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

Geometric Invariants of Recursive Group Orbit Stratification

Xiping Zhang xzhmath@gmail.com SCMS, Fudan University, Shanghai, China
(Date: 05/07/2021)
Abstract.

The local Euler obstructions and the Euler characteristics of linear sections with all hyperplanes on a stratified projective variety are key geometric invariants in the study of singularity theory. Despite their importance, in general it is very hard to compute them. In this paper we consider a special type of singularity: the recursive group orbits. They are the group orbits of a sequence of GnG_{n} representations VnV_{n} satisfying certain assumptions. We introduce a new intrinsic invariant called the csmc_{sm} invariant, and use it to give explicit formulas to the local Euler obstructions and the sectional Euler characteristics of such orbits. In particular, the matrix rank loci are examples of recursive group orbits. Thus as applications, we explicitly compute these geometry invariants for ordinary, skew-symmetric and symmetric rank loci. Our method is systematic and algebraic, thus works for algebraically closed field of characteristic 0. Moreover, in the complex setting we also compute the stalk Euler characteristics of the Intersection Cohomology Sheaf complexes for all three types of rank loci.

MSC Classification: 14C17, 14J17, 32S05, 55S35

1. Introduction

The geometric invariants on singular varieties have been an important subject for us to understand the singularities. In 1973 MacPherson introduced a local measurement for complex varieties (later published as [31]), and named it the local Euler obstruction. Defined as the obstruction to extend the distance 11-form after lifting to the Nash transform; it is the key ingredient in MacPherson’s proof of the existence and uniqueness of Chern class on singular spaces. An equivalent definition was given by Brasselet and Schwartz in [6] using vector fields. Later González-Sprinberg showed in [18] that there is an algebraic formula for the local Euler obstruction function, thus this definition extends to arbitrary algebraically closed field. For more about local Euler obstructions we refer to [5] and [8].

In the same year M. Kashiwara published his paper introducing the famous index theorem for holonomic D-modules [26]. In the paper he defined certain local topological invariants, as the weighted sum of the Euler characteristics of certain link spaces. He named them local characteristics. Although the two definitions were defined by two different flavors, and were introduced in two different branches of mathematics, it was proved in [10] that surprisingly the two definitions are equivalent. As the key ingredients in both singularity theory and Kashiwara’s index theorem, the local Euler obstructions of stratified spaces have been intensively studied in many different fields. It is one of the most important invariants in singularity theory.

However, it is very hard to compute the local Euler obstructions in general. Many authors have been working on formulas that make the computation easier. For example, González-Sprinberg’s formula allows one to use intersection theory method; the formula in [38] reduced the computation to the knowledge of local polar multiplicities. In [7] the authors provided a recursive formula using the Euler characteristics of the link spaces of strata. Recently in [24] the authors proved the relation between the local Euler obstructions and the maximal likelihood degree, and proposed an algorithm to compute local Euler obstructions by computer. Despite the difficulty, in many cases we have known the Euler obstructions very well. For Schubert cells in Grassmannnians they were computed in [9][25]; for skew-symmetric and ordinary rank loci they were computed in [35] and [17][39]. Recently in [32] the authors provide an algorithm for local Euler obstructions of Schubert varieties in all cominuscule spaces. Based on the examples computed Mihalcea and Singh conjectured [32, Conjecture 10.2] the non-negativity of the local Euler obstructions.

Another fundamental geometric invariant is the Euler characteristic. Over \mathbb{C} it’s simply the topological Euler characteristic, over arbitrary field of characteristic 0 it can be defined via the integration of MacPherson’s Chern class. When XX is a projective variety, we can access more refined invariants by consider linear sections, i.e., the Euler characteristic of XX intersects with a given hyperplane HH. We call them sectional Euler characteristics. When HH is generic, such invariants can be obtained from MacPherson’s Chern class (Cf. [2]). However, for ‘special’ linear sections, the geometry of the intersection is determined by both the singularity of XX and the position of HH, and concrete formulas are unknown in general. Also, it’s quite subtle to determine when a hypersurface is generic. These sectional Euler characteristics are very important invariants of XX. For example, when XX is a hypersurface, the generic sectional Euler characteristics of the complement correspond to another fundamental invariant of XX: the Milnor numbers of XX (Cf. [23][12]). Also, from computational algebraic geometry perspective, many algorithms involves the step of ‘choose a generic hyperplane section’, where one needs to compute the invariants on a generic linear section. For example, in [24] the authors prove a formula of local Euler obstructions in terms of the Maximal likelihood degrees of generic linear sections. As the case of local Euler obstructions, the sectional Euler characteristics are important but very hard to compute in general .

In this paper we consider a special type of singularities: the recursive group orbits over algebraically closed field of characteristic 0. By recursive group orbits we mean a sequence of group actions GnGL(Vn)G_{n}\to GL(V_{n}) satisfying certain conditions (Assumption 1). For such group orbits, we define an intrinsic invariant called the csmc_{sm} invariants, using the Chern classes of the projectivized orbits. They are discussed in §3. The main result in §3 is Theorem 3.1:

Theorem.

Assume that we have sequence of group actions as in Assumption 1. For any 1kn11\leq k\leq n-1 we denote Sn,k=(𝒪n,k)(Vn)S_{n,k}=\mathbb{P}(\mathcal{O}_{n,k})\subset\mathbb{P}(V_{n}) to be the projectivizations. Denote the csmc_{sm} invariants of 𝒪n,k\mathcal{O}_{n,k} by Smn,i{Sm}_{n,i}. We have the following formula for the local Euler obstructions :

Eu𝒪¯n,k(𝒪n,r)=Eu𝒪r,k(0)=(μ1>μ2>>μlk)Smnμ1Smμ1μ2SmμlkEu_{\bar{\mathcal{O}}_{n,k}}(\mathcal{O}_{n,r})=Eu_{\mathcal{O}_{r,k}}(0)=\sum_{(\mu_{1}>\mu_{2}>\cdots>\mu_{l}\geq k)}{Sm}_{n\mu_{1}}\cdot{Sm}_{\mu_{1}\mu_{2}}\cdots\cdot{Sm}_{\mu_{l}k}

Here the sum is over all (partial and full) flags (μ1>μ2>>μl)(\mu_{1}>\mu_{2}>\cdots>\mu_{l}) such that μi1=μi+1\mu_{i}-1=\mu_{i+1} for every ii and μlk\mu_{l}\geq k.

The formula requires only the knowledge of our csmc_{sm} invariants, which in many cases can be directly computed. Our result is purely algebraic and works for general field kk. In particular, when the base field is \mathbb{C}, we prove that our csmc_{sm} invariants agree with Kashiwara’s local Euler characteristics, i.e., the Euler characteristic of the complex link spaces of the orbits to 0. We prove an algebraic version of [36, Theorem 0.2] using our csmc_{sm} invariants, which specialize to a Brasselet-Lê-Seade type formula (Proposition 2.8) .

In §4 we discuss the sectional Euler characteristics of group orbits satisfy Assumption 2. The main result is Theorem 4.1, in which we prove a formula to the Euler characteristics of such orbits with any given hyperplane in terms of their Euler obstructions:

Theorem.

With the Assumption 2. Let en,ke_{n,k} and en,ke^{\prime}_{n,k} be their local Euler obstructions of the pair of orbits and their dual orbits. Let LrSrL_{r}\in S^{\prime}_{r} be the corresponding hyperplane, and let A=[αi,j]n1×n1A=[\alpha_{i,j}]_{n-1\times n-1} be the inverse matrix of the matrix E=[ei,j]n1×n1E=[e_{i,j}]_{n-1\times n-1} of local Euler obstructions. Then we have

χSkLr=\displaystyle\chi_{S_{k}\cap L_{r}}= i=rn1αk,i(1)di+dni+Neni,r+χ(Sk)Sm(Sk).\displaystyle\sum_{i=r}^{n-1}\alpha_{k,i}\cdot(-1)^{d_{i}+d^{\prime}_{n-i}+N}e^{\prime}_{n-i,r}+\chi(S_{k})-{Sm}(S_{k})\/.

We also give explicit description on when a hypersurface is generic with respect to sectional Euler characteristic.

In particular, for reflective recursive group orbits (group orbits satisfy Assumption 1 and Assumption 3) we prove that, the local Euler obstructions of such orbits are equivalent to their sectional Euler characteristics, i.e., they can be deduced from each other. Thus for such group orbits we have proved that the three sets of local invariants-the local Euler obstructions, the sectional Euler characteristics and the csmc_{sm} invariants-are equivalent.

As application we apply the formulas to a special type of reflective recursive group orbits: the ordianry, skew-symmetric and symmetric matrix rank loci. In §5 we briefly review the basic properties of the matrix rank loci, and in §6 we explicitly compute the local Euler obstructions for all three types rank loci. The main results are Theorem 6.2, Theorem 6.4 and Theorem 6.6. When the base field is \mathbb{C}, the formula is known for ordinary rank loci in [17] and for skew-symmetric rank loci by [34], proved by concrete topological computation. It is not known for symmetric case. Our method, however, is systematic and algebraic: we translate the computation of the topological Euler characteristic of the link spaces into standard Schubert calculus: the computations of tautological Chern classes over Grassmannians. Then we take advantage of the nice enumerative properties on Grassmannians.

With the knowledge of local Euler obstructions, in §7 we apply Theorem 4.1 and compute the sectional Euler characteristics for the rank loci. For all three cases, we give explicit formulas to the Euler characteristics of the intersections of the rank loci with any given hyperplanes. The formulas in Theprem 7.1, Theorem 7.2 and Theorem 7.3 are summations of simple binomials, which can be easily carried out by hand.

The famous Kashiwara index theorem shows that the local Euler obstructions connect as a bridge from the microlocal multiplicities to the stalk Euler characteristics for constructible sheaf complexes. For stratified singular complex varieties we are particularly interested in the Intersection cohomology sheaf complexes: they are bounded complexes of constructible sheaves that reflect the limiting behaviors from strata to their boundaries. As an application in Theorem 8.3 we compute the stalk Euler characteristic of the Intersection cohomology sheaves for all three types of rank loci. The computations indicate that, these topological invariant of ordinary rank loci behave very similarly to the skew-symmetric rank loci, but very differently from symmetric rank loci. For further discussions of this phenomenon we refer to [4] and [9].

As pointed in [17], for ordinary rank loci the local Euler obstructions are, surprisingly, Newton binomials fitting into the Pascal triangle. The skew-symmetric rank loci behaves the same with ordinary case. For symmetric case this is almost true, except that the Euler obstructions are zero on even size but odd rank rank loci. For other cases the local Euler obstructions are also Newton binomials. When the base field is \mathbb{C}, via MacPherson’s definitions this vanishing phenomenon says that for such cases we can always extend the distance 11-form on the Nash transform. It should be interesting to study from topology why such cases are special. Moreover, as singularities in the rank stratification are the same as singularities of certain Schubert varieties in the Grassmannians (for symmetric and skew-symmetric case the Grassmannians should be chosen as Grassmannians of isotropic subspaces with a symplectic or symmetric bilinear form), our results in §6 support the Conjecture 10.2 in [32].

Acknowledgment

The author is grateful to Leonardo Mihalcea, Richard Rimányi, Fei Si, Changjian Su and Dingxin Zhang for many useful discussions and suggestions. The author also would like to thank the referees for all the comments. The author is supported by China Postdoctoral Science Foundation (Grant No.2019M661329).

2. Preliminary

Unless specifically mentioned, all varieties in this note are closed projective, and the base field kk is algebraically closed of characteristic 0.

2.1. Characteristic Classes

Let X(V)X\subset\mathbb{P}(V) be a closed subvariety. A constructible function on XX is a finite sum WmW𝟙W\sum_{W}m_{W}\mathds{1}_{W} over closed subvarieties WW of XX, where the coefficients mWm_{W} are integers and 𝟙W\mathds{1}_{W} is the indicator function on WW. Let F(X)F(X) be the addition group of constructible functions on XX. It is isomorphic to the group of cycles 𝒵(X)\mathcal{Z}(X).

Let f:XYf\colon X\to Y be a proper morphism. One defines a homomorphism Ff:F(X)F(Y)Ff\colon F(X)\to F(Y) by setting, for all pYp\in Y, Ff(𝟙W)(p):=χ(f1(p)W)Ff(\mathds{1}_{W})(p):=\chi(f^{-1}(p)\cap W), and extending this definition by linearity. This makes FF into a functor from the category of kk-varieties 𝒱𝒜\mathcal{VAR} to 𝒜\mathcal{AB}, the category of abelian groups. The following theorem was first proved in [31] by MacPherson, to answer a conjecture proposed by Deligne and Grothendieck, then generalized to characteristic 0 algebraically field by G. Kennedy in [29].

Theorem 2.1 (MacPherson, Kennedy).

Let XX be a projective variety. There is a unique natural transformation cc_{*} from the functor FF to the Chow functor AA such that if XX is smooth, then c(𝟙X)=c(TX)[X]c_{*}(\mathds{1}_{X})=c(T_{X})\cap[X], where TXT_{X} is the tangent bundle of XX.

The proof of the theorem may be summarized as the following steps.

  1. (1)

    For any closed projective variety XX, there is a local measurement of the singularity called local Euler obstruction. This is a constructible function on XX denoted by EuXEu_{X}, i.e., EuX=WeW𝟙WEu_{X}=\sum_{W}e_{W}\mathds{1}_{W} for some sub-varieties WW of XX.

  2. (2)

    For any XX, the Euler obstruction functions along all closed subvarieties {EuW}\{Eu_{W}\} form a basis for F(X)F(X).

  3. (3)

    For any closed subvariety WW, there is an assigned characteristic class cM(W)A(W)c_{M}(W)\in A_{*}(W) defined via Nash blowup, called Chern-Mather class.

  4. (4)

    Let i:WXi\colon W\to X be the closed embedding. Define c(EuW)=i(cM(W))c_{*}(Eu_{W})=i_{*}(c_{M}(W)) to be the pushforward of the Chern-Mather class of WW in A(X)A_{*}(X). This is the unique natural transformation that matches the desired normalization property.

The Euler obstruction was originally defined over \mathbb{C} via obstruction theory to extend the lift of a non-zero vector field to the Nash transform, we refer [31] for details. In [18] the author gave an algebraic formula works for arbitrary algebraically closed field. Here we use the formula in [18] as our definition. Let XMX\subset M be a closed subvariety of dimension dd, we define the Nash transform of XX to be

X^:=closure of {(x,TxX)|x is a smooth point}Gd(TM)\hat{X}:=\text{closure of }\{(x,T_{x}X)|x\text{ is a smooth point}\}\subset G_{d}(TM)

The projection map p:X^Xp\colon\hat{X}\to X is birational and is isomorphic over the smooth locus XsmX_{sm}. The universal subbundle SS of Gd(TM)G_{d}(TM) restricts to a rank dd vector bundle on X^\hat{X}, denoted by 𝒯X\mathcal{T}_{X}.

Definition.

The local Euler obstruction of XX is defined as follows: for any xXx\in X we define

EuX(x):=p1(x)c(𝒯X)s(p1(x),X^).Eu_{X}(x):=\int_{p^{-1}(x)}c(\mathcal{T}_{X})\cap s(p^{-1}(x),\hat{X})\/.

The Chern-Mather class of XX is defined as

cM(X):=p(c(𝒯X)[X^]).c_{M}(X):=p_{*}(c(\mathcal{T}_{X})\cap[\hat{X}])\/.

The Chern-Schwartz-MacPherson class of XX is defined as csm(X):=c(𝟙X)c_{sm}(X):=c_{*}(\mathds{1}_{X}). Thus we have

csm(X)=kakcM(Wk)c_{sm}(X)=\sum_{k}a_{k}c_{M}(W_{k})

providing that 𝟙X=kakEuWk\mathds{1}_{X}=\sum_{k}a_{k}Eu_{W_{k}}.

Proposition 2.2.

We have the following properties.

  1. (1)

    EuX(x)Eu_{X}(x) is a local invariant, thus only depends on an open neighborhood of xx in XX.

  2. (2)

    If xXx\notin X, then EuX(x)=0Eu_{X}(x)=0.

  3. (3)

    If xXx\in X is a smooth point, then EuX(x)=1Eu_{X}(x)=1. But notice that EuX(x)=1Eu_{X}(x)=1 does NOT imply that xx is smooth (Cf. [33] and [32], and Theorem 6.6).

  4. (4)

    The Euler obstruction has the product property, i.e., EuX×Y(x×y)=EuX(x)×EuY(y)Eu_{X\times Y}(x\times y)=Eu_{X}(x)\times Eu_{Y}(y).

Definition.

Let XX be a projective or affine variety. By a stratification on XX we mean a disjoint union X=αSαX=\cup_{\alpha}S_{\alpha} such that

  1. (1)

    Each SαS_{\alpha} is a smooth quasi-projective variety.

  2. (2)

    For any pp and qq in SαS_{\alpha}, we have EuX(p)=EuX(q)Eu_{X}(p)=Eu_{X}(q).

  3. (3)

    If SαS¯βS_{\alpha}\cap\bar{S}_{\beta}\neq\emptyset, then we have SαS¯βS_{\alpha}\subset\bar{S}_{\beta}.

Example 1.

Let XX be a proejctive variety, we define X0=XsmX_{0}=X_{sm} to be the smooth locus of XX, and X¯k+1\bar{X}_{k+1} to be the singularity of X¯k\bar{X}_{k}. This process will stop in finite steps, since X¯k+1\bar{X}_{k+1} is closed in X¯k\bar{X}_{k} and thus dimX¯k+1<dimX¯k\dim\bar{X}_{k+1}<\dim\bar{X}_{k}. The strata Xk:=Xk¯X¯k+1X_{k}:=\bar{X_{k}}\smallsetminus\bar{X}_{k+1} satisfy the above assumptions. We call this the Singularity Stratification of XX.

Remark 1.

Assuming that XX is complete. If we consider the constant map k:X{p}k\colon X\to\{p\}, then the covariance property of cc_{*} shows that

Xcsm(Y)=\displaystyle\int_{X}c_{sm}(Y)= {p}Afc(𝟙Y)={p}cFf(𝟙Y)\displaystyle\leavevmode\nobreak\ \int_{\{p\}}Afc_{*}(\mathds{1}_{Y})=\int_{\{p\}}c_{*}Ff(\mathds{1}_{Y})
=\displaystyle= {p}χ(Y)c(𝟙{p})=χ(Y).\displaystyle\leavevmode\nobreak\ \int_{\{p\}}\chi(Y)c_{*}(\mathds{1}_{\{p\}})=\chi(Y).

This observation gives a generalization of the classical Poincaré-Hopf Theorem to possibly singular varieties.

2.2. Sectional Euler Characteristic

In [2] Aluffi shows that, for projective varieties the generic sectional Euler characteristic can be obtained by studying their Chern-Schwzrtz-MacPherson classes. Let X(V)X\subset\mathbb{P}(V) be a projective variety of dimension nn. For any r0r\geq 0 we define

Xr=XH1HrX_{r}=X\cap H_{1}\cap\cdots\cap H_{r}

to be the intersection of XX with rr generic hyperplanes. Let χ(Xr)=Xrcsm(Xr)\chi(X_{r})=\int_{X_{r}}c_{sm}(X_{r}) be its Euler characteristic, we define χX(t)=iχ(Xr)(t)r\chi_{X}(t)=\sum_{i}\chi(X_{r})\cdot(-t)^{r} to be the corresponding polynomial. This is a polynomial of degree n\leq n.

Also, notice that A((V))=[H]/Hn+1A_{*}(\mathbb{P}(V))=\mathbb{Z}[H]/H^{n+1} is a polynomial ring, thus we can write icsm(X)=i0γi[i]=i0γNiHii_{*}c_{sm}(X)=\sum_{i\geq 0}\gamma_{i}[\mathbb{P}^{i}]=\sum_{i\geq 0}\gamma_{N-i}H^{i}. Let γX(t)=iγiti\gamma_{X}(t)=\sum_{i}\gamma_{i}t^{i} and csmX(t)=iγNitic_{sm}^{X}(t)=\sum_{i}\gamma_{N-i}t^{i} be the assigned polynomials respectively. They are polynomials of degree n\leq n.

