X, where UEU_{E} is an open set in \mathcal{M}, the sheaf KK is a subbundle of EE, which on the line LXL\subset X given by lUEl\in U_{E} has the form

K|Lj=1t𝒪L(aj).K|_{L}\cong\oplus_{j=1}^{t}\mathcal{O}_{L}(a_{j}).
Proof.

The Lemma 4.13, 4.3, 4.4, 4.5 and Proposition 4.11 play important roles in our proof and our proof applies almost verbatim the proof of Theorem 2.4 given in 23. ∎

Similarly, we have the following corollary.

Corollary 4.19.

Let XX be a generalized Grassmannian with short root αk\alpha_{k}. For a semistable rr-bundle EE over XX of type a¯E=(a1,,ar),a1ar\underline{a}_{E}=(a_{1},\ldots,a_{r}),a_{1}\geq\cdots\geq a_{r} with respect to \mathcal{M}. we have

aiai+13fori=1,,r1.a_{i}-a_{i+1}\leq 3~{}~{}for~{}i=1,\ldots,r-1.

In particular, if 𝒟G2\mathcal{D}\neq G_{2}, aiai+12fori=1,,r1.a_{i}-a_{i+1}\leq 2~{}~{}for~{}i=1,\ldots,r-1.

From now, Let

X=G/PG1/PI1×G2/PI2××Gm/PIm,X=G/P\simeq G_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}},

where GiG_{i} is a simple Lie group with Dynkin diagram 𝒟i\mathcal{D}_{i} whose set of nodes is DiD_{i} and PIiP_{I_{i}} is a parabolic subgroup of GiG_{i} corresponding to IiDiI_{i}\subset D_{i}. We set F(Ii):=Gi/PIiF(I_{i}):=G_{i}/P_{I_{i}} by marking on the Dynkin diagram 𝒟i\mathcal{D}_{i} of GiG_{i} the nodes corresponding to IiI_{i}. Let δi\delta_{i} be a node in 𝒟i\mathcal{D}_{i} and N(δi)N(\delta_{i}) be the set of nodes in 𝒟i\mathcal{D}_{i} that are connected to δi\delta_{i}.

If δiIi\delta_{i}\in I_{i}, we call

iδic:=Gi/Piδic×Gi/PIi^(1im),\mathcal{M}_{i}^{\delta_{i}^{c}}:=G_{i}/P_{i}^{\delta_{i}^{c}}\times\widehat{G_{i}/P_{I_{i}}}~{}(1\leq i\leq m),

the ii-th special family of lines of class δˇi\check{\delta}_{i}, where Piδic:=PIi\δiN(δi)P_{i}^{\delta_{i}^{c}}:=P_{I_{i}\backslash\delta_{i}\cup N(\delta_{i})} and Gi/PIi^\widehat{G_{i}/P_{I_{i}}} is G1/PI1×G2/PI2××Gm/PImG_{1}/P_{I_{1}}\times G_{2}/P_{I_{2}}\times\cdots\times G_{m}/P_{I_{m}} by deleting ii-th term Gi/PIiG_{i}/P_{I_{i}}. Denote by

𝒰iδic:=Gi/PIiN(δi)×Gi/PIi^\mathcal{U}_{i}^{\delta_{i}^{c}}:=G_{i}/P_{I_{i}\cup N(\delta_{i})}\times\widehat{G_{i}/P_{I_{i}}}

the ii-th universal family of class δˇi\check{\delta}_{i}, which has a natural 1\mathbb{P}^{1}-bundle structure over iδic\mathcal{M}_{i}^{\delta_{i}^{c}}.

We separate our discussion into two cases:

Case I: N(δi)IiN(\delta_{i})\subseteq I_{i}, then 𝒰iδic=X\mathcal{U}_{i}^{\delta_{i}^{c}}=X and we have the natural projection XiδicX\rightarrow\mathcal{M}_{i}^{\delta_{i}^{c}};

Case II: N(δi)IiN(\delta_{i})\nsubseteq I_{i}, then we have the standard diagram:

(4.89)

Notice that for xXx\in X, xδic=q2(q11(x))\mathcal{M}^{\delta_{i}^{c}}_{x}={q_{2}}({q_{1}}^{-1}(x)) coincides with H/PN(δi)H/P_{N(\delta_{i})} where D(H)D(H) is the components of (Di\Ii¯)\δi(\overline{D_{i}\backslash I_{i}})\backslash\delta_{i} containing an element of N(δi)N(\delta_{i}) by Theorem 2.6.