We have the following beautiful theorem connecting csmc_{sm} class and sectional Euler chracteristics:

Theorem 2.3 (Aluffi).

We consider the transformation SM:[t]n[t]n\mathcal{I}_{SM}\colon\mathbb{Z}[t]^{\leq n}\to\mathbb{Z}[t]^{\leq n} defined by

P(t)SM(P)(t):=tp(t1)+p(0)t+1.P(t)\mapsto\mathcal{I}_{SM}(P)(t):=\frac{t\cdot p(-t-1)+p(0)}{t+1}\/.

This is a \mathbb{Z}-linear involution, i.e., SM2(P)=P\mathcal{I}_{SM}^{2}(P)=P. Moreover, we have

SM(γX(t))=χX(t);SM(χX(t))=γX(t).\mathcal{I}_{SM}(\gamma_{X}(t))=\chi_{X}(t);\quad\mathcal{I}_{SM}(\chi_{X}(t))=\gamma_{X}(t)\/.

In particular, we have the following property.

Proposition 2.4.

For a generic hyperplane HH, we have

csmX(1)=(1)n(χ(X)χ(XH)).c_{sm}^{X}(-1)=(-1)^{n}(\chi(X)-\chi(X\cap H))\/.
Proof.

By the theorem, we have

γX(t)=\displaystyle\gamma_{X}(t)= tχX(1t)+χ(0)t+1\displaystyle\frac{t\cdot\chi_{X}(-1-t)+\chi(0)}{t+1}
=\displaystyle= t(χX(0)χ1(1t)+χ2(1t)2)+χX(0)1+t\displaystyle\frac{t\cdot(\chi_{X}(0)-\chi_{1}(-1-t)+\chi_{2}(-1-t)^{2}-\cdots)+\chi_{X}(0)}{1+t}
=\displaystyle= (t+1)χ0+tχ1(t+1)+tχ2(1+t)2+1+t\displaystyle\frac{(t+1)\chi_{0}+t\cdot\chi_{1}\cdot(t+1)+t\cdot\chi_{2}\cdot(1+t)^{2}+\cdots}{1+t}
=\displaystyle= χ0+tχ1+t(1+t)F(t)\displaystyle\chi_{0}+t\cdot\chi_{1}+t(1+t)\cdot F(t)

Thus

csmX(1)=(1)nγX(1)=(1)n(χ(X)χ(XH)).c_{sm}^{X}(-1)=(-1)^{n}\cdot\gamma_{X}(-1)=(-1)^{n}(\chi(X)-\chi(X\cap H))\/.

Thus if we know the csmc_{sm} class of XX, then we can obtain the generic sectional Euler characteristics χ(Xr)\chi(X_{r}). In general, the characteristic classes of a variety may be hard to compute. However, as in the property, we only need the alternating sum csmX(1)c_{sm}^{X}(-1), which in practice is much easier to compute.

Definition.

Let X(V)X\subset\mathbb{P}(V) be a (quasi)-projective variety, and let ΣV\Sigma\subset V be its affine cone. We define the csmc_{sm} invariant of XX (or Σ\Sigma) as

SmX=SmΣ:=(1)dimV1csmX(1)=χ(X)χ(XH).{Sm}_{X}={Sm}_{\Sigma}:=(-1)^{\dim V-1}c_{sm}^{X}(-1)=\chi(X)-\chi(X\cap H)\/.

In fact, if we write csm(X)=i=0dimXXi;c_{sm}(X)=\sum_{i=0}^{\dim X}X_{i}; where XiAi(X)X_{i}\in A_{i}(X) denotes the dimension ii component of csm(X)c_{sm}(X). One can see that Sm(X)=i=0dimX(1)idegXi{Sm}(X)=\sum_{i=0}^{\dim X}(-1)^{i}\deg X_{i}. Thus this definition is intrinsic, i.e., independent of the embedding of XX.

2.3. Dual Varieties and Radon Transform

Let Xn=(V)X\in\mathbb{P}^{n}=\mathbb{P}(V) be a projective variety, and let n=(V)\mathbb{P}^{*n}=\mathbb{P}(V^{*}) be the (projective) dual space of hyperplanes in n\mathbb{P}^{n}, i.e., any point in (V)\mathbb{P}(V^{*}) is a hyperplane in (V)\mathbb{P}(V). The dual variety of XX, denoted by XX^{\vee}, is defined as the closure of the following

{Hn|there is some xXsm such that TxXsmH}.\{H\in\mathbb{P}^{*n}|\text{there is some }x\in X_{sm}\text{ such that }T_{x}X_{sm}\subset H\}\/.

Here XsmX_{sm} denotes the smooth locus of XX. The variety XX and its dual are related as follows. Consider the product (V)×(V)\mathbb{P}(V)\times\mathbb{P}(V^{*}) and define IXI_{X} to be the closure of

IX:={(x,h)|xXsm;H(V) s.t. TxXsmH}I_{X}^{\circ}:=\{(x,h)|x\in X_{sm};H\in\mathbb{P}(V^{*})\text{ s.t. }T_{x}X_{sm}\subset H\}

to the the incidence correspondence. Let pp and qq denote the first and second projection. Then one can check that p(IX)=Xp(I_{X})=X, and q(IX)=Xq(I_{X})=X^{\vee}. The correspondence induces the definition of Radon transform:

Definition.

The topological Radon transform is the group homomorphism

F(X)F(X):λλ:=qpλ.F(X)\to F(X^{\vee})\colon\quad\lambda\mapsto\lambda^{\vee}:=q_{*}p^{*}\lambda\/.

If we write λ=ini𝟙Wi\lambda=\sum_{i}n_{i}\mathds{1}_{W_{i}} , then by the definition of proper pushforward and pull back we have

λ(L)=iniχ(WiL)\lambda^{\vee}(L)=\sum_{i}n_{i}\chi(W_{i}\cap L)

for any point L(V)L\in\mathbb{P}(V^{*}).

In particular, as proved in [15], the Radon transform takes the Euler obstruction function in XX to the Euler obstruction function in XX^{\vee}, up to an error term.

Theorem 2.5 (Ernström).

Let XX be of dimension nn, and XX^{\vee} be of dimension mm. Let VV be of dimension N+1N+1. Then

(EuX)=(1)n+(N1)mEuX+eX𝟙(V)(Eu_{X})^{\vee}=(-1)^{n+(N-1)-m}Eu_{X^{\vee}}+e_{X}\mathds{1}_{\mathbb{P}(V^{*})}

Here eX:=(EuX)(H)e_{X}:=(Eu_{X})^{\vee}(H) for generic hyperplanes HH.

This theorem shows that the Euler obstruction of the dual variety and the (signed) dual of the Euler obstruction differ by the Euler characteristic of generic hyperplane section.

Remark 2.

Recall that there is a unique expression 𝟙X=ieiEuWi\mathds{1}_{X}=\sum_{i}e_{i}Eu_{W_{i}} for closed subvarieties WiW_{i}. Then the generic value exe_{x} can be expressed as

eX=ieiχ(WiH)e_{X}=\sum_{i}e_{i}\chi(W_{i}\cap H)

for a generic hyperplane HH. Here by generic we mean that, by Bertini’s theorem there is an open dense subset UU in (V)\mathbb{P}(V^{*}) such that χ(WiH)\chi(W_{i}\cap H) is constant for any HUH\in U.

Corollary 1.

Let XX and XX^{\vee} be a pair of dual varieties. Assume that X=i=1mSiX=\cup_{i=1}^{m}S_{i} is a stratification. Let LL be a hyperplane such that LL lives outside XX^{\vee}. Then LL is Euler characteristic generic with respect to XX, i.e.,

χ(XL)=χ(XH)\chi(X\cap L)=\chi(X\cap H)

for generic hyperplanes HUH\in U. In particular, if LS¯iL\notin\bar{S}_{i}^{\vee}, then LL is Euler characteristic generic with respect to SiS_{i}: χ(SiL)=χ(SiH)\chi(S_{i}\cap L)=\chi(S_{i}\cap H).

Moreover, when XX^{\vee} admits a stratification Ti\cup T_{i}, we have χ(XL1)=χ(XL2)\chi(X\cap L_{1})=\chi(X\cap L_{2}) for any LiL_{i} and LiL^{\prime}_{i} living in the same stratum TiT_{i}.

Proof.

Recall that lXl\notin X^{\vee} implies that EuX(l)=0Eu_{X^{\vee}}(l)=0. Let eje_{j} be the local Euler obstruction EuX(Sj)Eu_{X}(S_{j}), then the theorem says:

(EuX)(L)=jejχ(SjL)=0+iejχ(SiH)(Eu_{X})^{\vee}(L)=\sum_{j}e_{j}\chi(S_{j}\cap L)=0+\sum_{i}e_{j}\chi(S_{i}\cap H)

Since there are only finitely many strata, by induction on the dimension of strata it suffice to prove the argument for smooth XX, otherwise its singularity locus induces a smaller stratum. When XX is smooth, we have EuX=𝟙XEu_{X}=\mathds{1}_{X}, and the theorem shows χ(XL)=χ(XH)\chi(X\cap L)=\chi(X\cap H).

When X=TiX^{\vee}=\cup T_{i} is a stratification, and {Li,Li}Ti\{L_{i},L_{i}^{\prime}\}\subset T_{i} are two hyperplanes of (V)\mathbb{P}(V). The theorem shows that (EuX)(Li)=(EuX)(Li)(Eu_{X})^{\vee}(L_{i})=(Eu_{X})^{\vee}(L^{\prime}_{i}). Thus we have

jejχ(SjLi)=jejχ(SjLi).\sum_{j}e_{j}\chi(S_{j}\cap L_{i})=\sum_{j}e_{j}\chi(S_{j}\cap L^{\prime}_{i})\/.

Then the same induction on the dimensions of SiS_{i} completes the proof. ∎

2.4. Projective Duality

As pointed out in [2, Prop 3.20], the two-way projections from conormal space induce an involution of Chern-Mather classes of dual varieties:

Theorem 2.6 (Aluffi).

Let X(V)X\subset\mathbb{P}(V) be a proper projective variety of dimension dd, and let XX^{\vee} be its dual variety of dimension dd^{*}. We write cMX(H)A((V))c_{M}^{X}(H)\in A_{*}(\mathbb{P}(V)) and cMX(H)A((V))c_{M}^{X^{\vee}}(H)\in A_{*}(\mathbb{P}(V^{*})) as polynomials in HH. Then

𝒥n((1)dcMX(H))=(1)dcMX(H);𝒥n((1)dcMX(H))=(1)dcMX(H)\mathcal{J}_{n}((-1)^{d}c_{M}^{X}(H))=(-1)^{d^{*}}c_{M}^{X^{\vee}}(H);\quad\mathcal{J}_{n}((-1)^{d^{*}}c_{M}^{X^{\vee}}(H))=(-1)^{d}c_{M}^{X}(H)

Here dimV=n+1\dim V=n+1, and 𝒥n:[H][H]\mathcal{J}_{n}\colon\mathbb{Z}[H]\to\mathbb{Z}[H] is defined as

𝒥n(p(H))=p(1H)p(1)((1+H)n+1Hn+1).\mathcal{J}_{n}(p(H))=p(-1-H)-p(-1)\left((1+H)^{n+1}-H^{n+1}\right)\/.

As corollary of the duality theorem, we have the following result.

Proposition 2.7.

Let X(V)X\subset\mathbb{P}(V) be a proper projective variety. Let CXC_{X} to be the affine cone of XX in VV. Then

EuCX(0)=(1)dimV1cMX(1).Eu_{C_{X}}(0)=(-1)^{\dim V-1}c_{M}^{X}(-1)\/.

Moreover, let XX^{\vee} be the dual variety of XX in (V)\mathbb{P}(V^{*}), then we have

(1)d+d+dimVEuCX(0)=EuCX(0).(-1)^{d+d^{*}+\dim V}\cdot Eu_{C_{X}}(0)=Eu_{C_{X^{\vee}}}(0)\/.
Proof.

The first argument is Proposition 3.17 in [2]. For the second part, plug H=1H=-1 into the involution formula we have

EuCX(0)=\displaystyle Eu_{C_{X^{\vee}}}(0)= (1)dimV1cMX(1)=(1)dimV1+d𝒥n((1)dcMX(H))(1)\displaystyle(-1)^{\dim V-1}c_{M}^{X^{\vee}}(-1)=(-1)^{\dim V-1+d^{*}}\mathcal{J}_{n}((-1)^{d}c_{M}^{X}(H))(-1)
=\displaystyle= (1)dimV1+d+d(cMX(0)+cMX(1)(1)dimV)\displaystyle(-1)^{\dim V-1+d^{*}+d}\left(c_{M}^{X}(0)+c_{M}^{X}(-1)\cdot(-1)^{\dim V}\right)
=\displaystyle= (1)d+d+dimVEuCX(0)\displaystyle(-1)^{d+d^{*}+\dim V}\cdot Eu_{C_{X}}(0)

And we have the following algebraic Brasselet-Lê-Seade type formula.

Proposition 2.8.

Let X(V)X\subset\mathbb{P}(V) be a projective variety, and let X=iSiX=\cup_{i}S_{i} be a stratification. Let ΣV\Sigma\subset V and ViV_{i} be the affine cones of XX and SiS_{i} respectively. Recall that SmSi:=(1)dimV1csmSi(1){Sm}_{S_{i}}:=(-1)^{\dim V-1}c_{sm}^{S_{i}}(-1). Then we have

EuΣ(0)=iEuX(Si)SmSi.Eu_{\Sigma}(0)=\sum_{i}Eu_{X}(S_{i})\cdot{Sm}_{S_{i}}\/.
Proof.

This follows directly from the definition of the natural transformation cc_{*}:

EuΣ(0)=\displaystyle Eu_{\Sigma}(0)= (1)dimV1cMX(1)\displaystyle(-1)^{\dim V-1}c_{M}^{X}(-1)
=\displaystyle= (1)dimV1iEuX(Si)csmSi(1).\displaystyle(-1)^{\dim V-1}\sum_{i}Eu_{X}(S_{i})c_{sm}^{S_{i}}(-1)\/.

More generally, let α\alpha be a constructible function on Σ\Sigma, i.e., α=iaiEuVi¯\alpha=\sum_{i}a_{i}Eu_{\bar{V_{i}}}. Then from previous Proposition we have

α(0)=iaiSmSi.\alpha(0)=\sum_{i}a_{i}\cdot{Sm}_{S_{i}}\/.

As we will see, this is exactly [36, Theorem 0.2] when restricted to base field \mathbb{C}.

In the analytic setting, let VV be a \mathbb{C}-vector space of dimension nn. Let X(V)X\subset\mathbb{P}(V) be a closed subvariety with a Whitney stratification i=1mSi\cup_{i=1}^{m}S_{i}. Let ViVV_{i}\subset V be the affine cones of SiS_{i}, and let C(X)C(X) be the affine cone over XX. For each ViV_{i} we consider the space Ll.t(Vi):=ViBϵl1(t)L_{l.t}(V_{i}):=V_{i}\cap B_{\epsilon}\cap l^{-1}(t). Here ll is a generic linear function VV\to\mathbb{C}, BϵB_{\epsilon} is the small ball centered at 0 and tt is sufficiently close to 0. The spaces Ll.t(Vi)L_{l.t}(V_{i}) are homotopic equivalent for generic ll and tt small enough, and are called the link space of ViV_{i} to 0. We show that the Euler characteristics of the link spaces are exactly our csmc_{sm} invariants SmVi{Sm}_{V_{i}}.

Proposition 2.9.

When the base field kk is \mathbb{C}, for any imi\leq m we have

SmVi=χ(ViBϵl1(t)).{Sm}_{V_{i}}=\chi(V_{i}\cap B_{\epsilon}\cap l^{-1}(t))\/.

In particular, we build the following affine-projective connection:

χ(C(X)Bϵl1(t))=χ(X)χ(XH)\chi(C(X)\cap B_{\epsilon}\cap l^{-1}(t))=\chi(X)-\chi(X\cap H)

Here HH is a generic hyperplane.

Proof.

Note that for any kmk\leq m, the decomposition V¯k=ikVi{0}\bar{V}_{k}=\cup_{i\geq k}V_{i}\cup\{0\} is a Whitney stratification of V¯k\bar{V}_{k} such that 0V¯i0\in\bar{V}_{i} for every kk. We recall the famous Brasselet-Lê-Seade formula in [7]:

EuV¯k(0)=ikχ(ViBϵl1(t))EuV¯k(Vi).Eu_{\bar{V}_{k}}(0)=\sum_{i\geq k}\chi(V_{i}\cap B_{\epsilon}\cap l^{-1}(t))\cdot Eu_{\bar{V}_{k}}(V_{i})\/.

Combine with previous corollary we then have

ikmEuS¯k(Si)csmSi(1)=\displaystyle\sum_{i\geq k}^{m}Eu_{\bar{S}_{k}}(S_{i})\cdot c_{sm}^{S_{i}}(-1)= cMS¯k(1)=(1)n1EuV¯k(0)\displaystyle c_{M}^{\bar{S}_{k}}(-1)=(-1)^{n-1}Eu_{\bar{V}_{k}}(0)
=\displaystyle= (1)n1ikmχ(ViBϵl1(t))EuV¯k(Vi).\displaystyle(-1)^{n-1}\sum_{i\geq k}^{m}\chi(V_{i}\cap B_{\epsilon}\cap l^{-1}(t))\cdot Eu_{\bar{V}_{k}}(V_{i})\/.

For any point xSix\in S_{i}, notice that locally we can view VkV_{k} as the product Sk×S_{k}\times\mathbb{C}^{*}. Thus the product property for the local Euler obstruction shows that: EuV¯k(Vi)=EuS¯k(Si)Eu_{\bar{V}_{k}}(V_{i})=Eu_{\bar{S}_{k}}(S_{i}), and thus

ikmEuS¯k(Si)(1)n1csmSi(1)=ikmχ(ViBϵl1(t))EuS¯k(Si).\sum_{i\geq k}^{m}Eu_{\bar{S}_{k}}(S_{i})\cdot(-1)^{n-1}\cdot c_{sm}^{S_{i}}(-1)=\sum_{i\geq k}^{m}\chi(V_{i}\cap B_{\epsilon}\cap l^{-1}(t))\cdot Eu_{\bar{S}_{k}}(S_{i})\/.

This equation holds for every k{1,2,,m}k\in\{1,2,\cdots,m\}, thus we have the following linear system

ikmEuS¯k(Si)ζi=0;k=1,2,m\sum_{i\geq k}^{m}Eu_{\bar{S}_{k}}(S_{i})\cdot\zeta_{i}=0;\quad k=1,2,\cdots m

has ζi=(csmSi(1)(1)n1χ(ViBϵl1(t)))\zeta_{i}=\left(c_{sm}^{S_{i}}(-1)-(-1)^{n-1}\chi(V_{i}\cap B_{\epsilon}\cap l^{-1}(t))\right) as a solution. However, the matrix represents the linear system is upper triangular with 11 on the diagonal, the solution must be the zero vector. This proves the first formula. The rest of the proposition follows from the inclusion-exclusion property of both csmc_{sm} class and Euler characteristic:

χ(C(X)Bϵl1(t))=\displaystyle\chi(C(X)\cap B_{\epsilon}\cap l^{-1}(t))= i=1mχ(ViBϵl1(t))=i=1m(1)n1csmSi(1)\displaystyle\sum_{i=1}^{m}\chi(V_{i}\cap B_{\epsilon}\cap l^{-1}(t))=\sum_{i=1}^{m}(-1)^{n-1}c_{sm}^{S_{i}}(-1)
=\displaystyle= (1)n1csmX(1)=χ(X)χ(XH).\displaystyle(-1)^{n-1}c_{sm}^{X}(-1)=\chi(X)-\chi(X\cap H)\/.

The last equality comes from Proposition 2.4

2.5. A Segre Computation

We close this section with a technical Lemma that will be used in later computations.

Lemma 1.

Let EE be a rank ee vector bundle on XX, and let =𝒪(1)\mathcal{L}=\mathcal{O}(1) be the tautological line bundle on =(E)\mathbb{P}=\mathbb{P}(E). Let pp be the projection map from (E)\mathbb{P}(E) to XX. Then we have

p(c(E)c()[(E)])=[X]ce(E)[X]c(E).p_{*}\left(\frac{c(E\otimes\mathcal{L})}{c(\mathcal{L})}\cap[\mathbb{P}(E)]\right)=[X]-\frac{c_{e}(E^{\vee})\cap[X]}{c(E^{\vee})}\/.
Proof.

Following the tensor formula [16, Example 3.2.2] we have

c(E)=\displaystyle c(E\otimes\mathcal{L})= k=0e(i=0k(eiki)ci(E)c1()ki)\displaystyle\sum_{k=0}^{e}\left(\sum_{i=0}^{k}\binom{e-i}{k-i}c_{i}(E)\cdot c_{1}(\mathcal{L})^{k-i}\right)
=\displaystyle= k=0e(i=0ek(eki)c1()i)ck(E)\displaystyle\sum_{k=0}^{e}\left(\sum_{i=0}^{e-k}\binom{e-k}{i}c_{1}(\mathcal{L})^{i}\right)\cdot c_{k}(E)
=\displaystyle= k=0eck(E)(1+c1())ek\displaystyle\sum_{k=0}^{e}c_{k}(E)\cdot(1+c_{1}(\mathcal{L}))^{e-k}
=\displaystyle= k=0ecek(E)(1+c1())k;\displaystyle\sum_{k=0}^{e}c_{e-k}(E)\cdot(1+c_{1}(\mathcal{L}))^{k}\/;

Thus we have

c(E)c()=\displaystyle\frac{c(E\otimes\mathcal{L})}{c(\mathcal{L})}= k=0ecek(E)(1+c1())k1+c1()\displaystyle\frac{\sum_{k=0}^{e}c_{e-k}(E)\cdot(1+c_{1}(\mathcal{L}))^{k}}{1+c_{1}(\mathcal{L})}
=\displaystyle= ce(E)1+c1()+k=1ecek(E)(1+c1())k1;\displaystyle\frac{c_{e}(E)}{1+c_{1}(\mathcal{L})}+\sum_{k=1}^{e}c_{e-k}(E)\cdot(1+c_{1}(\mathcal{L}))^{k-1}\/;

Recall the definition of Segre class:

p(c1()e+i[(E)])=si+1(E)[X].p_{*}(c_{1}(\mathcal{L})^{e+i}\cap[\mathbb{P}(E)])=s_{i+1}(E)\cap[X]\/.

The pushforward then becomes

p(c(E)c()[(E)])=\displaystyle p_{*}\left(\frac{c(E\otimes\mathcal{L})}{c(\mathcal{L})}\cap[\mathbb{P}(E)]\right)= p(ce(E)1+c1()[(E)])+s0(E)[X]\displaystyle p_{*}(\frac{c_{e}(E)}{1+c_{1}(\mathcal{L})}\cap[\mathbb{P}(E)])+s_{0}(E)\cap[X]
=\displaystyle= (1)e1s(E)ce(E)[X]+[X]\displaystyle(-1)^{e-1}s(E^{\vee})c_{e}(E)\cap[X]+[X]
=\displaystyle= [X]s(E)ce(E)[X].\displaystyle[X]-s(E^{\vee})c_{e}(E^{\vee})\cap[X]\/.

3. Local Euler Obstruction of Recursive Group Orbits

Now we consider the following situation: for each n1n\geq 1, the group GnG_{n} is a connected linear algebraic group and GnG_{n} acts algebraically on a vector space VnV_{n} of dimension lnl_{n}. Assume that for every such nn there are exactly nn orbits.

Assumption 1.

We assume the following

  1. (1)

    All actions contain sub-actions by kk^{*} multiplications. Thus the orbits are necessarily cones.

  2. (2)

    The orbits can be labeled by 𝒪n,0,𝒪n,1,,𝒪n,n\mathcal{O}_{n,0},\mathcal{O}_{n,1},\cdots,\mathcal{O}_{n,n} such that 𝒪n,j𝒪¯n,i\mathcal{O}_{n,j}\subset\bar{\mathcal{O}}_{n,i} whenever j>ij>i. Thus 𝒪n,0\mathcal{O}_{n,0} is the largest orbit and 𝒪n,n={0}\mathcal{O}_{n,n}=\{0\}.

  3. (3)

    The largest strata 𝒪n,0\mathcal{O}_{n,0} is dense in VnV_{n}.

  4. (4)

    For every nn, kk, and for any point p𝒪n,rp\in\mathcal{O}_{n,r}, the transverse normal slice pair (Np𝒪n,k,p)(N_{p}\cap\mathcal{O}_{n,k},p) is isomorphic to (𝒪r,k,0)(\mathcal{O}_{r,k},0). Thus we have local product structure Np×𝒪r,k𝒪n,kN_{p}\times\mathcal{O}_{r,k}\sim\mathcal{O}_{n,k}.

Example 2.

Let Gn=GLn×GLnG_{n}=GL_{n}\times GL_{n} act on Vn=knknV_{n}=k^{n}\otimes k^{n}, the space of n×nn\times n matrices by AXBA\cdot X\cdot B. The orbits are matrices of fixed corank kk, denoted by Σn,k\Sigma_{n,k}. Then for each point pp in Σn,jΣn,k\Sigma_{n,j}^{\circ}\subset\Sigma_{n,k}, the normal slice NpN_{p} intersect with Σn,k\Sigma_{n,k} at pp. Moreover we have NpΣn,k=Σj,kN_{p}\cap\Sigma_{n,k}=\Sigma_{j,k}, and isomorphism Σj,k×kNΣn,k\Sigma_{j,k}\times k^{N}\cong\Sigma_{n,k}. Here kNk^{N} is centered at pp of complementary dimension. Similar arguments for symmetric matrices also give such sequence of actions.

For skew-symmetric matrices, we consider the sequence of Gn=GL2nG_{n}=GL_{2n} actions on Vn=2k2nV_{n}=\wedge^{2}k^{2n} and the sequence of Gn=GL2n+1G_{n}=GL_{2n+1} action on Vn=2k2n+1V_{n}=\wedge^{2}k^{2n+1} separately. Each of them form a recursive group actions.

Remark 3.

We can loose the assumption as follows. In fact we don’t need all the orbits to match the assumption, among all the orbits we only concentrate a subset, i.e., a flag of them. Thus we can only assume the following: For each nn, there is a maximal flag of orbits labeled by 𝒪n,0,𝒪n,1,,𝒪n,n\mathcal{O}_{n,0},\mathcal{O}_{n,1},\cdots,\mathcal{O}_{n,n} such that 𝒪n,j𝒪¯n,i\mathcal{O}_{n,j}\subset\bar{\mathcal{O}}_{n,i} whenever j>ij>i, and 𝒪¯n,0=i=0n𝒪n,i\bar{\mathcal{O}}_{n,0}=\cup_{i=0}^{n}\mathcal{O}_{n,i}. Here 𝒪n,0\mathcal{O}_{n,0} is the largest orbit. We only require the sequence of flags to satisfy the rest of Assumption 1, except that we want 𝒪n,0\mathcal{O}_{n,0} to be dense in its closure, not VV.

Theorem 3.1 (Local Euler Obstructions and csmc_{sm} Invariants).

Assume that we have a sequence of group actions as in Assumption 1. For any 1kn11\leq k\leq n-1 we denote Sn,k=(𝒪n,k)(Vn)S_{n,k}=\mathbb{P}(\mathcal{O}_{n,k})\subset\mathbb{P}(V_{n}) to be the projectivizations. Denote the csmc_{sm} invariants of 𝒪n,k\mathcal{O}_{n,k} (or Sn,kS_{n,k}) by Smn,i{Sm}_{n,i}, i.e.,

Smn,i:=(1)ln1csmSn,i(1){Sm}_{n,i}:=(-1)^{l_{n}-1}c_{sm}^{S_{n,i}}(-1)

Then we have the following equivalence:

{Smr,k|k,r}\xlongleftrightarrowequivalent{ek,r|k,r}.\{{Sm}_{r,k}|k,r\}\xlongleftrightarrow{\text{equivalent}}\{e_{k,r}|k,r\}\/.

Here by equivalence we mean that they can be derived from each other. Moreover, the local Euler obstructions Eu𝒪n,r(𝒪n,k)Eu_{\mathcal{O}_{n,r}}(\mathcal{O}_{n,k}) can be computed as

Eu𝒪¯n,k(𝒪n,r)=Eu𝒪r,k(0)=(μ1>μ2>>μlk)Smnμ1Smμ1μ2SmμlkEu_{\bar{\mathcal{O}}_{n,k}}(\mathcal{O}_{n,r})=Eu_{\mathcal{O}_{r,k}}(0)=\sum_{(\mu_{1}>\mu_{2}>\cdots>\mu_{l}\geq k)}{Sm}_{n\mu_{1}}\cdot{Sm}_{\mu_{1}\mu_{2}}\cdots\cdot{Sm}_{\mu_{l}k}

Here the sum is over all (partial and full) flags (μ1>μ2>>μl)(\mu_{1}>\mu_{2}>\cdots>\mu_{l}) such that μi1=μi+1\mu_{i}-1=\mu_{i+1} for every ii and μlk\mu_{l}\geq k.

Proof.

Recall the product property of local Euler obstructions, then the normal slice assumption (Assumption I (4)(4)) shows that Eu𝒪¯n,k(𝒪n,r)=Eu𝒪¯r,k(𝒪r,r)Eu_{\bar{\mathcal{O}}_{n,k}}(\mathcal{O}_{n,r})=Eu_{\bar{\mathcal{O}}_{r,k}}(\mathcal{O}_{r,r}) is independent of nn. We denote this number by ek,re_{k,r}.

Since GnG_{n} acts linearly on VnV_{n}, the orbits Vn=i=0n𝒪n,iV_{n}=\cup_{i=0}^{n}\mathcal{O}_{n,i} form a stratification of VnV_{n} such that 0𝒪¯n,k0\in\bar{\mathcal{O}}_{n,k} for every kk. Proposition 2.7 shows that for any k=1,,n1k=1,\cdots,n-1 we have

ek,n:=Eu𝒪¯n,k(0)=(1)ln1cMS¯n,k(1)=(1)ln1i=kn1EuS¯n,k(Sn,i)csmSn,i(1)=i=kn1Smn,iekie_{k,n}:=Eu_{\bar{\mathcal{O}}_{n,k}}(0)=(-1)^{l_{n}-1}c_{M}^{\bar{S}_{n,k}}(-1)=(-1)^{l_{n}-1}\sum_{i=k}^{n-1}Eu_{\bar{S}_{n,k}}(S_{n,i})c_{sm}^{S_{n,i}}(-1)=\sum_{i=k}^{n-1}{Sm}_{n,i}\cdot e_{ki}

This formula applies to all nn and kk, thus we obtain a recursive formula. On the initial case k=nk=n, notice that 𝒪¯k,k={0}\bar{\mathcal{O}}_{k,k}=\{0\} and thus ekk=Eu𝒪¯k,k(0)=1.e_{kk}=Eu_{\bar{\mathcal{O}}_{k,k}}(0)=1. Then we have

ek,n=\displaystyle e_{k,n}= i=kn1Smn,ieki=i=kn1Smn,ij=ki1Smi,jeij\displaystyle\sum_{i=k}^{n-1}{Sm}_{n,i}\cdot e_{ki}=\sum_{i=k}^{n-1}{Sm}_{n,i}\sum_{j=k}^{i-1}{Sm}_{i,j}e_{ij}
=\displaystyle= \displaystyle\quad\cdots\cdots
=\displaystyle= (μ1>μ2>>μlk)Smm,μ1Smμ1,μ2Smμl,k\displaystyle\sum_{(\mu_{1}>\mu_{2}>\cdots>\mu_{l}\geq k)}{Sm}_{m,\mu_{1}}\cdot{Sm}_{\mu_{1},\mu_{2}}\cdots\cdot{Sm}_{\mu_{l},k}

The sum is over all the flags (μ1>μ2>>μl)(\mu_{1}>\mu_{2}>\cdots>\mu_{l}) such that μi1=μi+1\mu_{i}-1=\mu_{i+1} for every ii and μlk\mu_{l}\geq k.

Now we show that the local Euler obstructions recover the csmc_{sm} invariants. By definition we have

Smn,k=\displaystyle{Sm}_{n,k}= (1)ln1csmSn,k(1)=(1)ln1r=kn1arncMSn,r(1)\displaystyle(-1)^{l_{n}-1}c_{sm}^{S_{n,k}}(-1)=(-1)^{l_{n}-1}\sum_{r=k}^{n-1}a^{n}_{r}c_{M}^{S_{n,r}}(-1)
=\displaystyle= r=kn1arnEu𝒪¯r,k(0)=r=kn1arnek,r.\displaystyle\sum_{r=k}^{n-1}a^{n}_{r}Eu_{\bar{\mathcal{O}}_{r,k}}(0)=\sum_{r=k}^{n-1}a^{n}_{r}e_{k,r}\/.

Here the arna^{n}_{r} are the base change coefficients between 𝟙(𝒪n,r)\mathds{1}_{\mathbb{P}(\mathcal{O}_{n,r})} and Eu(𝒪n,r)Eu_{\mathbb{P}(\mathcal{O}_{n,r})}, which are completely determined by {er,n}\{e_{r,n}\}. This completes the proof. ∎

Remark 4.

When the base field is \mathbb{C}, this actually correspond to [11, Definition 4,1,36] via Proposition 2.9. The fact that the recursive definition using Kashiwara’s local Euler characteristics is equivalent to MacPherson’s orbstruction theoretic definition are equivalent was proved in [10]. Thus this theorem generalizes the equivalence from \mathbb{C} to arbitrary algebraically closed field of characteristic 0, using González-Sprinberg’s algebraic formula for local Euler obstructions and our csmc_{sm} invariants for local Euler characteristics.

4. Sectional Euler Characteristic of Group Orbits

In this section we consider the following situation.

Assumption 2.

Let GG be a connected linear algebraic group acting on VV with finitely many orbits. We label them by 𝒪0,𝒪1,,𝒪n\mathcal{O}_{0},\mathcal{O}_{1},\cdots,\mathcal{O}_{n} such that 𝒪i𝒪¯j\mathcal{O}_{i}\not\subset\bar{\mathcal{O}}_{j} whenever j>ij>i. As shown in [37], the dual action of GG induces nn orbits in VV^{*}. We can label them such that (𝒪¯i)=(𝒪¯ni)\mathbb{P}(\bar{\mathcal{O}}_{i})^{\vee}=\mathbb{P}(\bar{\mathcal{O}}^{\prime}_{n-i}) for i1i\geq 1. Then we also have 𝒪j𝒪¯i\mathcal{O}^{\prime}_{j}\not\subset\bar{\mathcal{O}^{\prime}}_{i} whenever i>ji>j. Thus the dense open orbits by 𝒪0\mathcal{O}_{0} and 𝒪0\mathcal{O}^{\prime}_{0}, and we have 𝒪n=𝒪n={0}\mathcal{O}_{n}=\mathcal{O}^{\prime}_{n}=\{0\}. Let did_{i} and did^{\prime}_{i} denote the dimensions of 𝒪i\mathcal{O}_{i} and 𝒪i\mathcal{O}^{\prime}_{i} respectively. We have d0=d0=dimV=Nd_{0}=d^{\prime}_{0}=\dim V=N.

Definition.

Let SiS_{i} and SiS^{\prime}_{i} be the projectivizations of 𝒪i\mathcal{O}_{i} and 𝒪i\mathcal{O}^{\prime}_{i}. For any i,ji,j in {1,,n1}\{1,\cdots,n-1\} we define the following numerical indices

ei,j:=EuS¯i(p);χi,j:=χ(SiLj)e_{i,j}:=Eu_{\bar{S}_{i}}(p);\quad\chi_{i,j}:=\chi(S_{i}\cap L_{j})

for any point pSjp\in S_{j}, and Lj𝒪jL_{j}\in\mathcal{O}^{\prime}_{j}. Symmetrically, we define

ei,j:=EuS¯i(p);χi,j:=χ(SiLj)e^{\prime}_{i,j}:=Eu_{\bar{S}^{\prime}_{i}}(p);\quad\chi^{\prime}_{i,j}:=\chi(S^{\prime}_{i}\cap L^{\prime}_{j})

for any point pSjp\in S^{\prime}_{j}, and LjSjL^{\prime}_{j}\in S_{j}. Let E=[ei,j]E=[e_{i,j}] and E=[ei,j]E^{\prime}=[e^{\prime}_{i,j}] be the n1×n1n-1\times n-1 matrices.

Knowing the local Euler obstructions of the orbits lead us to the information of the sectional Euler characteristics of the dual orbits:

Theorem 4.1.

Assume that a GG representation VV satisfies Assumption 2. Let A=[αi,j]n1×n1A=[\alpha_{i,j}]_{n-1\times n-1} and A=[αi,j]n1×n1A^{\prime}=[\alpha^{\prime}_{i,j}]_{n-1\times n-1} be the inverse matrix of EE and EE^{\prime} respectively, then we have

χk,r=\displaystyle\chi_{k,r}= i=rn1αk,i(1)di+dni+Neni,r+χ(Sk)Sm(Sk);\displaystyle\sum_{i=r}^{n-1}\alpha_{k,i}\cdot(-1)^{d_{i}+d^{\prime}_{n-i}+N}e^{\prime}_{n-i,r}+\chi(S_{k})-{Sm}(S_{k})\/;
χk,r=\displaystyle\chi^{\prime}_{k,r}= i=rn1αk,i(1)di+dni+Neni,r+χ(Sk)Sm(Sk).\displaystyle\sum_{i=r}^{n-1}\alpha^{\prime}_{k,i}\cdot(-1)^{d_{i}+d^{\prime}_{n-i}+N}e_{n-i,r}+\chi(S^{\prime}_{k})-{Sm}(S^{\prime}_{k})\/.
Proof.

Notice that the matrix E=[ei,j]E=[e_{i,j}] is upper triangular with 11 on the diagonal, thus it is invertible with inverse matrix AA also being upper triangular. This shows that αi,j=0\alpha_{i,j}=0 whenever i>ji>j. As pointed out in Corollary 1, for any SiS_{i} we have χi,0=χ(SiH)\chi_{i,0}=\chi(S_{i}\cap H) and χi,0=χ(SiH)\chi^{\prime}_{i,0}=\chi(S^{\prime}_{i}\cap H) equal the generic sectional Euler characteristic. Thus recall Theorem 2.5, for any lsSsl_{s}\in S^{\prime}_{s} we have

ikn1eki(χisχi0)=(1)dk+dnk+Neks\sum_{i\geq k}^{n-1}e_{ki}(\chi_{is}-\chi_{i0})=(-1)^{d_{k}+d^{\prime}_{n-k}+N}e^{\prime}_{ks}

Thus written in matrix form we have

E(χ1,1χ1,0χ12χ1,0χ1,n1χ1,0χ2,1χ2,0χ22χ2,0χ2,n1χ2,0χn1,1χn1,0χn1,2χn1,0χn1,n1χn1,0)=(00e^n1,n10e^n2,n2e^n2,n1e^1,1e^1,2e^1,n1).E\cdot\left(\begin{array}[]{cccc}\chi_{1,1}-\chi_{1,0}&\chi_{12}-\chi_{1,0}&\cdots&\chi_{1,n-1}-\chi_{1,0}\\ \chi_{2,1}-\chi_{2,0}&\chi_{22}-\chi_{2,0}&\cdots&\chi_{2,n-1}-\chi_{2,0}\\ \cdots&\cdots&\cdots&\cdots\\ \chi_{n-1,1}-\chi_{n-1,0}&\chi_{n-1,2}-\chi_{n-1,0}&\cdots&\chi_{n-1,n-1}-\chi_{n-1,0}\end{array}\right)=\left(\begin{array}[]{cccc}0&0&\cdots&\hat{e}^{\prime}_{n-1,n-1}\\ 0&\cdots&\hat{e}^{\prime}_{n-2,n-2}&\hat{e}^{\prime}_{n-2,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ \hat{e}^{\prime}_{1,1}&\hat{e}^{\prime}_{1,2}&\cdots&\hat{e}^{\prime}_{1,n-1}\end{array}\right).