Let 𝒢δi\mathcal{G}^{\delta_{i}} be a generalized Grassmannian whose Dynkin diagram 𝒟δi\mathcal{D}^{\delta_{i}} is the maximal sub-diagram of (𝒟i,Ii)(\mathcal{D}_{i},I_{i}) with the only marked point δi\delta_{i}. Let’s consider the standard diagram associated to 𝒢δi\mathcal{G}^{\delta_{i}}

(4.94)

It’s not hard to see that xδic\mathcal{M}^{\delta_{i}^{c}}_{x} is isomorphic to y=q(p1(y))\mathcal{M}_{y}=q(p^{-1}(y)) for every y𝒢δiy\in\mathcal{G}^{\delta_{i}} and we have the following Lemma.

Lemma 4.20.

Let L~=q21(l)𝒰iδic\widetilde{L}=q_{2}^{-1}(l)\subset\mathcal{U}_{i}^{\delta_{i}^{c}} for liδicl\in\mathcal{M}_{i}^{\delta_{i}^{c}}. For the relative tangent bundle T𝒰iδic/XT_{\mathcal{U}_{i}^{\delta_{i}^{c}}/X}, we have

T𝒰iδic/X|L~=T𝒰/𝒢δi|L~.T_{\mathcal{U}_{i}^{\delta_{i}^{c}}/X}|_{\widetilde{L}}=T_{\mathcal{U}/\mathcal{G}^{\delta_{i}}}|_{\widetilde{L}}.

If δi\delta_{i} is an exposed short root of (𝒟i,Ii)(\mathcal{D}_{i},I_{i}), then δi\delta_{i} is a short root of 𝒟δi\mathcal{D}^{\delta_{i}} and vice versa. Combining this fact and the above Lemma, we get the following Proposition:

Proposition 4.21.

Let L~=q21(l)𝒰iδic\widetilde{L}=q_{2}^{-1}(l)\subset\mathcal{U}_{i}^{\delta_{i}^{c}} for liδicl\in\mathcal{M}_{i}^{\delta_{i}^{c}}. If δi\delta_{i} is not an exposed short root, then for the relative tangent bundle T𝒰iδic/XT_{\mathcal{U}_{i}^{\delta_{i}^{c}}/X}, we have

T𝒰iδic/X|L~=𝒪L~(1)N,N=dim𝒰iδicdimX.T_{\mathcal{U}_{i}^{\delta_{i}^{c}}/X}|_{\widetilde{L}}=\mathcal{O}_{\widetilde{L}}(-1)^{N},~{}N=dim~{}\mathcal{U}_{i}^{\delta_{i}^{c}}-dimX.
Proof.

It’s obviously from Proposition 4.12. ∎

Similarly, by combining with Lemma 4.3, 4.4, 4.5, 4.20 and Proposition 4.11, we can immediately draw the following proposition.

Proposition 4.22.

Let L~=q21(l)𝒰iδic\widetilde{L}=q_{2}^{-1}(l)\subset\mathcal{U}_{i}^{\delta_{i}^{c}} for liδicl\in\mathcal{M}_{i}^{\delta_{i}^{c}}. If δi\delta_{i} is an exposed short root, then for the relative tangent bundle T𝒰iδic/XT_{\mathcal{U}_{i}^{\delta_{i}^{c}}/X}, the splitting type of T𝒰iδic/XT_{\mathcal{U}_{i}^{\delta_{i}^{c}}/X} takes several forms: (1,,1,2)(-1,\ldots,-1,-2), (2,,2)(-2,\ldots,-2), (2,2,1)(-2,-2,-1), (2,2,2,1,1,1)(-2,-2,-2,-1,-1,-1) or (3)(-3).

According to Theorem 2.3, 2.6, if δi\delta_{i} is not an exposed short root, then iδic\mathcal{M}_{i}^{\delta_{i}^{c}} is the space of of lines of class δˇi\check{\delta}_{i} and xδic\mathcal{M}_{x}^{\delta_{i}^{c}} is the space of lines of class δˇi\check{\delta}_{i} through xx.

Using the above propositions, we have the similar results for rational homogeneous spaces.

Theorem 4.23.

Fix δiIi\delta_{i}\in I_{i} and assume that δi\delta_{i} is not an exposed short root. Let EE be a holomorphic rr-bundle over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),~{}a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. If for some t<rt<r,

at(δi)at+1(δi){1,andN(δi)fits Case I2,andN(δi)fits Case II,a_{t}^{(\delta_{i})}-a_{t+1}^{(\delta_{i})}\geq\left\{\begin{array}[]{ll}1,&and~{}N(\delta_{i})~{}\text{fits Case I}\\ 2,&and~{}N(\delta_{i})~{}\text{fits Case II},\end{array}\right.

then there is a normal subsheaf KEK\subset E of rank tt with the following properties: over the open set VE=q1(q21(UE(δi)))XV_{E}=q_{1}(q_{2}^{-1}(U_{E}^{(\delta_{i})}))\subset X, where UE(δi)U_{E}^{(\delta_{i})} is an open set in δic\mathcal{M}^{\delta_{i}^{c}}, the sheaf KK is a subbundle of EE, which on the line LXL\subset X given by lUE(δi)l\in U_{E}^{(\delta_{i})} has the form

K|Ls=1t𝒪L(as(δi)).K|_{L}\cong\oplus_{s=1}^{t}\mathcal{O}_{L}(a_{s}^{(\delta_{i})}).
Proof.

The Proposition 4.21 and the Lemma 4.13 play important roles in our proof and our proof applies almost verbatim the proof of Theorem 5.7 given in 6. ∎

Corollary 4.24.

With the same assumption as Theorem 4.23. For a δi\delta_{i}-semistable rr-bundle EE over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}, we have

as(δi)as+1(δi)1for alls=1,,r1.a_{s}^{(\delta_{i})}-a_{s+1}^{(\delta_{i})}\leq 1~{}~{}\text{for all}~{}s=1,\ldots,r-1.

In particular, if N(δi)N(\delta_{i}) fits Case I, then we have as(δi)a_{s}^{(\delta_{i})}’s are constant for all 1sr1\leq s\leq r.

When αj\alpha_{j} is an exposed short root of (𝒟,I)(\mathcal{D},I), iδic\mathcal{M}_{i}^{\delta_{i}^{c}} is not the space of of lines of class δˇi\check{\delta}_{i}, but only a closed GG-orbit.

Theorem 4.25.

Fix δiIi\delta_{i}\in I_{i} and assume that δi\delta_{i} is an exposed short root. Let EE be a holomorphic rr-bundle over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),~{}a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}. If for some t<rt<r,

at(δi)at+1(δi){1,andN(δi)fits Case I4,andN(δi)fits Case II,a_{t}^{(\delta_{i})}-a_{t+1}^{(\delta_{i})}\geq\left\{\begin{array}[]{ll}1,&and~{}N(\delta_{i})~{}\text{fits Case I}\\ 4,&and~{}N(\delta_{i})~{}\text{fits Case II},\end{array}\right.

then there is a normal subsheaf KEK\subset E of rank tt with the following properties: over the open set VE=q1(q21(UE(δi)))XV_{E}=q_{1}(q_{2}^{-1}(U_{E}^{(\delta_{i})}))\subset X, where UE(δi)U_{E}^{(\delta_{i})} is an open set in δic\mathcal{M}^{\delta_{i}^{c}}, the sheaf KK is a subbundle of EE, which on the line LXL\subset X given by lUE(δi)l\in U_{E}^{(\delta_{i})} has the form

K|Ls=1t𝒪L(as(δi)).K|_{L}\cong\oplus_{s=1}^{t}\mathcal{O}_{L}(a_{s}^{(\delta_{i})}).
Proof.

The Proposition 4.22 and the Lemma 4.13 play important roles in our proof and our proof applies almost verbatim the proof of Theorem 5.7 given in 6. ∎

Corollary 4.26.