Here e^i,j=(1)di+dni+Nei,j\hat{e}^{\prime}_{i,j}=(-1)^{d_{i}+d^{\prime}_{n-i}+N}e^{\prime}_{i,j}. Thus we have

χk,rχk,0=i=rn1αk,i(1)di+dni+Neni,r.\chi_{k,r}-\chi_{k,0}=\sum_{i=r}^{n-1}\alpha_{k,i}\cdot(-1)^{d_{i}+d^{\prime}_{n-i}+N}e^{\prime}_{n-i,r}\/.

Recall from Proposition 2.4 shows χk,0=χ(SkH)=χ(Sk)Sm(Sk)\chi_{k,0}=\chi(S_{k}\cap H)=\chi(S_{k})-{Sm}(S_{k}). Then we have

χk,r=\displaystyle\chi_{k,r}= i=rn1αk,i(1)di+dni+Neni,r+χ(Sk)Sm(Sk).\displaystyle\sum_{i=r}^{n-1}\alpha_{k,i}\cdot(-1)^{d_{i}+d^{\prime}_{n-i}+N}e^{\prime}_{n-i,r}+\chi(S_{k})-{Sm}(S_{k})\/.

The other formula is obtained directly by symmetry. ∎

In fact, as shown in the following proposition, the dual sectional Euler characteristics in (V)\mathbb{P}(V^{*}) is principally the same with the ones in (V)\mathbb{P}(V).

Proposition 4.2.

The following two matrices are inverse to each other.

(χ1,1χ1,0χ1,n1χ1,0χ2,1χ2,0χ2,n1χ2,0χn1,1χn1,0χn1,n1χn1,0)1=(χ1,1χ1,0χ1,n1χ1,0χ2,1χ2,0χ2,n1χ2,0χn1,1χn1,0χn1,n1χn1,0).\left(\begin{array}[]{ccc}\chi_{1,1}-\chi_{1,0}&\cdots&\chi_{1,n-1}-\chi_{1,0}\\ \chi_{2,1}-\chi_{2,0}&\cdots&\chi_{2,n-1}-\chi_{2,0}\\ \cdots&\cdots&\cdots\\ \chi_{n-1,1}-\chi_{n-1,0}&\cdots&\chi_{n-1,n-1}-\chi_{n-1,0}\end{array}\right)^{-1}=\left(\begin{array}[]{ccc}\chi^{\prime}_{1,1}-\chi^{\prime}_{1,0}&\cdots&\chi^{\prime}_{1,n-1}-\chi^{\prime}_{1,0}\\ \chi^{\prime}_{2,1}-\chi^{\prime}_{2,0}&\cdots&\chi^{\prime}_{2,n-1}-\chi^{\prime}_{2,0}\\ \cdots&\cdots&\cdots\\ \chi^{\prime}_{n-1,1}-\chi^{\prime}_{n-1,0}&\cdots&\chi^{\prime}_{n-1,n-1}-\chi^{\prime}_{n-1,0}\end{array}\right)\/.
Proof.

Set βk,r=χk,rχk,0\beta_{k,r}=\chi_{k,r}-\chi_{k,0} and βk,r=χk,rχk,0\beta^{\prime}_{k,r}=\chi^{\prime}_{k,r}-\chi^{\prime}_{k,0}, then we have

(e1,1e1,2e1,n10e2,2e2,n100en1,n1)(β1,1β1,n1β2,1β2,n1βn1,1βn1,n1)(β1,1β1,n1β2,1β2,n1βn1,1βn1,n1)\displaystyle\left(\begin{array}[]{cccc}e_{1,1}&e_{1,2}&\cdots&e_{1,n-1}\\ 0&e_{2,2}&\cdots&e_{2,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ 0&0&\cdots&e_{n-1,n-1}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta_{1,1}&\cdots&\beta_{1,n-1}\\ \beta_{2,1}&\cdots&\beta_{2,n-1}\\ \cdots&\cdots&\cdots\\ \beta_{n-1,1}&\cdots&\beta_{n-1,n-1}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{\prime}_{1,1}&\cdots&\beta^{\prime}_{1,n-1}\\ \beta^{\prime}_{2,1}&\cdots&\beta^{\prime}_{2,n-1}\\ \cdots&\cdots&\cdots\\ \beta^{\prime}_{n-1,1}&\cdots&\beta^{\prime}_{n-1,n-1}\end{array}\right)
=(00e^n1,n10e^n2,n2e^n2,n1e^1,1e^1,2e^1,n1)(β1,1β1,n1β2,1β2,n1βn1,1βn1,n1)\displaystyle=\left(\begin{array}[]{cccc}0&0&\cdots&\hat{e}^{\prime}_{n-1,n-1}\\ 0&\cdots&\hat{e}^{\prime}_{n-2,n-2}&\hat{e}^{\prime}_{n-2,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ \hat{e}^{\prime}_{1,1}&\hat{e}^{\prime}_{1,2}&\cdots&\hat{e}^{\prime}_{1,n-1}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{\prime}_{1,1}&\cdots&\beta^{\prime}_{1,n-1}\\ \beta^{\prime}_{2,1}&\cdots&\beta^{\prime}_{2,n-1}\\ \cdots&\cdots&\cdots\\ \beta^{\prime}_{n-1,1}&\cdots&\beta^{\prime}_{n-1,n-1}\end{array}\right)
=((1)Tn100(1)Tn20(1)T1)(0en1,n10en2,n1e1,1e1,n1)(β1,1β1,n1β2,1β2,n1βn1,1βn1,n1)\displaystyle=\left(\begin{array}[]{ccc}(-1)^{T_{n-1}}&\cdots&0\\ 0&(-1)^{T_{n-2}}&\cdots\\ \cdots&\cdots&\cdots\\ 0&\cdots&(-1)^{T_{1}}\end{array}\right)\cdot\left(\begin{array}[]{ccc}0&\cdots&e^{\prime}_{n-1,n-1}\\ 0&\cdots&e^{\prime}_{n-2,n-1}\\ \cdots&\cdots&\cdots\\ e^{\prime}_{1,1}&\cdots&e^{\prime}_{1,n-1}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{\prime}_{1,1}&\cdots&\beta^{\prime}_{1,n-1}\\ \beta^{\prime}_{2,1}&\cdots&\beta^{\prime}_{2,n-1}\\ \cdots&\cdots&\cdots\\ \beta^{\prime}_{n-1,1}&\cdots&\beta^{\prime}_{n-1,n-1}\end{array}\right)
=((1)Tn100(1)Tn20(1)T1)(e^1,1e^1,2e^1,n10e^2,2e^2,n100e^n1,n1)\displaystyle=\left(\begin{array}[]{ccc}(-1)^{T_{n-1}}&\cdots&0\\ 0&(-1)^{T_{n-2}}&\cdots\\ \cdots&\cdots&\cdots\\ 0&\cdots&(-1)^{T_{1}}\end{array}\right)\cdot\left(\begin{array}[]{cccc}\hat{e}_{1,1}&\hat{e}_{1,2}&\cdots&\hat{e}_{1,n-1}\\ 0&\hat{e}_{2,2}&\cdots&\hat{e}_{2,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ 0&0&\cdots&\hat{e}_{n-1,n-1}\end{array}\right)
=((1)Tn1+T100(1)Tn2+T20(1)T1+Tn1)(e1,1e1,2e1,n10e2,2e2,n100en1,n1)\displaystyle=\left(\begin{array}[]{ccc}(-1)^{T_{n-1}+T^{\prime}_{1}}&\cdots&0\\ 0&(-1)^{T_{n-2}+T^{\prime}_{2}}&\cdots\\ \cdots&\cdots&\cdots\\ 0&\cdots&(-1)^{T_{1}+T^{\prime}_{n-1}}\end{array}\right)\cdot\left(\begin{array}[]{cccc}e_{1,1}&e_{1,2}&\cdots&e_{1,n-1}\\ 0&e_{2,2}&\cdots&e_{2,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ 0&0&\cdots&e_{n-1,n-1}\end{array}\right)

Notice that Tk+Tnk=dk+dnk+N+dnk+dk+NT_{k}+T^{\prime}_{n-k}=d_{k}+d^{\prime}_{n-k}+N+d^{\prime}_{n-k}+d_{k}+N is even, thus we complete the proof. ∎

In particular, for the reflective recursive group orbits, the χ\chi indices of orbits 𝒪n,k\mathcal{O}_{n,k} are equivalent to their local Euler obstructions: they also determine the local Euler obstructions. Here by reflective we mean the following assumption:

Assumption 3 (Reflective Assumption).

For any GnG_{n} action on VnV_{n}, the dual orbits in VnV_{n}^{*} also have the normal slice property. More precisely, we assume that for any point p𝒪n,rp\in\mathcal{O}_{n,r}, the normal slice pair (Np𝒪n,k,p)(N_{p}\cap\mathcal{O}^{\prime}_{n,k},p) is isomorphic to (𝒪r,k,0)(\mathcal{O}^{\prime}_{r,k},0).

The normal slice assumptions shows that Eu𝒪¯n,k(𝒪n,r)=Eu𝒪¯r,k(𝒪r,r)Eu_{\bar{\mathcal{O}^{\prime}}_{n,k}}(\mathcal{O}^{\prime}_{n,r})=Eu_{\bar{\mathcal{O}^{\prime}}_{r,k}}(\mathcal{O}^{\prime}_{r,r}) are independent of also nn. Thus we can welly define the indices ek,r=Eu𝒪¯n,k(𝒪n,r)e_{k,r}=Eu_{\bar{\mathcal{O}}_{n,k}}(\mathcal{O}_{n,r}) and ek,r=Eu𝒪¯n,k(𝒪n,r)e^{\prime}_{k,r}=Eu_{\bar{\mathcal{O}^{\prime}}_{n,k}}(\mathcal{O}^{\prime}_{n,r}) respectively. Let Sn,kS_{n,k} and Sn,kS^{\prime}_{n,k} be the projectivizations, and we denote their dimensions by dn,kd_{n,k} and dn,kd^{\prime}_{n,k} respectively. We define χk,rn=χ(Sn,kLn,r)\chi^{n}_{k,r}=\chi(S_{n,k}\cap L_{n,r}) for ln,rSn,rl_{n,r}\in S^{\prime}_{n,r}. Let lnl_{n} denotes the dimension of VnV_{n}. The following Lemma shows that the local Euler obstruction of the dual orbits are the ‘same’ with the original orbits.

Lemma 2 (Duality).

For any recursive group orbits satisfying Assumption 1 and Assumption 3, we have

ei,j=e^ji,j=(1)ln+dn,kdn,kenk,n.e^{\prime}_{i,j}=\hat{e}_{j-i,j}=(-1)^{l_{n}+d^{\prime}_{n,k}-d_{n,k}}\cdot e_{n-k,n}\/.
Proof.

Recall from Proposition 2.7 that, denote ln=dimVnl_{n}=\dim V_{n} we have

ek,n=\displaystyle e^{\prime}_{k,n}= Eu𝒪¯n,k(𝒪n,n)=(1)ln1cMSn,k(1)\displaystyle Eu_{\bar{\mathcal{O}^{\prime}}_{n,k}}(\mathcal{O}^{\prime}_{n,n})=(-1)^{l_{n}-1}c_{M}^{S^{\prime}_{n,k}}(-1)
=\displaystyle= (1)ln1(1)ln+dn,kdn,kcMSn,nk(1)\displaystyle(-1)^{l_{n}-1}\cdot(-1)^{l_{n}+d^{\prime}_{n,k}-d_{n,k}}c_{M}^{S_{n,n-k}}(-1)
=\displaystyle= (1)ln+dn,kdn,kEu𝒪¯n,nk(𝒪n,n)\displaystyle(-1)^{l_{n}+d^{\prime}_{n,k}-d_{n,k}}Eu_{\bar{\mathcal{O}^{\prime}}_{n,n-k}}(\mathcal{O}_{n,n})
=\displaystyle= (1)ln+dn,kdn,kenk,n=e^nk,n\displaystyle(-1)^{l_{n}+d^{\prime}_{n,k}-d_{n,k}}\cdot e_{n-k,n}=\hat{e}_{n-k,n}

Corollary 2.

We consider recursive group orbits satisfying Assumption 1 and Assumption 3. Let χk,rn\chi^{n}_{k,r} denote the Euler characteristic χ(Sn,kLn,r)\chi(S_{n,k}\cap L_{n,r}) for any hyperplane ln,rSn,rl_{n,r}\in S^{\prime}_{n,r}. Then we have the following equivalence for each nn: they can be deduced from each other.

{χk,rn|k,r=0,1,,n1}\xlongleftrightarrowequivalent{ei,j|i,j=1,n}.\{\chi^{n}_{k,r}|k,r=0,1,\cdots,n-1\}\xlongleftrightarrow{\text{equivalent}}\{e_{i,j}|i,j=1,\cdots n\}\/.

In particular, the nnth level sectional Euler characteristics {χk,rn|k,r=0,1,,n1}\{\chi^{n}_{k,r}|k,r=0,1,\cdots,n-1\} control the n1n-1th level {χk,rn1|k,r=0,1,,n2}\{\chi^{n-1}_{k,r}|k,r=0,1,\cdots,n-2\}.

Proof.

We have shown how to obtain the sectional Euler characteristic χk,rn\chi^{n}_{k,r} from local Euler obstructions ek,re^{\prime}_{k,r} and Smk,r{Sm}_{k,r}. By the previous Duality Lemma and Theorem 3.1, both of them are completely determined by ek,re_{k,r}. Thus we have proved the right-to-left arrow, and we just need to show that we can compute ei,je_{i,j} from χk,rn\chi^{n}_{k,r}. Applying Ernström’s theorem we have

(e1,1e1,2e1,n10e2,2e2,n100en1,n1)[χk,rnχk,0n]k,r=(00e0,n10e0,n2e1,n1e0,1e1,2en2,n1).\left(\begin{array}[]{cccc}e_{1,1}&e_{1,2}&\cdots&e_{1,n-1}\\ 0&e_{2,2}&\cdots&e_{2,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ 0&0&\cdots&e_{n-1,n-1}\end{array}\right)\cdot[\chi^{n}_{k,r}-\chi^{n}_{k,0}]_{k,r}=-\left(\begin{array}[]{cccc}0&0&\cdots&e_{0,n-1}\\ 0&\cdots&e_{0,n-2}&e_{1,n-1}\\ \cdots&\cdots&\cdots&\cdots\\ e_{0,1}&e_{1,2}&\cdots&e_{n-2,n-1}\end{array}\right).

Here the two triangular matrices are invertible, thus the matrix [βk,rn:=χk,rnχk,0n]k,r[\beta^{n}_{k,r}:=\chi^{n}_{k,r}-\chi^{n}_{k,0}]_{k,r} is also invertible. Then for 1kn11\leq k\leq n-1 we have

(β1,1β2,1βn1,1β1,2β2,2βn1,2β1,n1β2,n1βn1,n1)(00ek,kek,k+1ek,n1)=(00e0,nke1,nk+1ek1,n1)\left(\begin{array}[]{cccc}\beta_{1,1}&\beta_{2,1}&\cdots&\beta_{n-1,1}\\ \beta_{1,2}&\beta_{2,2}&\cdots&\beta_{n-1,2}\\ \cdots&\cdots&\cdots&\cdots\\ \beta_{1,n-1}&\beta_{2,n-1}&\cdots&\beta_{n-1,n-1}\end{array}\right)\cdot\left(\begin{array}[]{c}0\\ \cdots\\ 0\\ e_{k,k}\\ e_{k,k+1}\\ \cdots\\ e_{k,n-1}\end{array}\right)=\left(\begin{array}[]{c}0\\ \cdots\\ 0\\ e_{0,n-k}\\ e_{1,n-k+1}\\ \cdots\\ e_{k-1,n-1}\end{array}\right)

This shows that if {ei,j}\{e_{i,j}\} are known for ij<ki\leq j<k, then ei,ke_{i,k} are uniquely determined by [βk,rn]k,r[\beta^{n}_{k,r}]_{k,r}. When k=1k=1, e0,j=1e_{0,j}=1. Thus by induction on kk we have proved that ei,je_{i,j} can be obtained from χk,rn\chi^{n}_{k,r}. ∎

5. Matrix Rank Loci

For the rest of the paper we discuss the matrix rank loci. By a matrix rank loci we mean the following three cases. Let MnM_{n}, MnM^{\wedge}_{n} and MnSM^{S}_{n} be the space of n×nn\times n ordinary, skew-symmetric and symmetric matrices over kk respectively. We know that Mn=(kn)2M_{n}=(k^{n})^{\otimes 2}, Mn=2knM_{n}^{\wedge}=\wedge^{2}k^{n} and MnS=Sym2knM_{n}^{S}=Sym^{2}k^{n} are of dimension n2n^{2}, (n+12)\binom{n+1}{2} and (n2)\binom{n}{2} respectively. The group GLn(k)×GLn(k)GL_{n}(k)\times GL_{n}(k) and GLn(k)GL_{n}(k) act on MnM_{n}, MnM^{\wedge}_{n} and MnSM^{S}_{n} by sending a matrix AA to PAQPAQ, PAPtPAP^{t} and PAPtPAP^{t}. For all three cases, there are only finitely orbits consisting of matrices of the same rank. For any 0kn0\leq k\leq n, we denote the orbits as

Σn,k:=\displaystyle\Sigma^{\circ}_{n,k}:= {AMn|rkA=nk};\displaystyle\{A\in M_{n}|\operatorname{rk}A=n-k\};
Σn,kS:=\displaystyle\Sigma^{S\circ}_{n,k}:= {AMnS|rkA=nk}.\displaystyle\{A\in M^{S}_{n}|\operatorname{rk}A=n-k\}\/.

Since a skew-symmetric matrix can only be of even rank, we denote

Σ2n+1,2k+1:={AM2n+1|rkA=2n2k};Σ2n,2k:={AM2n|rkA=2n2k}.\Sigma^{\wedge\circ}_{2n+1,2k+1}:=\{A\in M^{\wedge}_{2n+1}|\operatorname{rk}A=2n-2k\};\quad\Sigma^{\wedge\circ}_{2n,2k}:=\{A\in M^{\wedge}_{2n}|\operatorname{rk}A=2n-2k\}\/.

We call those orbits the ordinary, symmetric and skew-symmetric rank loci. We denote Σn,k\Sigma_{n,k}, Σn,k\Sigma^{\wedge}_{n,k} and Σn,kS\Sigma^{S}_{n,k} to be there closure.

Note that all three GLnGL_{n} actions contain kk^{*} multiplications, thus the actions pass to projectived spaces (Mn)\mathbb{P}(M_{n}), (Mn)\mathbb{P}(M^{\wedge}_{n}) and (MnS)\mathbb{P}(M^{S}_{n}). We denote the projectived orbits (and their closure) by τn,k\tau^{\circ}_{n,k}, τn,k\tau^{\wedge\circ}_{n,k} and τn,kS\tau^{S\circ}_{n,k} (and τn,k\tau_{n,k}, τn,k\tau^{\wedge}_{n,k} and τn,kS\tau^{S}_{n,k}) respectively. We list here some basic properties of rank loci, for details we refer to [35], [34], [14] and [22].