For a δi\delta_{i}-semistable rr-bundle EE over XX of type a¯E(δi)=(a1(δi),,ar(δi)),a1(δi)ar(δi)\underline{a}_{E}^{(\delta_{i})}=(a_{1}^{(\delta_{i})},\ldots,a_{r}^{(\delta_{i})}),a_{1}^{(\delta_{i})}\geq\cdots\geq a_{r}^{(\delta_{i})} with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}}, we have

as(δi)as+1(δi)3for alls=1,,r1.a_{s}^{(\delta_{i})}-a_{s+1}^{(\delta_{i})}\leq 3~{}~{}\text{for all}~{}s=1,\ldots,r-1.

In particular, if N(δi)N(\delta_{i}) fits Case I, then we have as(δi)a_{s}^{(\delta_{i})}’s are constant for all 1sr1\leq s\leq r.

From Proposition 3.12, Corollary 4.24 and Corollary 4.26, we can have the following result.

Corollary 4.27.

Let X=G/BX=G/B, where GG is a semi-simple Lie group and BB is a Borel subgroup of GG. If an rr-bundle EE is δi\delta_{i}-semistable for all ii and δi(1im)\delta_{i}~{}(1\leq i\leq m) over XX, then EE splits as a direct sum of line bundles.

Proof.

The assumption tells us that EE is poly-uniform with respect to iδic\mathcal{M}_{i}^{\delta_{i}^{c}} for each ii and δi\delta_{i} and the splitting type is (a(δi),,a(δi))(a^{(\delta_{i})},\ldots,a^{(\delta_{i})}) by Corollary 4.24 and Corollary 4.26. We therefore conclude by Proposition 3.12. ∎

Acknowledgements

We would like to show our great appreciation to professor B. Fu, L. Manivel, R. Muñoz, G. Occhetta, L. E. Solá Conde and K. Watanabe for useful discussions. In particular, we would like to thank professor L. E. Solá Conde for recommending us their interesting papers and explaining the details in their papers with patience, especially for the method of calculating the relative tangent bundles between two generalized flag manifolds.