Proposition 5.1.

With the notations mentioned above.

  1. (1)

    The rank loci are reduced irreducible quasi-projective varieties.

  2. (2)

    ((Dimension)) dimΣn,k=n2k2\dim\Sigma_{n,k}=n^{2}-k^{2}, dimΣn,kS=(nk)(n+k+1)2\dim\Sigma^{S}_{n,k}=\frac{(n-k)(n+k+1)}{2} and dimΣn,k=(nk)(n+k1)2\dim\Sigma^{\wedge}_{n,k}=\frac{(n-k)(n+k-1)}{2} .

  3. (3)

    ((Singularity)) When 1kn11\leq k\leq n-1, Σn,k\Sigma_{n,k}, Σn,k\Sigma^{\wedge}_{n,k} and Σn,kS\Sigma^{S}_{n,k} are singular with singularity Σn,k+1\Sigma_{n,k+1}, Σn,k+2\Sigma^{\wedge}_{n,k+2} and Σn,k+1S\Sigma^{S}_{n,k+1} respectively. For k=0,nk=0,n they are smooth.

  4. (4)

    ((Product Structure)) For any AΣn,rA\in\Sigma_{n,r}, the normal slice NAN_{A} is isomorphic to MrM_{r}, and intersects Σn,r\Sigma_{n,r} at {A}=0\{A\}=0. Moreover the intersection Σn,kNA\Sigma_{n,k}\cap N_{A} is isomorphic to Σr,k\Sigma_{r,k}. This gives a local product structure (Σn,k,A)(Σr,k,0)×kN(\Sigma_{n,k},A)\cong(\Sigma_{r,k},0)\times k^{N}. The same property holds for Σn,k\Sigma^{\wedge}_{n,k} and Σn,kS\Sigma^{S}_{n,k}.

  5. (5)

    ((Duality)) For k=1,,n1k=1,\cdots,n-1, the dual variety of τn,k\tau_{n,k}, τn,kS\tau^{S}_{n,k} and τ2n+1,2k+1\tau^{\wedge}_{2n+1,2k+1} (or τ2n,2k)(\text{or }\tau^{\wedge}_{2n,2k}) are isomorphic to τn,nk\tau_{n,n-k} , τn,nkS\tau^{S}_{n,n-k} and τ2n+1,2n+12k\tau^{\wedge}_{2n+1,2n+1-2k} (or τ2n,2n2k)(\text{or }\tau^{\wedge}_{2n,2n-2k}) respectively.

  6. (6)

    For ordinary matrix and skew-symmetric matrices, we have

    χ(τn,k)={0k<n2(n2)k=n2;χ(τn,k)={0k<n1n2k=n1\displaystyle\chi(\tau^{\wedge\circ}_{n,k})=\begin{cases}0&k<n-2\\ \binom{n}{2}&k=n-2\end{cases};\quad\chi(\tau^{\circ}_{n,k})=\begin{cases}0&k<n-1\\ n^{2}&k=n-1\end{cases}
  7. (7)

    For symmetric matrices, we have the following slightly different result

    χn,kS:=χ(τn,kS)=\displaystyle\chi^{S}_{n,k}:=\chi(\tau^{S\circ}_{n,k})= {0k<n2(n2)k=n2(n1)k=n1\displaystyle\begin{cases}0&k<n-2\\ \binom{n}{2}&k=n-2\\ \binom{n}{1}&k=n-1\end{cases}

Most importantly, all three types rank loci admit natural resolutions of singularities: the Tjurina transforms defined as the incidence of Grassmannians. Let G(k,n)G(k,n) be the Grassmannian of kk-planes in V=knV=k^{n}, we denote SS and QQ to be the universal sub and quotient bundles. For ordinary rank loci τn,k\tau_{n,k}, the (projective) Tjurina transform is defined by

τ^n,k:={(Λ,A)|ΛkerA}(Mn)×G(k,n).\hat{\tau}_{n,k}:=\{(\Lambda,A)|\Lambda\subset\ker A\}\subset\mathbb{P}(M_{n})\times G(k,n)\/.

The skew-symmetric and symmetric Tjurina transforms are defined as

τ^n,k:=\displaystyle\hat{\tau}^{\wedge}_{n,k}:= {(Λ,X)|X|Λ=0}(Mn)×G(k,n);\displaystyle\{(\Lambda,X)|X|_{\Lambda}=0\}\subset\mathbb{P}(M^{\wedge}_{n})\times G(k,n);
τ^n,kS:=\displaystyle\hat{\tau}^{S}_{n,k}:= {(Λ,X)|X|Λ=0}(MnS)×G(k,n);\displaystyle\{(\Lambda,X)|X|_{\Lambda}=0\}\subset\mathbb{P}(M^{S}_{n})\times G(k,n);

For all three cases, set X=τX=\tau, X=τX=\tau^{\wedge} and X=τSX=\tau^{S}, and set N\mathbb{P}^{N} by (Mn)\mathbb{P}(M_{n}), (Mn)\mathbb{P}(M^{\wedge}_{n}) and (MnS)\mathbb{P}(M^{S}_{n}) we have commutative diagrams

X^n,k{\hat{X}_{n,k}}G(k,n)×N{G(k,n)\times\mathbb{P}^{N}}G(k,n){G(k,n)}Xn,k{X_{n,k}}N.{\mathbb{P}^{N}.}p\scriptstyle{p}q\scriptstyle{q}

The second projection pp is a resolution of singularity, and is isomorphic over τn,k\tau^{\circ}_{n,k}. The first projections qq identify the Tjurina transforms with projectivized bundles:

τ^n,k(Qn);τ^n,k(2Q);τ^n,kS(Sym2Q).\hat{\tau}_{n,k}\cong\mathbb{P}(Q^{\vee n});\quad\hat{\tau}^{\wedge}_{n,k}\cong\mathbb{P}(\wedge^{2}Q^{\vee});\quad\hat{\tau}^{S}_{n,k}\cong\mathbb{P}(Sym^{2}Q^{\vee})\/.

6. Local Euler Obstruction of Rank Loci

In this section we apply Theorem 4.1 to the case of matrix rank loci.

6.1. Ordinary Matrix

For ordinary matrix cells, the local Euler obstruction are computed over \mathbb{C} in [17] and over arbitrary algebraically closed field in [39]. In [17] the authors used recursive method based on the knowledge of Smn,i{Sm}_{n,i} computed in [21] via topology method. In [39] the formula came from direct computation via the knowledge of Nash blowup and the Nash bundle. Here we propose a proof to the formula for Smn,i{Sm}_{n,i} different with [21] and [39]. The proof is algebraic, thus works for arbitrary algebraically closed field of characteristic 0.

Proposition 6.1.

For ordinary matrix rank loci we have

Smn,i=(1)n+1k(nk).{Sm}_{n,i}=(-1)^{n+1-k}\binom{n}{k}\/.
Proof.

Recall the Tjurina diagram

τ^n,k{\hat{\tau}_{n,k}}G(k,n)×N{G(k,n)\times\mathbb{P}^{N}}G(k,n){G(k,n)}τn,k{\tau_{n,k}}N.{\mathbb{P}^{N}.}p\scriptstyle{p}q\scriptstyle{q}

In [40] we proved the following formula

csmτn,k(H)=i=kn1(1)ik(ik)pcsmτ^n,i(H).c_{sm}^{\tau_{n,k}^{\circ}}(H)=\sum_{i=k}^{n-1}(-1)^{i-k}\binom{i}{k}p_{*}c_{sm}^{\hat{\tau}_{n,i}}(H).

Thus it amounts to compute pcsmτ^n,i(1)p_{*}c_{sm}^{\hat{\tau}_{n,i}}(-1). Recall that τ^n,k\hat{\tau}_{n,k} is isomorphic to the projectivized bundle (Qn)\mathbb{P}(Q^{\vee n}). Let SS and QQ be the universal sub and quotient bundles on Grassmannian G(k,n)G(k,n), and let 𝒪(1)\mathcal{O}(1) be the tautological line bundle. We have

(1)n21pcsmτ^n,k(1)=(Qn)c(SQ)c(Qn𝒪(1))c(𝒪(1))[(Qn)](-1)^{n^{2}-1}p_{*}c_{sm}^{\hat{\tau}_{n,k}}(-1)=\int_{\mathbb{P}(Q^{\vee n})}\frac{c(S^{\vee}\otimes Q)c(Q^{\vee n}\otimes\mathcal{O}(1))}{c(\mathcal{O}(1))}\cap[\mathbb{P}(Q^{\vee n})]

By Lemma 1 we have

(1)n21pcsmτ^n,k(1)=\displaystyle(-1)^{n^{2}-1}p_{*}c_{sm}^{\hat{\tau}_{n,k}}(-1)= (Qn)c(SQ)c(Qn𝒪(1))c(𝒪(1))[(Qn)]\displaystyle\int_{\mathbb{P}(Q^{\vee n})}\frac{c(S^{\vee}\otimes Q)c(Q^{\vee n}\otimes\mathcal{O}(1))}{c(\mathcal{O}(1))}\cap[\mathbb{P}(Q^{\vee n})]
=\displaystyle= G(k,n)(1cn(nk)(Qn))c(Qn)))[G(k,n)]\displaystyle\int_{G(k,n)}\left(1-\frac{c_{n(n-k)}(Q^{n}))}{c(Q^{n}))}\right)\cap[G(k,n)]
=\displaystyle= G(k,n)c(TG(k,n))[G(k,n)]=(nk)\displaystyle\int_{G(k,n)}c(T_{G(k,n)})\cap[G(k,n)]=\binom{n}{k}

Thus we have

Smn,k=\displaystyle{Sm}_{n,k}= (1)n21csmτn,k(1)=i=kn1(1)ik(ik)(1)n21pcsmτ^n,i(1)\displaystyle(-1)^{n^{2}-1}c_{sm}^{\tau_{n,k}^{\circ}}(-1)=\sum_{i=k}^{n-1}(-1)^{i-k}\binom{i}{k}(-1)^{n^{2}-1}p_{*}c_{sm}^{\hat{\tau}_{n,i}}(-1)
=\displaystyle= i=kn1(1)ik(ni)(ik)=(nk)i=kn1(1)ik(nkik)\displaystyle\sum_{i=k}^{n-1}(-1)^{i-k}\binom{n}{i}\binom{i}{k}=\binom{n}{k}\cdot\sum_{i=k}^{n-1}(-1)^{i-k}\binom{n-k}{i-k}
=\displaystyle= (1)n+1k(nk)\displaystyle(-1)^{n+1-k}\binom{n}{k}

Follow from Theorem 3.1 we have:

Theorem 6.2.

The local Euler obstruction function of τn,k\tau_{n,k} is

ei,j=Euτn,i(τn,j)=(ji).e_{i,j}=Eu_{\tau_{n,i}}(\tau_{n,j}^{\circ})=\binom{j}{i}\/.
Proof.

We prove by inductions. For the initial case, we have ei,i=1e_{i,i}=1 for any ii. Assume that ei,j=(ji)e_{i,j}=\binom{j}{i} for all ij<ni\leq j<n. Then Proposition 2.4 we have

ek,n=\displaystyle e_{k,n}= r=kn1ek,rSmn,r=r=kn1(rk)(1)n+1r(nr)\displaystyle\sum_{r=k}^{n-1}e_{k,r}\cdot{Sm}_{n,r}=\sum_{r=k}^{n-1}\binom{r}{k}(-1)^{n+1-r}\binom{n}{r}
=\displaystyle= (nk)(1)j=0nk1(1)nkj(nkj)\displaystyle\binom{n}{k}\cdot(-1)\cdot\sum_{j=0}^{n-k-1}(-1)^{n-k-j}\binom{n-k}{j}
=\displaystyle= (nk)(1)(0nk(1)(nk)(nk))=(nk)\displaystyle\binom{n}{k}\cdot(-1)\cdot(0^{n-k}-(-1)^{(n-k)-(n-k)})=\binom{n}{k}

6.2. Skew-Symmetric Matrix

For skew-symmetric rank loci we have Vn=2knV_{n}=\wedge^{2}k^{n} and Gn=GLn(k)G_{n}=GL_{n}(k). we denote to orbits of of rank nkn-k matrices in 2kn\wedge^{2}k^{n} and (2kn)\mathbb{P}(\wedge^{2}k^{n}) by Σn,k\Sigma^{\wedge\circ}_{n,k} and τn,k\tau^{\wedge\circ}_{n,k} respectively. Here nk=0,2,4,n2n-k=0,2,4,\cdots\lfloor\frac{n}{2}\rfloor. Let Σn,k\Sigma_{n,k} and τn,k\tau_{n,k} be their closures. The local product structure shows that, for mnm\geq n we have

Euτn,k(τn,k+j)=Euτm,k(τm,k+j).Eu_{\tau^{\wedge}_{n,k}}(\tau^{\wedge\circ}_{n,k+j})=Eu_{\tau^{\wedge}_{m,k}}(\tau^{\wedge\circ}_{m,k+j})\/.

For complex skew-symmetric rank loci, the local Euler obstruction are computed in [34, Chapter 9], in which the author computed Smm,i{Sm}_{m,i} using topology method. Here we propose an algebraic formula which works for general kk of characteristic 0.

Proposition 6.3.

The csmc_{sm} invariants are given by

Sm2n,2k=Sm2n+1,2k+1=(1)nk+1(nk).{Sm}_{2n,2k}={Sm}_{2n+1,2k+1}=(-1)^{n-k+1}\binom{n}{k}\/.
Proof.

Recall the Tjurina diagram for skew-symmetric rank loci.

τ^m,k{\hat{\tau}^{\wedge}_{m,k}}G(k,n)×(2km){G(k,n)\times\mathbb{P}(\wedge^{2}k^{m})}G(k,n){G(k,n)}τm,k{\tau^{\wedge}_{m,k}}(2km).{\mathbb{P}(\wedge^{2}k^{m}).}p\scriptstyle{p}q\scriptstyle{q}

Here τ^m,k\hat{\tau}^{\wedge}_{m,k} is isomorphic to the projective bundle (2Q)\mathbb{P}(\wedge^{2}Q^{\vee}). Thus we have

(1)(m2)1p(csmτ^m,k)(1)=(2Q)c(SQ)c(2Q)𝒪(1))c(𝒪(1))[(2Q)](-1)^{\binom{m}{2}-1}p_{*}(c_{sm}^{\hat{\tau}^{\wedge}_{m,k}})(-1)=\int_{\mathbb{P}(\wedge^{2}Q^{\vee})}\frac{c(S^{\vee}\otimes Q)c(\wedge^{2}Q^{\vee})\otimes\mathcal{O}(1))}{c(\mathcal{O}(1))}\cap[\mathbb{P}(\wedge^{2}Q^{\vee})]

By Lemma 1 we have

(1)(m2)1p(csmτ^m,k)(1)=\displaystyle(-1)^{\binom{m}{2}-1}p_{*}(c_{sm}^{\hat{\tau}^{\wedge}_{m,k}})(-1)= G(k,m)c(SQ)(1ctop(2Q)c(2Q))[G(k,m)]\displaystyle\int_{G(k,m)}c(S^{\vee}\otimes Q)\cdot\left(1-\frac{c_{top}(\wedge^{2}Q)}{c(\wedge^{2}Q)}\right)\cap[G(k,m)]
=\displaystyle= (mk)G(k,m)c(SQ)ctop(2Q)c(2Q)[G(k,m)]\displaystyle\binom{m}{k}-\int_{G(k,m)}\frac{c(S^{\vee}\otimes Q)c_{top}(\wedge^{2}Q)}{c(\wedge^{2}Q)}\cap[G(k,m)]

For any Aτm,rA\in\tau^{\wedge\circ}_{m,r}, the fiber p1(A)p^{-1}(A) is isomorphic to the Grassmannian G(k,r)G(k,r). Notice that AA can only have even degree, we need to consider m=2nm=2n and m=2n+1m=2n+1 cases separately. First we consider the even case. The pushforward of constructible functions shows that

(1)(2n2)1p(csmτ^2n,2k)(1)=\displaystyle(-1)^{\binom{2n}{2}-1}p_{*}(c_{sm}^{\hat{\tau}^{\wedge}_{2n,2k}})(-1)= r=kn1(2r2k)(1)(2n2)1csmτ2n,2r(1)\displaystyle\sum_{r=k}^{n-1}\binom{2r}{2k}(-1)^{\binom{2n}{2}-1}c_{sm}^{\tau^{\wedge\circ}_{2n,2r}}(-1)
=\displaystyle= r=kn1(2r2k)Sm2n,2r.\displaystyle\sum_{r=k}^{n-1}\binom{2r}{2k}{Sm}_{2n,2r}\/.

When m=2n+1m=2n+1, we have

(1)(2n+12)1p(csmτ^2n+1,2k+1)(1)=r=kn1(2r+12k+1)Sm2n+1,2r+1.(-1)^{\binom{2n+1}{2}-1}p_{*}(c_{sm}^{\hat{\tau}^{\wedge}_{2n+1,2k+1}})(-1)=\sum_{r=k}^{n-1}\binom{2r+1}{2k+1}{Sm}_{2n+1,2r+1}\/.

Notice that the matrices [(2r2k)]1r,kn1[\binom{2r}{2k}]_{1\leq r,k\leq n-1} and [(2r+12k+1)]1r,kn1[\binom{2r+1}{2k+1}]_{1\leq r,k\leq n-1} are invertible, thus it amounts to prove the following Schubert identities:

G(2k,2n)c(SQ)c(2n2k2)(2Q)c(2Q)[G(2k,2n)]=\displaystyle\int_{G(2k,2n)}\frac{c(S^{\vee}\otimes Q)c_{\binom{2n-2k}{2}}(\wedge^{2}Q)}{c(\wedge^{2}Q)}\cap[G(2k,2n)]= r=kn(1)nk(2r2k)(nr)\displaystyle\sum_{r=k}^{n}(-1)^{n-k}\binom{2r}{2k}\binom{n}{r}
G(2k+1,2n+1)c(SQ)c(2n2k2)(2Q)c(2Q)[G(2k+1,2n+1)]=\displaystyle\int_{G(2k+1,2n+1)}\frac{c(S^{\vee}\otimes Q)c_{\binom{2n-2k}{2}}(\wedge^{2}Q)}{c(\wedge^{2}Q)}\cap[G(2k+1,2n+1)]= r=kn(1)nk(2r+12k+1)(nr)\displaystyle\sum_{r=k}^{n}(-1)^{n-k}\binom{2r+1}{2k+1}\binom{n}{r}

When the base field is \mathbb{C}, this was proved in [34, Chapter 9]. Since the equations are nothing but binomial identities, thus they hold for arbitrary field kk of charcteristic 0. ∎

Mimicking the proof for the ordinary rank loci case we have:

Theorem 6.4.

The local Euler obstructions of skew-symmetric rank loci are

Euτ2n,2k(τ2n,2r)=Euτ2n+1,2k+1(τ2n+1,2r+1)=(rk).Eu_{\tau^{\wedge}_{2n,2k}}(\tau^{\wedge\circ}_{2n,2r})=Eu_{\tau^{\wedge}_{2n+1,2k+1}}(\tau^{\wedge\circ}_{2n+1,2r+1})=\binom{r}{k}\/.

6.3. Symmetric Matrix

Unlike the ordinary and skew-symmetric case, the Euler obstructions of complex symmetric rank loci are unknown. Thus we compute them here.