5 Appendix

Table 3: E6E_{6}-type
GG node kk H/PN(k)H/P_{N(k)} tag weights
E6E_{6} 11 D5/P5D_{5}/P_{5} (00001)(00001)
{(12211),(11211),(01211),(11111),(11101),\{(12211),(11211),(01211),(11111),(11101),
(01111),(01101),(00111),(00101),(00001)}(01111),(01101),(00111),(00101),(00001)\}
E6E_{6} 22 A5/P3A_{5}/P_{3} (00100)(00100)
{(11100),(11110),(11111),(01100),(01110),\{(11100),(11110),(11111),(01100),(01110),
(01111),(00100),(00110),(00111)}(01111),(00100),(00110),(00111)\}
E6E_{6} 33 A1/P1×A4/P2A_{1}/P_{1}\times A_{4}/P_{2} (1,0100)(1,0100)
{(1)}×{(1100),(1110),(1111),\{(1)\}\times\{(1100),(1110),(1111),
(0100),(0110),(0111)}(0100),(0110),(0111)\}
E6E_{6} 44 A2/P2×A1/P1×A2/P1A_{2}/P_{2}\times A_{1}/P_{1}\times A_{2}/P_{1} (01,1,10)(01,1,10) {(01),(11)}×{(1)}×{(10),(11)}\{(01),(11)\}\times\{(1)\}\times\{(10),(11)\}
E6E_{6} 55 A4/P3×A1/P1A_{4}/P_{3}\times A_{1}/P_{1} (0010,1)(0010,1)
{(1110),(1111),(0110),\{(1110),(1111),(0110),
(0111),(0010),(0011)}×{(1)}(0111),(0010),(0011)\}\times\{(1)\}
E6E_{6} 66 D5/P5D_{5}/P_{5} (00001)(00001)
{(12211),(11211),(01211),(11111),(11101),\{(12211),(11211),(01211),(11111),(11101),
(01111),(01101),(00111),(00101),(00001)}(01111),(01101),(00111),(00101),(00001)\}
Table 4: E7E_{7}-type
GG node kk H/PN(k)H/P_{N(k)} tag weights
E7E_{7} 11 D6/P6D_{6}/P_{6} (000001)(000001)
{(122211),(112211),(012211),(111111),\{(122211),(112211),(012211),(111111),
(111101),(011111),(001101),(000111),(111101),(011111),(001101),(000111),
(000101),(000001),(111211),(011211),(000101),(000001),(111211),(011211),
(011101),(001211),(001111)}(011101),(001211),(001111)\}
E7E_{7} 22 A6/P3A_{6}/P_{3} (001000)(001000)
{(111000),(111100),(111110),(111111),\{(111000),(111100),(111110),(111111),
(011000),(011100),(011110),(011111),(011000),(011100),(011110),(011111),
(001000),(001100),(001110),(001111)}(001000),(001100),(001110),(001111)\}
E7E_{7} 33 A1/P1×A5/P2A_{1}/P_{1}\times A_{5}/P_{2} (1,01000)(1,01000)
{(1)}×{(11000),(11100),\{(1)\}\times\{(11000),(11100),
(11110),(11111),(01000),(11110),(11111),(01000),
(01100),(01110),(01111)}(01100),(01110),(01111)\}
E7E_{7} 44 A2/P2×A1/P1×A3/P1A_{2}/P_{2}\times A_{1}/P_{1}\times A_{3}/P_{1} (01,1,100)(01,1,100) {(01),(11)}×{(1)}×{(100),(110),(111)}\{(01),(11)\}\times\{(1)\}\times\{(100),(110),(111)\}
E7E_{7} 55 A4/P3×A2/P1A_{4}/P_{3}\times A_{2}/P_{1} (0010,10)(0010,10)
{(1110),(1111),(0110),(0111)\{(1110),(1111),(0110),(0111)
(0010),(0011)}×{(10),(11)}(0010),(0011)\}\times\{(10),(11)\}
E7E_{7} 66 D5/P5×A1/P1D_{5}/P_{5}\times A_{1}/P_{1} (00001,1)(00001,1)
{(12211),(11211),(01211),(11111),\{(12211),(11211),(01211),(11111),
(11101),(01111),(01101),(00111),(11101),(01111),(01101),(00111),
(00101),(00001)}×{(1)}(00101),(00001)\}\times\{(1)\}
E7E_{7} 77 E6/P6E_{6}/P_{6} (000001)(000001)
{(000001),(000011),(000111),(010111),\{(000001),(000011),(000111),(010111),
(001111),(101111),(011111),(111111)(001111),(101111),(011111),(111111)
(011211),(111211),(011221),(112211),(011211),(111211),(011221),(112211),
(111211),(112221),(112321),(122321)}(111211),(112221),(112321),(122321)\}
Table 5: E8E_{8}-type
GG node kk H/PN(k)H/P_{N(k)} tag weights
E8E_{8} 11 D7/P7D_{7}/P_{7} (0000001)(0000001)
{(1222211),(1122211),(1112211),\{(1222211),(1122211),(1112211),