In this section we have Vn=Sym2knV_{n}=Sym^{2}k^{n} to be the space of all symmetric n×nn\times n matrices, and Gn=GLn(k)G_{n}=GL_{n}(k) acts on VnV_{n} by PAPtPAP^{t}. The group orbits consist of symmetric matrices of rank nkn-k for k=0,1,n1k=0,1,\cdots n-1, and are denoted by Σn,kS\Sigma^{S\circ}_{n,k}. Let τn,kS\tau^{S\circ}_{n,k} be their projectivizations in (Sym2kn)\mathbb{P}(Sym^{2}k^{n}). Let Σn,kS\Sigma^{S}_{n,k} and τn,kS\tau^{S}_{n,k} be their closures. The product structure shows that

EuΣn,kS(Σn,k+jS)=EuΣm,kS(Σm,k+jS)=Euτn,kS(τn,k+jS).Eu_{\Sigma^{S}_{n,k}}(\Sigma^{S\circ}_{n,k+j})=Eu_{\Sigma^{S}_{m,k}}(\Sigma^{S\circ}_{m,k+j})=Eu_{\tau^{S}_{n,k}}(\tau^{S\circ}_{n,k+j})\/.

for k+j<nk+j<n. First we compute the csmc_{sm} indices Smn,k{Sm}_{n,k}.

Proposition 6.5.

For any 1km11\leq k\leq m-1, we have the following formula

Smm,k:=(1)(m+12)1csmτm,kS(1)={(1)ni+1(ni)k=2i,m=2n0k=2i+1,m=2n(1)ni(ni)k=2i,m=2n+1(1)ni+1(ni)k=2i+1,m=2n+1\displaystyle{Sm}_{m,k}:=(-1)^{\binom{m+1}{2}-1}\cdot c_{sm}^{\tau^{S\circ}_{m,k}}(-1)=\begin{cases}(-1)^{n-i+1}\cdot\binom{n}{i}&k=2i,m=2n\\ 0&k=2i+1,m=2n\\ (-1)^{n-i}\binom{n}{i}&k=2i,m=2n+1\\ (-1)^{n-i+1}\binom{n}{i}&k=2i+1,m=2n+1\end{cases}
Proof.

Consider the Tjurina transform diagram of symmetric rank loci:

τ^n,kS{\hat{\tau}^{S}_{n,k}}G(k,n)×(Sym2kn){G(k,n)\times\mathbb{P}(Sym^{2}k^{n})}G(k,n){G(k,n)}τn,kS{\tau^{S}_{n,k}}(Sym2kn).{\mathbb{P}(Sym^{2}k^{n})\/.}p\scriptstyle{p}q\scriptstyle{q}

Here τ^n,kS\hat{\tau}^{S}_{n,k} is isomorphic to the projective bundle (Sym2Q)\mathbb{P}(Sym^{2}Q^{\vee}) over G(k,n)G(k,n), for QQ being the universal quotient bundle. Notice that for any Aτm,k+jSA\in\tau^{S\circ}_{m,k+j}, its fiber p1(A)p^{-1}(A) is isomorphic to G(k,k+j)G(k,k+j). Thus we have

pcsmτ^m,kS(H)=j=km1(jk)csmτ^m,jS(H)p_{*}c_{sm}^{\hat{\tau}^{S}_{m,k}}(H)=\sum_{j=k}^{m-1}\binom{j}{k}c_{sm}^{\hat{\tau}^{S\circ}_{m,j}}(H)

Following the same argument in ordinary case we have

csmτm,kS(H)=j=km1(1)jk(jk)pcsmτ^m,jS(H).c_{sm}^{\tau^{S\circ}_{m,k}}(H)=\sum_{j=k}^{m-1}(-1)^{j-k}\binom{j}{k}p_{*}c_{sm}^{\hat{\tau}^{S}_{m,j}}(H).

Thus it amounts to compute the indices pcsmτ^m,k+iS(1)p_{*}c_{sm}^{\hat{\tau}^{S}_{m,k+i}}(-1). Since τ^m,kS(Sym2Q)\hat{\tau}^{S}_{m,k}\cong\mathbb{P}(Sym^{2}Q^{\vee}), let 𝒪(1)\mathcal{O}(1) be the tautological line bundle we have

(1)(m+12)1pcsmτ^m,kS(1)=(Sym2Q)c(SQ)c(Sym2Q)𝒪(1))c(𝒪(1))[(Sym2Q)](-1)^{\binom{m+1}{2}-1}p_{*}c_{sm}^{\hat{\tau}^{S}_{m,k}}(-1)=\int_{\mathbb{P}(Sym^{2}Q^{\vee})}\frac{c(S^{\vee}\otimes Q)c(Sym^{2}Q^{\vee})\otimes\mathcal{O}(1))}{c(\mathcal{O}(1))}\cap[\mathbb{P}(Sym^{2}Q^{\vee})]

By Lemma 1 we have

(1)(m+12)1pcsmτ^n,k(1)=\displaystyle(-1)^{\binom{m+1}{2}-1}p_{*}c_{sm}^{\hat{\tau}_{n,k}}(-1)= (Sym2Q)c(SQ)c(Sym2Q𝒪(1))c(𝒪(1))[(Sym2Q)]\displaystyle\int_{\mathbb{P}(Sym^{2}Q^{\vee})}\frac{c(S^{\vee}\otimes Q)c(Sym^{2}Q^{\vee}\otimes\mathcal{O}(1))}{c(\mathcal{O}(1))}\cap[\mathbb{P}(Sym^{2}Q^{\vee})]
=\displaystyle= G(k,n)c(SQ)(1ctop(Sym2Q)c(Sym2Q))[G(k,n)]\displaystyle\int_{G(k,n)}c(S^{\vee}\otimes Q)\cdot\left(1-\frac{c_{top}(Sym^{2}Q)}{c(Sym^{2}Q)}\right)\cap[G(k,n)]
=\displaystyle= (nk)G(k,n)c(SQ)ctop(Sym2Q)c(Sym2Q)[G(k,n)]\displaystyle\binom{n}{k}-\int_{G(k,n)}\frac{c(S^{\vee}\otimes Q)c_{top}(Sym^{2}Q)}{c(Sym^{2}Q)}\cap[G(k,n)]
Lemma 3.

We have the following Schubert formula.

G(k,2n)c(SQ)ctop(Sym2Q)c(Sym2Q)=\displaystyle\int_{G(k,2n)}\frac{c(S^{\vee}\otimes Q)c_{top}(Sym^{2}Q)}{c(Sym^{2}Q)}= r=knn(nr)(rkn)=22nk(nkn)\displaystyle\sum_{r=k-n}^{n}\binom{n}{r}\binom{r}{k-n}=2^{2n-k}\cdot\binom{n}{k-n}
G(k,2n+1)c(SQ)ctop(Sym2Q)c(Sym2Q)=\displaystyle\int_{G(k,2n+1)}\frac{c(S^{\vee}\otimes Q)c_{top}(Sym^{2}Q)}{c(Sym^{2}Q)}= r=kn1n(nr)(rkn1)=22nk+1(nkn1).\displaystyle\sum_{r=k-n-1}^{n}\binom{n}{r}\binom{r}{k-n-1}=2^{2n-k+1}\cdot\binom{n}{k-n-1}\/.
Proof of Lemma.

Recall the standard duality isomorphism

ϕ:G1:=G(k,n)=G(k,V)G(nk,n)=G(nk,V)=:G2.\phi\colon G_{1}:=G(k,n)=G(k,V)\to G(n-k,n)=G(n-k,V^{\vee})=:G_{2}\/.

For i=1,2i=1,2 we denote Si,QiS_{i},Q_{i} to be the universal sub and quotient bundles. Then ϕS2=Q1\phi^{*}S_{2}=Q_{1}^{\vee} and ϕQ2=S1\phi^{*}Q_{2}=S_{1}^{\vee}. Thus we have

G(k,2n)c(SQ)ctop(Sym2Q)c(Sym2Q)=G(2nk,2n)c(QS)ctop(Sym2S)c(Sym2S).\int_{G(k,2n)}\frac{c(S^{\vee}\otimes Q)c_{top}(Sym^{2}Q)}{c(Sym^{2}Q)}=\int_{G(2n-k,2n)}\frac{c(Q\otimes S^{\vee})c_{top}(Sym^{2}S^{\vee})}{c(Sym^{2}S^{\vee})}\/.

Thus it amounts to prove that

G(2nk,2n)c(QS)ctop(Sym2S)c(Sym2S)\displaystyle\int_{G(2n-k,2n)}\frac{c(Q\otimes S^{\vee})c_{top}(Sym^{2}S^{\vee})}{c(Sym^{2}S^{\vee})} =22nk(nkn)\displaystyle=2^{2n-k}\cdot\binom{n}{k-n}
G(2nk+1,2n+1)c(QS)ctop(Sym2S)c(Sym2S)\displaystyle\int_{G(2n-k+1,2n+1)}\frac{c(Q\otimes S^{\vee})c_{top}(Sym^{2}S^{\vee})}{c(Sym^{2}S^{\vee})} =22nk+1(nkn1).\displaystyle=2^{2n-k+1}\cdot\binom{n}{k-n-1}\/.

We leave the proof of above Schubert identities to Corollary 4, where we identify such Schubert integrations as the Euler characteristics of certain moduli spaces, and compute them using geometry method. ∎

Thus from the Lemma we have

(1)(m+12)1pcsmτ^2n,kS(1)=\displaystyle(-1)^{\binom{m+1}{2}-1}p_{*}c_{sm}^{\hat{\tau}^{S}_{2n,k}}(-1)= (2nk)22nk(nkn)\displaystyle\binom{2n}{k}-2^{2n-k}\cdot\binom{n}{k-n}
(1)(m+12)1pcsmτ^2n+1,kS(1)=\displaystyle(-1)^{\binom{m+1}{2}-1}p_{*}c_{sm}^{\hat{\tau}^{S}_{2n+1,k}}(-1)= (2n+1k)22nk+1(nkn1).\displaystyle\binom{2n+1}{k}-2^{2n-k+1}\cdot\binom{n}{k-n-1}\/.

Here we make the convention that (nk)=0\binom{n}{k}=0 whenever k<0k<0.

Then for m=2nm=2n we have

Sm2n,k=\displaystyle{Sm}_{2n,k}= j=k2n1(1)jk(jk)pcsmτ^2n,j(1)\displaystyle\sum_{j=k}^{2n-1}(-1)^{j-k}\binom{j}{k}p_{*}c_{sm}^{\hat{\tau}_{2n,j}}(-1)
=\displaystyle= j=k2n1(1)jk(jk)((2nj)22nj(njn))\displaystyle\sum_{j=k}^{2n-1}(-1)^{j-k}\binom{j}{k}\left(\binom{2n}{j}-2^{2n-j}\cdot\binom{n}{j-n}\right)
=\displaystyle= {(1)ni+1(ni)k=2i0k=2i+1.\displaystyle\begin{cases}(-1)^{n-i+1}\cdot\binom{n}{i}&k=2i\\ 0&k=2i+1\end{cases}\/.

For m=2n+1m=2n+1 we have

Sm2n+1,k=\displaystyle{Sm}_{2n+1,k}= j=k2n(1)jk(jk)pcsmτ^2n+1,j(1)\displaystyle\sum_{j=k}^{2n}(-1)^{j-k}\binom{j}{k}p_{*}c_{sm}^{\hat{\tau}_{2n+1,j}}(-1)
=\displaystyle= j=k2n(1)jk(jk)((2n+1j)22n+1j(njn1))\displaystyle\sum_{j=k}^{2n}(-1)^{j-k}\binom{j}{k}\left(\binom{2n+1}{j}-2^{2n+1-j}\cdot\binom{n}{j-n-1}\right)
=\displaystyle= {(1)ni(ni)k=2i(1)ni+1(ni)k=2i+1.\displaystyle\begin{cases}(-1)^{n-i}\binom{n}{i}&k=2i\\ (-1)^{n-i+1}\binom{n}{i}&k=2i+1\end{cases}\/.

The last steps are binomial identities that can be proved by induction. We leave the details to the readers. ∎

Now we proceed to compute the local Euler obstructions.

Theorem 6.6.

Let ek,ne_{k,n} be the local Euler obstruction of Σn,kS\Sigma^{S}_{n,k} at 0. Then we have

ek,m={(ni)k=2i,m=2n0k=2i+1,m=2n(ni)k=2i,m=2n+1(ni)k=2i+1,m=2n+1\displaystyle e_{k,m}=\begin{cases}\binom{n}{i}&k=2i,m=2n\\ 0&k=2i+1,m=2n\\ \binom{n}{i}&k=2i,m=2n+1\\ \binom{n}{i}&k=2i+1,m=2n+1\\ \end{cases}
Proof.

We prove by induction on nn. When n=1n=1, Proposition 2.8 shows that

e1,2=\displaystyle e_{1,2}= e1,1Sm2,1=0\displaystyle e_{1,1}\cdot{Sm}_{2,1}=0
e1,3=\displaystyle e_{1,3}= e1,2Sm3,2+e1,1Sm3,1=01+11=1=(10)\displaystyle e_{1,2}\cdot{Sm}_{3,2}+e_{1,1}\cdot{Sm}_{3,1}=0\cdot 1+1\cdot 1=1=\binom{1}{0}
e2,3=\displaystyle e_{2,3}= e2,2Sm3,2=11=(11)\displaystyle e_{2,2}\cdot{Sm}_{3,2}=1\cdot 1=\binom{1}{1}

Assume that the theorem holds true for all nNn\leq N and all k2N+1k\leq 2N+1. For n=N+1n=N+1, combine Proposition 2.8 and Proposition 6.5 we have

e2k+1,2n=\displaystyle e_{2k+1,2n}= j=2k+12n1e2k+1,jSm2n,j\displaystyle\sum_{j=2k+1}^{2n-1}e_{2k+1,j}\cdot{Sm}_{2n,j}
=\displaystyle= j=k+1n1e2k+1,2jSm2n,2j+j=kn1e2k+1,2j+1Sm2n,2j+1\displaystyle\sum_{j=k+1}^{n-1}e_{2k+1,2j}\cdot{Sm}_{2n,2j}+\sum_{j=k}^{n-1}e_{2k+1,2j+1}\cdot{Sm}_{2n,2j+1}
=\displaystyle= j=k+1n10Sm2n,2j+j=kn1e2k+1,2j+10\displaystyle\sum_{j=k+1}^{n-1}0\cdot{Sm}_{2n,2j}+\sum_{j=k}^{n-1}e_{2k+1,2j+1}\cdot 0
=\displaystyle= 0\displaystyle 0
e2k,2n=\displaystyle e_{2k,2n}= j=2k2n1e2k,jSm2n,j\displaystyle\sum_{j=2k}^{2n-1}e_{2k,j}\cdot{Sm}_{2n,j}
=\displaystyle= j=kn1e2k,2jSm2n,2j+j=kn1e2k,2j+1Sm2n,2j+1\displaystyle\sum_{j=k}^{n-1}e_{2k,2j}\cdot{Sm}_{2n,2j}+\sum_{j=k}^{n-1}e_{2k,2j+1}\cdot{Sm}_{2n,2j+1}
=\displaystyle= j=kn1e2k,2jSm2n,2j=j=kn1(1)jk+1(jk)(nj)\displaystyle\sum_{j=k}^{n-1}e_{2k,2j}\cdot{Sm}_{2n,2j}=\sum_{j=k}^{n-1}(-1)^{j-k+1}\cdot\binom{j}{k}\binom{n}{j}
=\displaystyle= j=kn1(1)nj+1(jk)(nj)=(nk)j=kn1(1)jk+1(nkjk)\displaystyle\sum_{j=k}^{n-1}(-1)^{n-j+1}\binom{j}{k}\cdot\binom{n}{j}=\binom{n}{k}\cdot\sum_{j=k}^{n-1}(-1)^{j-k+1}\binom{n-k}{j-k}
=\displaystyle= (nk)\displaystyle\binom{n}{k}

The proof for e2k,2n+1=e2k+1,2n+1=(nk)e_{2k,2n+1}=e_{2k+1,2n+1}=\binom{n}{k} are exactly the same. Thus we have proved the induction step, and hence the theorem. ∎

7. Sectional Euler Characteristic of Rank Loci

In this section we apply Theorem 2.3 to compute the sectional Euler characteristic of matrix rank loci.

7.1. Ordinary Matrix

Recall that the dual variety of τn,kn21\tau_{n,k}\subset\mathbb{P}^{n^{2}-1} is exactly τn,nkn21\tau_{n,n-k}\subset\mathbb{P}^{*n^{2}-1}, with dimensions d=dimτn,k=(n+k)(nk)1d=\dim\tau_{n,k}=(n+k)(n-k)-1 and d=dimτn,nk=(2nk)k1d^{*}=\dim\tau_{n,n-k}=(2n-k)k-1. We have the following.

Theorem 7.1.

Let LL be a hyperplane correspond to lτn,rn21l\in\tau_{n,r}^{\circ}\in\mathbb{P}^{{}^{*}n^{2}-1}. Then we have

χ(τn,kL)={i=kn(1)ik(ik)(rni)k<n1rn+n21k=n1.\chi(\tau_{n,k}^{\circ}\cap L)=\begin{cases}\sum_{i=k}^{n}(-1)^{i-k}\binom{i}{k}\binom{r}{n-i}&k<n-1\\ r-n+n^{2}-1&k=n-1\end{cases}\/.

Here we make the convention that (ab)=0\binom{a}{b}=0 whenever b<0b<0.

Proof.

Theorem 2.5 shows that

(Euτn,k)=Euτn,nk+e𝟙n21(Eu_{\tau_{n,k}})^{\vee}=Eu_{\tau_{n,n-k}}+e\mathds{1}_{\mathbb{P}^{n^{2}-1}}

Thus for any rr we have

j=0n1Euτn,k(τn,j)(χk,rnχk,0n)=Euτn,nk(τn,r).\sum_{j=0}^{n-1}Eu_{\tau_{n,k}}(\tau_{n,j}^{\circ})\left(\chi^{n}_{k,r}-\chi^{n}_{k,0}\right)=Eu_{\tau_{n,n-k}}(\tau_{n,r}^{\circ})\/.

Recall that Euτn,k(τn,j)=ek,j=(jk)Eu_{\tau_{n,k}}(\tau_{n,j}^{\circ})=e_{k,j}=\binom{j}{k}. Let βk,rn:=(χk,snχk,0n)\beta^{n}_{k,r}:=\left(\chi^{n}_{k,s}-\chi^{n}_{k,0}\right) we have

((11)(21)(31)(n11)0(22)(32)(n12)000(n1n1))(β1,1nβ1,n1nβ2,1nβ2,n1nβn1,1nβn1,n1n)=(000(n1n1)00(n2n2)(n1n2)(11)(21)(31)(n11)).\left(\begin{array}[]{ccccc}\binom{1}{1}&\binom{2}{1}&\binom{3}{1}&\cdots&\binom{n-1}{1}\\ 0&\binom{2}{2}&\binom{3}{2}&\cdots&\binom{n-1}{2}\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ 0&0&0&\cdots&\binom{n-1}{n-1}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{n}_{1,1}&\cdots&\beta^{n}_{1,n-1}\\ \beta^{n}_{2,1}&\cdots&\beta^{n}_{2,n-1}\\ \cdots&\cdots&\cdots\\ \beta^{n}_{n-1,1}&\cdots&\beta^{n}_{n-1,n-1}\end{array}\right)=\left(\begin{array}[]{ccccc}0&0&\cdots&0&\binom{n-1}{n-1}\\ 0&0&\cdots&\binom{n-2}{n-2}&\binom{n-1}{n-2}\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ \binom{1}{1}&\binom{2}{1}&\binom{3}{1}&\cdots&\binom{n-1}{1}\\ \end{array}\right).