(1111211),(1111111),(1111101),(1111211),(1111111),(1111101),
(0122211),(0112211),(0111211),(0122211),(0112211),(0111211),
(0111111),(0111101),(0012211),(0111111),(0111101),(0012211),
(0011211),(0011111),(0011101),(0011211),(0011111),(0011101),
(0001211),(0001111),(0001101),(0001211),(0001111),(0001101),
(0000111),(0000101),(0000001)}(0000111),(0000101),(0000001)\}
E8E_{8} 22 A7/P3A_{7}/P_{3} (0010000)(0010000)
{(1110000),(1111000),(1111110),\{(1110000),(1111000),(1111110),
(1111110),(1111111),(0110000),(1111110),(1111111),(0110000),
(0111000),(0111100),(0111110),(0111000),(0111100),(0111110),
(0111111),(0010000),(0011000),(0111111),(0010000),(0011000),
(0011100),(0011110),(0011111)}(0011100),(0011110),(0011111)\}
E8E_{8} 33 A1/P1×A6/P2A_{1}/P_{1}\times A_{6}/P_{2} (1,010000)(1,010000)
{(1)}×{(110000),(111000),\{(1)\}\times\{(110000),(111000),
(111100),(111110),(111111),(111100),(111110),(111111),
(010000),(011000),(011100),(010000),(011000),(011100),
(011110),(011111)}(011110),(011111)\}
E8E_{8} 44 A2/P2×A1/P1×A4/P1A_{2}/P_{2}\times A_{1}/P_{1}\times A_{4}/P_{1} (01,1,1000)(01,1,1000)
{(01),(11)}×{(1)}×{(1000),\{(01),(11)\}\times\{(1)\}\times\{(1000),
(1100),(1110),(1111)}(1100),(1110),(1111)\}
E8E_{8} 55 A4/P3×A3/P1A_{4}/P_{3}\times A_{3}/P_{1} (0010,100)(0010,100)
{(1110),(1111),(0110),(0111),\{(1110),(1111),(0110),(0111),
(0010),(0011)}×{(100),(110),(111)}(0010),(0011)\}\times\{(100),(110),(111)\}
E8E_{8} 66 D5/P5×A2/P1D_{5}/P_{5}\times A_{2}/P_{1} (00001,10)(00001,10)
{(12211),(11211),(01211),\{(12211),(11211),(01211),
(11111),(11101),(01111),(11111),(11101),(01111),
(01101),(00111),(00101),(01101),(00111),(00101),
(00001)}×{(10),(11)}(00001)\}\times\{(10),(11)\}
E8E_{8} 77 E6/P6×A1/P1E_{6}/P_{6}\times A_{1}/P_{1} (000001,1)(000001,1)
{(000001),(000011),(000111),\{(000001),(000011),(000111),
(010111),(001111),(101111),(010111),(001111),(101111),
(011111),(111111),(011211),(011111),(111111),(011211),
(111211),(011221),(112211),(111211),(011221),(112211),
(111211),(112221),(112321),(111211),(112221),(112321),
(122321)}×{(1)}(122321)\}\times\{(1)\}
E8E_{8} 88 E7/P7E_{7}/P_{7} (0000001)(0000001)
{(000001),(000011),(000111),\{(000001),(000011),(000111),
(0101111),(0001111),(0112221),(0101111),(0001111),(0112221),
(0112211),(0112111)(0111111),(0112211),(0112111)(0111111),
(0011111),(1112221),(1112211),(0011111),(1112221),(1112211),
(1112111),(1111111),(1011111),(1112111),(1111111),(1011111),
(1224321),(1223321),(1123321),(1224321),(1223321),(1123321),
(1223221),(1123221),(1122221),(1223221),(1123221),(1122221),
(1223211),(1123211),(1122211),(1223211),(1123211),(1122211),
(1122111),(1234321),(2234321)}(1122111),(1234321),(2234321)\}
Table 6: F4,G2F_{4},G_{2}-type
GG node kk H/PN(k)H/P_{N(k)} tag weights
F4F_{4} 11 C3/P3C_{3}/P_{3} (001)(001) {(121),(111),(011),(221),(021),(001)}\{(121),(111),(011),(221),(021),(001)\}
F4F_{4} 22 A1/P1×A2/P1A_{1}/P_{1}\times A_{2}/P_{1} (1,10)(1,10) {(1)}×{(10),(11)}\{(1)\}\times\{(10),(11)\}
F4F_{4} 33 A2/P2×A1/P1A_{2}/P_{2}\times A_{1}/P_{1} (02,1)(02,1) {(01),(11)}×{(1)}\{(01),(11)\}\times\{(1)\}
F4F_{4} 44 B3/P3B_{3}/P_{3} (001)(001) {(122),(112),(012),(111),(011),(001)}\{(122),(112),(012),(111),(011),(001)\}
G2G_{2} 11 A1/P1A_{1}/P_{1} (3)(3) {(1)}\{(1)\}
G2G_{2} 22 A1/P1A_{1}/P_{1} (1)(1) {(1)}\{(1)\}

References

  • [1] M. Andreatta and J. Wiśniewski. On manifolds whose tangent bundle contains an ample locally free subsheaf. Invent. Math., 146:209–217, 2001.
  • [2] E. Ballico. Uniform vector bundles on quadrics. Ann. Univ. Ferrara Sez. VII (N.S.), 27(1):135–146, 1982.
  • [3] E. Ballico. Uniform vector bundles of rank (n+1)(n+1) on n\mathbb{P}^{n}. Tsukuba J. Math., 7(2):215–226, 1983.
  • [4] A. Borel and R. Remmert. Über kompakte homogene kählersche mannigfaltigkeiten. Math. Ann., 145:429–439, 1961/1962.
  • [5] R. Carter. Lie Algebras of Finite and Affine Type. Cambridge studies in advanced mathematics, 2005.
  • [6] R. Du, X. Fang, and Y. Gao. Vector bundles on flag varieties. arXiv:1905.10151.
  • [7] H. Duan and X. Zhao. The chow ring of generalized grassmannians. Foundations of Computational Mathematics, 10(3):245–274, 2005.
  • [8] D. Eisenbud and J. Harris. 3264 All that-Interesection Theory in Algebraic Geometry. Cambridge University Press, 2011.
  • [9] G. Elencwajg. Les fibrés uniformes de rang 3 sur 2()\mathbb{P}^{2}(\mathbb{C}) sont homogénes. Math.Ann, 231:217–227, 1978.
  • [10] G. Elencwajg, A. Hirschowitz, and M. Schneider. Les fibrés uniformes de rang n sur n()\mathbb{P}^{n}(\mathbb{C}) sont ceux qu’on croit. Progr. Math., 7:37–63, 1980.
  • [11] P. Ellia. Sur les fibrés uniformes de rang (n+1)(n+1) sur n\mathbb{P}^{n}. Mém. Soc. Math. France (N.S.), 7:1–60, 1982.
  • [12] M. Guyot. Caractérisation par l’uniformité des fibrés universels sur la grassmanienne. Math. Ann., 270(1):47–62, 1985.
  • [13] D. Huybrechts and M. Lehn. The geometry of moduli spaces of sheaves, second ed. Cambridge University Press, 2010. xviii+325 pp.
  • [14] J-M. Hwang. Geometry of minimal rational curves on fano manifolds. School on Vanishing Theorems and Effective Results in Algebraic Geometry (Trieste, 2000), ICTP Lect. Notes, vol. 6, Abdus Salam Int. Cent. Theoret. Phys., Trieste, 2001, pages 335–393, 2001.
  • [15] Y. Kachi and E. Sato. Segre’s reflexivity and an inductive characterization of hyperquadrics. Mem. Amer. Math. Soc., 160(763), 2002.
  • [16] J.M. Landsberg and L. Manivel. On the projective geometry of rational homogeneous varieties. Comment. Math. Helv., 78:65–100, 2003.
  • [17] R. Muñoz, G. Occhetta, and L. E. Solá Conde. On uniform flag bundles on fano manifolds. arXiv:1610.05930.
  • [18] R. Muñoz, G. Occhetta, and L. E. Solá Conde. Uniform vector bundles on fano manifolds and applications. J. Reine Angew. Math. (Crelles Journal), 664:141–162, 2012.
  • [19] R. Muñoz, G. Occhetta, and L. E. Solá Conde. Splitting conjectures for uniform flag bundles. European Journal of Mathematics, 6:430–452, 2020.
  • [20] R. Muñoz, G. Occhetta, L. E. Solá Conde, and K. Watanabe. Rational curves,dynkin diagrams and fano manifolds with nef tangent bundle. Math.Ann., 361:583–609, 2015.
  • [21] G. Occhetta, L. E. Solá Conde, K. Watanabe, and J. A. Wiśniewski. Fano manifolds whose elementary contractions are smooth 1\mathbb{P}^{1}-fibrations: a geometric characterization of flag varieties. Ann. Sc. Norm. Super. Pica. Cl. Sci.
  • [22] G. Occhetta, L. E. Solá Conde, and J. A. Wiśniewski. Flag bundles on fano manifolds. Journal de Mathmatiques Pures et Appliques, 106(4):651–669, 2016.
  • [23] C. Okonek, M. Schneider, and H. Spindler. Vector bundles on complex projective spaces. Birkha¨\ddot{a}user/Springer Basel AG, Basel, 2011. viii+239 pp.
  • [24] E. Sato. Uniform vector bundles on a projective space. J. Math. Soc. Japan, 28(1):123–132, 1976.
  • [25] R. L. E. Schwarzenberger. Vector bundles on the projective plane. Proc. London Math. Soc., 11(3):623–640, 1961.
  • [26] H. Tamvakis. Quantum cohomology of isotropic grassmannians. Geometric Methods in Algebra and Number Theory, 235:311–338, 2006.
  • [27] A. Van de Ven. On uniform vector bundles. Math. Ann., 195:245–248, 1972.