Since the inverse matrix of [(ij)]1i,jn1[\binom{i}{j}]_{1\leq i,j\leq n-1} is [(1)ji(ij)]1i,jn1[(-1)^{j-i}\binom{i}{j}]_{1\leq i,j\leq n-1}, we then have

βk,rn=i=kn1(1)ik(ik)(rni)\displaystyle\beta^{n}_{k,r}=\sum_{i=k}^{n-1}(-1)^{i-k}\binom{i}{k}\binom{r}{n-i}

Recall that Proposition 2.4 shows:

χk,rn=βk,rn+χk,0n=βk,rn+χ(τn,k)Smn,k.\chi^{n}_{k,r}=\beta^{n}_{k,r}+\chi^{n}_{k,0}=\beta^{n}_{k,r}+\chi(\tau_{n,k}^{\circ})-{Sm}_{n,k}\/.

From Proposition 6.1 we have Smn,k=(1)n+1k(nk){Sm}_{n,k}=(-1)^{n+1-k}\binom{n}{k}, and hence

χk,rn=\displaystyle\chi^{n}_{k,r}= i=kn1(1)ik(ik)(rni)+(1)nk(nk)+χ(τn,k)\displaystyle\sum_{i=k}^{n-1}(-1)^{i-k}\binom{i}{k}\binom{r}{n-i}+(-1)^{n-k}\binom{n}{k}+\chi(\tau_{n,k}^{\circ})
=\displaystyle= {i=kn1(1)ik(ik)(rni)+(1)nk(nk)k<n1rn+n21k=n1\displaystyle\begin{cases}\sum_{i=k}^{n-1}(-1)^{i-k}\binom{i}{k}\binom{r}{n-i}+(-1)^{n-k}\binom{n}{k}&k<n-1\\ r-n+n^{2}-1&k=n-1\end{cases}

7.2. Skew-Symmetric Matrix

Now we consider skew-symmetric case. When m=2n+1m=2n+1, the total space (2k2n+1)\mathbb{P}(\wedge^{2}k^{2n+1}) is of dimension n(2n+1)1n(2n+1)-1. The rank loci τ2n+1,2k+1\tau^{\wedge}_{2n+1,2k+1} is dual to τ2n+1,2n2k1\tau^{\wedge}_{2n+1,2n-2k-1}, and they are of dimensions (nk)(2n+2k+1)1(n-k)(2n+2k+1)-1 and (k+1)(4n2k1)1(k+1)(4n-2k-1)-1 respectively. When m=2nm=2n is even, the total space (2k2n)\mathbb{P}(\wedge^{2}k^{2n}) is of dimension n(2n1)1n(2n-1)-1. The rank loci τ2n,2k\tau^{\wedge}_{2n,2k} is dual to τ2n,2n2k\tau^{\wedge}_{2n,2n-2k}, and they are of dimensions (nk)(2n+2k1)1(n-k)(2n+2k-1)-1 and (k+1)(4n2k1)1(k+1)(4n-2k-1)-1. Thus we have the following:

Theorem 7.2.

Let LBL_{B} be a hyperplane correspond to lτm,B(m2)1l\in\tau_{m,B}^{\circ}\in\mathbb{P}^{{}^{*}\binom{m}{2}-1}. Then we have

χ(τ2n,2kL2r)=\displaystyle\chi(\tau^{\wedge\circ}_{2n,2k}\cap L_{2r})= χ(τ2n+1,2k+1L2r+1)\displaystyle\chi(\tau^{\wedge\circ}_{2n+1,2k+1}\cap L_{2r+1})
=\displaystyle= {i=kn1(1)ik+1(ik)(rni)+(1)nk(nk)k<n12n21rk=n1\displaystyle\begin{cases}\sum_{i=k}^{n-1}(-1)^{i-k+1}\binom{i}{k}\binom{r}{n-i}+(-1)^{n-k}\binom{n}{k}&k<n-1\\ 2n^{2}-1-r&k=n-1\end{cases}

Here we make the convention that (ab)=0\binom{a}{b}=0 whenever b<0b<0.

Proof.

First we consider the odd case m=2n+1m=2n+1. Theorem 2.5 shows that

(Euτ2n+1,2k+1)=(1)Euτ2n+1,2n2k1+e𝟙(2k2n+1)(Eu_{\tau^{\wedge}_{2n+1,2k+1}})^{\vee}=(-1)\cdot Eu_{\tau^{\wedge}_{2n+1,2n-2k-1}}+e\mathds{1}_{\mathbb{P}(\wedge^{2}k^{2n+1})}

Thus for 1k,sn11\leq k,s\leq n-1 we have

r=kn1Euτ2n+1,2k+1(τ2n+1,2r+1)(χ2r+1,2s+12n+1χ2r+1,12n+1)=Euτ2n+1,2n2k1(τ2n+1,2s+1).\sum_{r=k}^{n-1}Eu_{\tau^{\wedge}_{2n+1,2k+1}}(\tau^{\wedge\circ}_{2n+1,2r+1})\left(\chi^{2n+1}_{2r+1,2s+1}-\chi^{2n+1}_{2r+1,1}\right)=-Eu_{\tau^{\wedge}_{2n+1,2n-2k-1}}(\tau^{\wedge\circ}_{2n+1,2s+1})\/.

Define β2k+1,2s+12n+1=χ2r+1,2s+12n+1χ2r+1,12n+1\beta^{2n+1}_{2k+1,2s+1}=\chi^{2n+1}_{2r+1,2s+1}-\chi^{2n+1}_{2r+1,1}. Recall from Theorem 6.4 that

Euτ2n+1,2k+1(τ2n+1,2r+1)=(rk).Eu_{\tau^{\wedge}_{2n+1,2k+1}}(\tau^{\wedge\circ}_{2n+1,2r+1})=\binom{r}{k}\/.

Then we have

((11)(21)(31)(n10)0(22)(32)(n12)000(n1n1))(β3,12n+1β3,2n12n+1β2n1,12n+1β2n1,2n12n+1)=(000(n1n1)00(n2n2)(n2n1)(11)(21)(31)(n10)).\left(\begin{array}[]{ccccc}\binom{1}{1}&\binom{2}{1}&\binom{3}{1}&\cdots&\binom{n-1}{0}\\ 0&\binom{2}{2}&\binom{3}{2}&\cdots&\binom{n-1}{2}\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ 0&0&0&\cdots&\binom{n-1}{n-1}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{2n+1}_{3,1}&\cdots&\beta^{2n+1}_{3,2n-1}\\ \cdots&\cdots&\cdots\\ \beta^{2n+1}_{2n-1,1}&\cdots&\beta^{2n+1}_{2n-1,2n-1}\end{array}\right)=-\left(\begin{array}[]{ccccc}0&0&\cdots&0&\binom{n-1}{n-1}\\ 0&0&\cdots&\binom{n-2}{n-2}&\binom{n-2}{n-1}\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ \binom{1}{1}&\binom{2}{1}&\binom{3}{1}&\cdots&\binom{n-1}{0}\end{array}\right).

Since the inverse matrix of [(ij)]1i,jn1[\binom{i}{j}]_{1\leq i,j\leq n-1} is [(1)ji(ij)]1i,jn1[(-1)^{j-i}\binom{i}{j}]_{1\leq i,j\leq n-1}, we then have

β2k+1,2r+12n+1=i=kn1(1)ik+1(ik)(rni)\displaystyle\beta^{2n+1}_{2k+1,2r+1}=\sum_{i=k}^{n-1}(-1)^{i-k+1}\binom{i}{k}\binom{r}{n-i}

Thus

χ2k+1,2r+12n+1=\displaystyle\chi^{2n+1}_{2k+1,2r+1}= β2k+1,2r+12n+1+χ2k+1,12n+1\displaystyle\beta^{2n+1}_{2k+1,2r+1}+\chi^{2n+1}_{2k+1,1}
=\displaystyle= i=kn1(1)ik+1(ik)(rni)+χ2k+1,12n+1\displaystyle\sum_{i=k}^{n-1}(-1)^{i-k+1}\binom{i}{k}\binom{r}{n-i}+\chi^{2n+1}_{2k+1,1}
=\displaystyle= i=kn1(1)ik+1(ik)(rni)+χτ2n+1,2k+1Sm2n+1,2k+1\displaystyle\sum_{i=k}^{n-1}(-1)^{i-k+1}\binom{i}{k}\binom{r}{n-i}+\chi_{\tau^{\wedge\circ}_{2n+1,2k+1}}-{Sm}_{2n+1,2k+1}
=\displaystyle= {i=kn1(1)ik+1(ik)(rni)+(1)nk(nk)k<n12n21rk=n1\displaystyle\begin{cases}\sum_{i=k}^{n-1}(-1)^{i-k+1}\binom{i}{k}\binom{r}{n-i}+(-1)^{n-k}\binom{n}{k}&k<n-1\\ 2n^{2}-1-r&k=n-1\end{cases}

The last step follows from Proposition 6.3 we have the formula in the statement. When m=2nm=2n is even, Theorem 2.5 shows that

(Euτ2n+1,2k+1)=(1)Euτ2n+1,2n2k1+e𝟙(2k2n+1)(Eu_{\tau^{\wedge}_{2n+1,2k+1}})^{\vee}=(-1)\cdot Eu_{\tau^{\wedge}_{2n+1,2n-2k-1}}+e\mathds{1}_{\mathbb{P}(\wedge^{2}k^{2n+1})}

The same argument shows that

χ2k,2r2n=\displaystyle\chi^{2n}_{2k,2r}= {i=kn1(1)ik+1(ik)(rni)+(1)nk(nk)k<n12n21rk=n1\displaystyle\begin{cases}\sum_{i=k}^{n-1}(-1)^{i-k+1}\binom{i}{k}\binom{r}{n-i}+(-1)^{n-k}\binom{n}{k}&k<n-1\\ 2n^{2}-1-r&k=n-1\end{cases}

7.3. Symmetric Matrix

Now we compute the sectional Euler characteristic of symmetric rank loci. We have the following theorem.

Theorem 7.3.

Let τm,AS\tau_{m,A}^{S\circ} be the orbits in (Sym2km)\mathbb{P}(Sym^{2}k^{m}). For any lB(τm,BS)(Sym2km)l_{B}\in(\tau^{S\circ}_{m,B})\subset\mathbb{P}(Sym^{2}k^{m\vee}), let LBL_{B} be the corresponding hyperplane. Then for A,BA,B from 11 to m1m-1 we have

χ(τm,ASLB)={l=1n(1)lk(lk)(rn+1l)(1)nk(n+1k)+χ2n+2,2kSA=2k,B=2r,m=2n+2χ2n+2,2k+1SA=2k+1,B=2r,m=2n+2l=1n(1)lk(lk)(r+1n+1l)(1)nk(n+1k)+χ2n+2,2kSA=2k,B=2r+1,m=2n+2l=1n(1)lk+1(lk)(rnl)+χ2n+2,2kSA=2k+1,B=2r+1,m=2n+2l=1n(1)lk(lk)(rn+1l)(1)nk(nk)+χ2n+1,2kSA=2kB=2r,m=2n+1(1)nk+1(nk)+χ2n+1,2k+1SA=2k+1,B=2r,m=2n+1l=0n1(1)lk+1(lk1)(rnl)(1)nk(nk)+χ2n+1,2kSA=2k,B=2r+1,m=2n+1l=1n1(1)lk(lk)(rnl)(1)nk+1(nk)+χ2n+1,2k+1SA=2k+1,B=2r+1,m=2n+1\displaystyle\chi(\tau_{m,A}^{S\circ}\cap L_{B})=\begin{cases}\sum_{l=1}^{n}(-1)^{l-k}\binom{l}{k}\binom{r}{n+1-l}-(-1)^{n-k}\binom{n+1}{k}+\chi^{S}_{2n+2,2k}&A=2k,B=2r,m=2n+2\\ \chi^{S}_{2n+2,2k+1}&A=2k+1,B=2r,m=2n+2\\ \sum_{l=1}^{n}(-1)^{l-k}\binom{l}{k}\binom{r+1}{n+1-l}-(-1)^{n-k}\binom{n+1}{k}+\chi^{S}_{2n+2,2k}&A=2k,B=2r+1,m=2n+2\\ \sum_{l=1}^{n}(-1)^{l-k+1}\binom{l}{k}\binom{r}{n-l}+\chi^{S}_{2n+2,2k}&A=2k+1,B=2r+1,m=2n+2\\ \sum_{l=1}^{n}(-1)^{l-k}\binom{l}{k}\binom{r}{n+1-l}-(-1)^{n-k}\binom{n}{k}+\chi^{S}_{2n+1,2k}&A=2kB=2r,m=2n+1\\ (-1)^{n-k+1}\binom{n}{k}+\chi^{S}_{2n+1,2k+1}&A=2k+1,B=2r,m=2n+1\\ \sum_{l=0}^{n-1}(-1)^{l-k+1}\binom{l}{k-1}\binom{r}{n-l}-(-1)^{n-k}\binom{n}{k}+\chi^{S}_{2n+1,2k}&A=2k,B=2r+1,m=2n+1\\ \sum_{l=1}^{n-1}(-1)^{l-k}\binom{l}{k}\binom{r}{n-l}-(-1)^{n-k+1}\binom{n}{k}+\chi^{S}_{2n+1,2k+1}&A=2k+1,B=2r+1,m=2n+1\end{cases}

Here χm,AS\chi^{S}_{m,A} are defined in Proposition 5.1-viivii. We make the convention that (ab)=0\binom{a}{b}=0 whenever b<0b<0.

Proof.

We follow the proof as in ordinary matrix case. Recall that for k=1,,m1k=1,\cdots,m-1 the variety τm,kS\tau^{S}_{m,k} is dual to τm,mkS\tau^{S}_{m,m-k}, with dimensions (mk)(m+k+1)21\frac{(m-k)(m+k+1)}{2}-1 and k(2mk+1)21\frac{k(2m-k+1)}{2}-1 respectively. Here (Sym2km)=N\mathbb{P}(Sym^{2}k^{m})=\mathbb{P}^{N}, whereN=(m+12)1N=\binom{m+1}{2}-1. Thus from Theorem 2.5 we have

(Euτm,kS)=(1)(mk)(m+1)Euτm,mkS+e𝟙N(Eu_{\tau^{S}_{m,k}})^{\vee}=(-1)^{(m-k)(m+1)}Eu_{\tau^{S}_{m,m-k}}+e\mathds{1}_{\mathbb{P}^{N}}

We denote βi,jm\beta^{m}_{i,j} to be the difference χ(τm,jSLj)χ(τm,jSH)\chi({\tau_{m,j}^{S\circ}\cap L_{j}})-\chi({\tau_{m,j}^{S\circ}\cap H}), where Ljτm,jSL_{j}\in\tau_{m,j}^{S\circ} in the dual space and HH is a generic hyperplane. First we consider the case that m=2n+2m=2n+2 is even. Apply Theorem 6.6 we have

((00)0(10)0(n0)0(11)(11)(n1)(n1)0000(nn))(β1,12n+2β1,2n+12n+2β2,12n+2β2,2n+12n+2β2n+1,12n+2β2n+1,2n+12n+2)=(0000(nn)0(11)(11)(n1)(n1)(00)0(10)0(n0)).\left(\begin{array}[]{cccccc}\binom{0}{0}&0&\binom{1}{0}&0&\cdots&\binom{n}{0}\\ 0&\binom{1}{1}&\binom{1}{1}&\cdots&\binom{n}{1}&\binom{n}{1}\\ \cdots&\cdots&\cdots&\cdots&\cdots&\cdots\\ 0&0&0&0&\cdots&\binom{n}{n}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{2n+2}_{1,1}&\cdots&\beta^{2n+2}_{1,2n+1}\\ \beta^{2n+2}_{2,1}&\cdots&\beta^{2n+2}_{2,2n+1}\\ \cdots&\cdots&\cdots\\ \beta^{2n+2}_{2n+1,1}&\cdots&\beta^{2n+2}_{2n+1,2n+1}\end{array}\right)=\left(\begin{array}[]{cccccc}0&0&\cdots&0&0&-\binom{n}{n}\\ \cdots&\cdots&\cdots&\cdots&\cdots&\cdots\\ 0&\binom{1}{1}&\binom{1}{1}&\cdots&\binom{n}{1}&\binom{n}{1}\\ -\binom{0}{0}&0&-\binom{1}{0}&0&\cdots&-\binom{n}{0}\end{array}\right).
Lemma 4.

The inverse matrix of

((00)0(10)0(n0)0(11)(11)(n1)(n1)0000(nn))\left(\begin{array}[]{cccccc}\binom{0}{0}&0&\binom{1}{0}&0&\cdots&\binom{n}{0}\\ 0&\binom{1}{1}&\binom{1}{1}&\cdots&\binom{n}{1}&\binom{n}{1}\\ \cdots&\cdots&\cdots&\cdots&\cdots&\cdots\\ 0&0&0&0&\cdots&\binom{n}{n}\end{array}\right)

is

((00)0(10)0(20)0(n0)0(11)(11)(21)(21)(n1)(n1)00(11)0(21)0(n1)000000(nn))\left(\begin{array}[]{cccccccc}\binom{0}{0}&0&-\binom{1}{0}&0&\binom{2}{0}&0&\cdots&\binom{n}{0}\\ 0&\binom{1}{1}&-\binom{1}{1}&-\binom{2}{1}&\binom{2}{1}&\cdots&-\binom{n}{1}&\binom{n}{1}\\ 0&0&\binom{1}{1}&0&-\binom{2}{1}&0&\cdots&-\binom{n}{1}\\ \cdots&\cdots&\cdots&\cdots&\cdots&\cdots&\cdots&\cdots\\ 0&0&0&0&0&0&\cdots&\binom{n}{n}\end{array}\right)

The proof of the Lemma is straight binomial computation, and we omit it here. For any A,BA,B from 11 to 2n+12n+1 we then have:

βA,B2n+2={l=1n(1)lk(lk)(rn+1l)A=2k,B=2r0A=2k+1,B=2rl=1n(1)lk(lk)(r+1n+1l)A=2k,B=2r+1l=1n(1)lk+1(lk)(rnl)A=2k+1,B=2r+1\displaystyle\beta^{2n+2}_{A,B}=\begin{cases}\sum_{l=1}^{n}(-1)^{l-k}\binom{l}{k}\binom{r}{n+1-l}&A=2k,B=2r\\ 0&A=2k+1,B=2r\\ \sum_{l=1}^{n}(-1)^{l-k}\binom{l}{k}\binom{r+1}{n+1-l}&A=2k,B=2r+1\\ \sum_{l=1}^{n}(-1)^{l-k+1}\binom{l}{k}\binom{r}{n-l}&A=2k+1,B=2r+1\end{cases}

For the case that m=2n+1m=2n+1, notice that

(Euτ2n+1,kS)=Euτ2n+1,2n+1kS+e𝟙N(Eu_{\tau^{S}_{2n+1,k}})^{\vee}=Eu_{\tau^{S}_{2n+1,2n+1-k}}+e\mathds{1}_{\mathbb{P}^{N}}

Thus we have

((00)0(10)0(n0)0(11)(11)(n1)(n1)0000(nn))(β1,12n+1β1,2n+12n+1β2,12n+1β2,2n+12n+1β2n+1,12n+1β2n+1,2n+12n+1)=(0000(nn)0(11)(11)(n1)(n1)(00)0(10)0(n0)).\left(\begin{array}[]{cccccc}\binom{0}{0}&0&\binom{1}{0}&0&\cdots&\binom{n}{0}\\ 0&\binom{1}{1}&\binom{1}{1}&\cdots&\binom{n}{1}&\binom{n}{1}\\ \cdots&\cdots&\cdots&\cdots&\cdots&\cdots\\ 0&0&0&0&\cdots&\binom{n}{n}\end{array}\right)\cdot\left(\begin{array}[]{ccc}\beta^{2n+1}_{1,1}&\cdots&\beta^{2n+1}_{1,2n+1}\\ \beta^{2n+1}_{2,1}&\cdots&\beta^{2n+1}_{2,2n+1}\\ \cdots&\cdots&\cdots\\ \beta^{2n+1}_{2n+1,1}&\cdots&\beta^{2n+1}_{2n+1,2n+1}\end{array}\right)=\left(\begin{array}[]{cccccc}0&0&\cdots&0&0&\binom{n}{n}\\ \cdots&\cdots&\cdots&\cdots&\cdots&\cdots\\ 0&\binom{1}{1}&\binom{1}{1}&\cdots&\binom{n}{1}&\binom{n}{1}\\ \binom{0}{0}&0&\binom{1}{0}&0&\cdots&\binom{n}{0}\end{array}\right).

Mimicking the computation for m=2n+2m=2n+2 we get

βA,B2n+1={l=1n(1)lk(lk)(rn+1l)A=2k,B=2r0A=2k+1,B=2rl=0n1(1)lk+1(lk1)(rnl)A=2k,B=2r+1l=1n1(1)lk(lk)(rnl)A=2k+1,B=2r+1\displaystyle\beta^{2n+1}_{A,B}=\begin{cases}\sum_{l=1}^{n}(-1)^{l-k}\binom{l}{k}\binom{r}{n+1-l}&A=2k,B=2r\\ 0&A=2k+1,B=2r\\ \sum_{l=0}^{n-1}(-1)^{l-k+1}\binom{l}{k-1}\binom{r}{n-l}&A=2k,B=2r+1\\ \sum_{l=1}^{n-1}(-1)^{l-k}\binom{l}{k}\binom{r}{n-l}&A=2k+1,B=2r+1\end{cases}

Recall that Smm,A=χ(τm,AS)χ(τm,ASH){Sm}_{m,A}=\chi(\tau_{m,A}^{S\circ})-\chi(\tau_{m,A}^{S\circ}\cap H). We then have

χ(τm,ASLB)=\displaystyle\chi(\tau_{m,A}^{S\circ}\cap L_{B})= βA,Bm+χ(τm,ASH)\displaystyle\beta^{m}_{A,B}+\chi(\tau_{m,A}^{S\circ}\cap H)
=\displaystyle= βA,Bm+χ(τm,AS)Smm,A\displaystyle\beta^{m}_{A,B}+\chi(\tau_{m,A}^{S\circ})-{Sm}_{m,A}

Proposition 5.1-77 shows that for symmetric rank loci we have

χ(τm,AS)=χm,AS\displaystyle\chi(\tau_{m,A}^{S\circ})=\chi^{S}_{m,A}

Thus combine with Proposition 6.5 we have the formula in the theorem. ∎

8. Stalk Euler Characteristic of the Intersection Cohomology Complex of Rank loci

In this section we compute the stalk Euler characteristic of the Intersection Cohomology complex of rank loci. In this section we assume k=k=\mathbb{C}.

8.1. Index Theorem

Let iISi\sqcup_{i\in I}S_{i} be a (finite) Whitney stratification of a complex projective variety XX in N\mathbb{P}^{N}. The theory of Chern-Schwartz-MacPherson classes of constructible functions can be viewed as the pushdown of the theory of characteristic cycles of constructible sheaves. In short, there is a map CCCC from the Grothendieck group of derived category of constructible sheaves to the group of (conical) Lagrangian cycles of TNT^{*}\mathbb{P}^{N} with support in XX, together with the following diagram

K0(Dcb(X)){K_{0}(D^{b}_{c}(X))}L(X){L(X)}Am1((TN)){A_{m-1}(\mathbb{P}(T^{*}\mathbb{P}^{N}))}F(X){F(X)}A(X){A_{*}(X)}A(N){A_{*}(\mathbb{P}^{N})}CC\scriptstyle{CC}χx\scriptstyle{\chi_{x}}\scriptstyle{\sim}[]\scriptstyle{[*]}sh\scriptstyle{sh}c\scriptstyle{c_{*}}i\scriptstyle{i_{*}}

Here [][*] denotes the fundamental classes. The first vertical map χx\chi_{x} takes the stalk Euler characteristic of a constructible sheaf complex, and the last vertical map sh:Am1((TM))A(M)sh\colon A_{m-1}(\mathbb{P}(T^{*}M))\to A_{*}(M) is given by ‘casting the shadow ’ process discussed in [3].

In fact the theory of characteristic cycles doesn’t require projectivization or compactness: the above projective assumption on XX is totally due to the Chern-Schwartz-MacPherson theory. The characteristic cycles on a projective stratification behave the same with the induced conical stratification on its affine cone. For the rest of this section we only assume X=iISiMX=\sqcup_{i\in I}S_{i}\subset M is a Whitney stratified space. For any stratum SiS_{i}, there is a naturally assigned constructible sheaf called the Intersection Cohomology Sheaf complex of SiS_{i}. We denote it by 𝒞S¯i\mathcal{IC}_{\overline{S}_{i}}, and call it the 𝒞\mathcal{IC} sheaf of SiS_{i} for short. This complex can be obtained by a sequence of (derived truncated) pushforward of the local system on SiS_{i} to its closure along all the smaller strata. For details about the intersection cohomology sheaf and intersection homology we refer to [19] and [20]. For any constructible sheaf \mathcal{F}^{\bullet} on XX, its characteristic cycle has the form

CC()=iIri()[TS¯iM].CC(\mathcal{F}^{\bullet})=\sum_{i\in I}r_{i}(\mathcal{F}^{\bullet})[T^{*}_{\overline{S}_{i}}M]\/.

Here for any ii, TS¯iMT^{*}_{\overline{S}_{i}}M is the conormal space of S¯i\overline{S}_{i} defied as the closure of

{(x,λ)|sSi,λ(TxSi)=0}TM.\{(x,\lambda)|s\in S_{i},\lambda(T_{x}S_{i})=0\}\subset T^{*}M\/.

The integer coefficients ri()r_{i}(\mathcal{F}^{\bullet}) are called the Microlocal Multiplicities: these are very important invariants in microlocal geometry, for details we refer to [28, Chapter 9]. The following deep theorem from [13, Theorem 3], [27, Theorem 6.3.1] reveals the relation among the microlocal multiplicities , the stalk Euler characterictic, and the local Euler obstructions :

Theorem 8.1 (Microlocal Index Formula).

For any i,jIi,j\in I, and any xSix\in S_{i} we define

χi()=p(1)pdimHp((x))\chi_{i}(\mathcal{F}^{\bullet})=\sum_{p}(-1)^{p}\dim H^{p}(\mathcal{F}^{\bullet}(x))

to be the stalk Euler characteristic of \mathcal{F}^{\bullet}. Let e(j,i)=EuSi¯(Sj)e(j,i)=Eu_{\overline{S_{i}}}(S_{j}) be the local Euler obstructions. We have the following formula:

χj()=iI(1)dimSie(j,i)ri().\chi_{j}(\mathcal{F}^{\bullet})=\sum_{i\in I}(-1)^{\dim S_{i}}e(j,i)r_{i}(\mathcal{F}^{\bullet})\/.

This theorem suggests that if one knows about any two sets of the indexes, then one can compute the third one.

8.2. Main Result

Now we consider rank loci. In [4] the authors studied concretely the category of perverse sheaves on rank loci, and showed the following properties in [4, Corollary 4.10, 4.11]:

Proposition 8.2.

For ordinary matrix rank loci with i1i\geq 1, and for skew-symmetric matrix rank loci with ni<2n2n-i<2\lfloor\frac{n}{2}\rfloor , we have

π1(Σi)=π1(Σi)=1\pi_{1}(\Sigma^{\circ}_{i})=\pi_{1}(\Sigma^{\wedge\circ}_{i})=1

Thus there are only trivial local systems. The characteristic cycles of the 𝒞\mathcal{IC} complexes are then irreducible, i.e., for j1j\geq 1 we have

CC(𝒞Σn,j())=[TΣn,jn2];CC(𝒞Σn,j())=[TΣn,j(n2)].CC(\mathcal{IC}_{\Sigma_{n,j}}(\mathbb{C}))=[T_{\Sigma_{n,j}}^{*}\mathbb{C}^{n^{2}}];\quad CC(\mathcal{IC}_{\Sigma^{\wedge}_{n,j}}(\mathbb{C}))=[T_{\Sigma^{\wedge}_{n,j}}^{*}\mathbb{C}^{\binom{n}{2}}]\/.

For symmetric rank loci, when n1i1n-1\geq i\geq 1 we have

π1(ΣiS)=/2;\pi_{1}(\Sigma^{S\circ}_{i})=\mathbb{Z}/2\mathbb{Z};

Thus for n1i1n-1\geq i\geq 1 there is a unique non-trivial rank 11 local system i\mathcal{L}_{i} on each ΣiS\Sigma^{S\circ}_{i}, such that i2=\mathcal{L}_{i}^{\otimes 2}=\mathbb{C}. The characteristic cycles of the 𝒞\mathcal{IC} complexes are given by

CC(𝒞Σn,jS())=\displaystyle CC(\mathcal{IC}_{\Sigma^{S}_{n,j}}(\mathbb{C}))= {[TΣn,jS(n+12)]j=2k[TΣn,jS(n+12)]+[TΣn,j+1N]j=2k+1\displaystyle\begin{cases}[T_{\Sigma^{S}_{n,j}}^{*}\mathbb{C}^{\binom{n+1}{2}}]&j=2k\\ [T_{\Sigma^{S}_{n,j}}^{*}\mathbb{C}^{\binom{n+1}{2}}]+[T_{\Sigma_{n,j+1}}^{*}\mathbb{P}^{N}]&j=2k+1\\ \end{cases}
CC(𝒞Σn,jS(j))=\displaystyle CC(\mathcal{IC}_{\Sigma^{S}_{n,j}}(\mathcal{L}_{j}))= {[TΣn,jS(n+12)]j=2k+1[TΣn,jS(n+12)]+[TΣn,j+1N]j=2k;\displaystyle\begin{cases}[T_{\Sigma^{S}_{n,j}}^{*}\mathbb{C}^{\binom{n+1}{2}}]&j=2k+1\\ [T_{\Sigma^{S}_{n,j}}^{*}\mathbb{C}^{\binom{n+1}{2}}]+[T_{\Sigma_{n,j+1}}^{*}\mathbb{P}^{N}]&j=2k\end{cases};

For j=nj=n we have Σn,nS={0}\Sigma^{S}_{n,n}=\{0\} and CC(𝒞Σn,n)=[T0(n+12)]CC(\mathcal{IC}_{\Sigma_{n,n}})=[T_{0}^{*}\mathbb{C}^{\binom{n+1}{2}}].

Thus combine the index theorem with above proposition, with the knowledge in §6 we obtain the following:

Theorem 8.3 (Stalk Euler Characteristic of 𝒞\mathcal{IC} complex).

Let 𝒞Σn,k()\mathcal{IC}_{\Sigma^{*}_{n,k}(\mathbb{C})} ((respectively, 𝒞Σn,kS(k))\mathcal{IC}_{\Sigma^{S}_{n,k}}(\mathcal{L}_{k})) be the intersection cohomology sheaf complex of Σn,k\Sigma^{*}_{n,k} ((respectively, Σn,kS)\Sigma^{S}_{n,k}). We use χΣn,r\chi_{\Sigma^{*\circ}_{n,r}} to denote χp\chi_{p} for any pΣn,rp\in\Sigma^{*\circ}_{n,r}. For ordinary matrix rank loci (=)(*=\emptyset), we have

(1)dimΣn,iχΣn,r(𝒞Σn,k())=(rk).(-1)^{\dim\Sigma_{n,i}}\chi_{\Sigma^{\circ}_{n,r}}(\mathcal{IC}_{\Sigma_{n,k}}(\mathbb{C}))=\binom{r}{k}\/.

For skew-symmetric matrix rank loci (=)(*=\wedge), we have:

(1)dimΣ2n,2kχΣ2n,2r(𝒞Σ2n,2k)=(1)dimΣ2n+1,2k+1χΣ2n+1,2r+1(𝒞Σ2n+1,2k+1)=(rk)\displaystyle(-1)^{\dim\Sigma^{\wedge}_{2n,2k}}\chi_{\Sigma^{\wedge\circ}_{2n,2r}}(\mathcal{IC}_{\Sigma^{\wedge}_{2n,2k}})=(-1)^{\dim\Sigma^{\wedge}_{2n+1,2k+1}}\chi_{\Sigma^{\wedge\circ}_{2n+1,2r+1}}(\mathcal{IC}_{\Sigma^{\wedge}_{2n+1,2k+1}})=\binom{r}{k}

For symmetric matrix rank loci (=S)(*=S), we have:

(1)dimΣn,iSχΣn,jS(𝒞Σn,iS(i))=\displaystyle(-1)^{\dim\Sigma^{S}_{n,i}}\chi_{\Sigma^{S\circ}_{n,j}}(\mathcal{IC}_{\Sigma^{S}_{n,i}}(\mathcal{L}_{i}))= {(ks)(i,j)=(2s,2k)0(i,j)=(2s+1,2k)0(i,j)=(2s,2k+1)(ks)(i,j)=(2s+1,2k+1)\displaystyle\begin{cases}\binom{k}{s}&(i,j)=(2s,2k)\\ 0&(i,j)=(2s+1,2k)\\ 0&(i,j)=(2s,2k+1)\\ \binom{k}{s}&(i,j)=(2s+1,2k+1)\end{cases}
(1)dimΣn,iSχΣn,jS(𝒞Σn,iS())=\displaystyle(-1)^{\dim\Sigma^{S}_{n,i}}\chi_{\Sigma^{S\circ}_{n,j}}(\mathcal{IC}_{\Sigma^{S}_{n,i}}(\mathbb{C}))= {(ks)(i,j)=(2s,2k)(ks+1)(i,j)=(2s+1,2k)(k+1s+1)(i,j)=(2s+1,2k)(ks)(i,j)=(2s+1,2k+1)\displaystyle\begin{cases}\binom{k}{s}&(i,j)=(2s,2k)\\ \binom{k}{s+1}&(i,j)=(2s+1,2k)\\ \binom{k+1}{s+1}&(i,j)=(2s+1,2k)\\ \binom{k}{s}&(i,j)=(2s+1,2k+1)\end{cases}
Proof.

Since the proof for two formulas are almost identical, we only prove the one for χΣn,jS(𝒞Σn,iS(i))\chi_{\Sigma^{S\circ}_{n,j}}(\mathcal{IC}_{\Sigma^{S}_{n,i}}(\mathcal{L}_{i})). Let dn,j=dimΣn,jSd_{n,j}=\dim\Sigma^{S\circ}_{n,j}, we have

χΣn,rS(𝒞Σn,kS(i))=\displaystyle\chi_{\Sigma^{S\circ}_{n,r}}(\mathcal{IC}_{\Sigma^{S}_{n,k}}(\mathcal{L}_{i}))= j(1)dn,jEuΣn,jS(Σn,rS)rj(𝒞Σn,kS(i))\displaystyle\sum_{j}(-1)^{d_{n,j}}Eu_{\Sigma^{S}_{n,j}}(\Sigma^{S\circ}_{n,r})\cdot r_{j}(\mathcal{IC}_{\Sigma^{S}_{n,k}}(\mathcal{L}_{i}))
=\displaystyle= {(1)dn,2AEuΣn,2AS(Σn,rS)+(1)dn,2A+1EuΣn,2A+1S(Σn,rS)k=2A;(1)dn,2A+1EuΣn,2A+1S(Σn,rS)k=2A+1\displaystyle\begin{cases}(-1)^{d_{n,2A}}Eu_{\Sigma^{S}_{n,2A}}(\Sigma^{S\circ}_{n,r})+(-1)^{d_{n,2A+1}}Eu_{\Sigma^{S}_{n,2A+1}}(\Sigma^{S\circ}_{n,r})&k=2A;\\ (-1)^{d_{n,2A+1}}Eu_{\Sigma^{S}_{n,2A+1}}(\Sigma^{S\circ}_{n,r})&k=2A+1\\ \end{cases}

Recall that dn,k=k(nk)+(nk+12)d_{n,k}=k(n-k)+\binom{n-k+1}{2}. Thus we have

(1)dn,2A+1dn,2A=(1)k1.(-1)^{d_{n,2A+1}-d_{n,2A}}=(-1)^{k-1}\/.

This proves that

(1)dn,kχΣn,rS(𝒞Σn,kS(i))=\displaystyle(-1)^{d_{n,k}}\chi_{\Sigma^{S\circ}_{n,r}}(\mathcal{IC}_{\Sigma^{S}_{n,k}}(\mathcal{L}_{i}))= {EuΣn,2AS(Σn,rS)EuΣn,2A+1S(Σn,rS)k=2A;EuΣn,2A+1S(Σn,rS)k=2A+1\displaystyle\begin{cases}Eu_{\Sigma^{S}_{n,2A}}(\Sigma^{S\circ}_{n,r})-Eu_{\Sigma^{S}_{n,2A+1}}(\Sigma^{S\circ}_{n,r})&k=2A;\\ Eu_{\Sigma^{S}_{n,2A+1}}(\Sigma^{S\circ}_{n,r})&k=2A+1\\ \end{cases}

The formula then follows directly from Theorem 6.6. ∎

Remark 5.

As shown in [20, §6.2], when a complex variety XX admits a small resolution, the 𝒞\mathcal{IC} complex is the ((shifted)) derived pushforward of the constant sheaf from the small resolution. Thus for any point xXx\in X, the signed stalk Euler characteristic of 𝒞X\mathcal{IC}_{X} equals the Euler characteristic of the fiber:

(1)dimXiχx(𝒞X)=χ(p1(x)).(-1)^{\dim X_{i}}\chi_{x}(\mathcal{IC}_{X})=\chi(p^{-1}(x)).

For ordinary matrix rank loci τn,k\tau_{n,k}, the Tjurina transform τ^n,k\hat{\tau}_{n,k} is exactly a small resolution. The fiber at τn,r\tau_{n,r} is isomorphic to G(k,r)G(k,r), which has Euler characteristic (rk)\binom{r}{k}. For symmetric and skew-symmetric rank loci, the Tjurina transforms fail to be small: the fiber Euler characteristics are too large.

In fact, assuming that there are canonically define small resolutions for symmetric rank loci, then the value χΣn,2k+1S(𝒞Σn,2s+1S())\chi_{\Sigma^{S\circ}_{n,2k+1}}(\mathcal{IC}_{\Sigma^{S}_{n,2s+1}}(\mathbb{C})) forces the fiber Euler characteristics to jump. This supports the following conjecture.

Conjecture 1.

There do not exist canonical small resolutions for symmetric rank loci.

9. An Enumerative Problem

In this section we finish the proof of Proposition 6.5. First we consider the following situation. Let VV be a complex vector space of dimension n+1n+1, and let X(V)X\subset\mathbb{P}(V) be an irreducible quadratic hypersurface defined by the vanishing of fXf_{X}. Let 𝔾(d,n)=G(d+1,n+1)\mathbb{G}(d,n)=G(d+1,n+1) be the Grassmannian of projective dd planes, with SS and QQ being universal sub and quotient bundles. We consider the space of projective dd planes contained in XX:

Zd,nX:={Λ|ΛX}𝔾(d,n)Z^{X}_{d,n}:=\{\Lambda|\Lambda\subset X\}\subset\mathbb{G}(d,n)

This space is called the Fano scheme of XX. Notice that quadratic hypersurfaces in (V)\mathbb{P}(V) correspond to sections of the sheaf Sym2(V)Sym^{2}(V^{\vee}), thus the function defining XX induces a section sXs_{X} from 𝔾(d,n)\mathbb{G}(d,n) to Hom(Sym2(S),)Hom(Sym^{2}(S),\mathbb{C}), sending Λ\Lambda to

sX|Λ:Sym2ΛSym2V.{s_{X}}|_{\Lambda}\colon Sym^{2}\Lambda\subset Sym^{2}V\to\mathbb{C}\/.

The last map is defined by fXf_{X}. One can check that the zero locus of sXs_{X} is exactly supported on Zd,nXZ^{X}_{d,n}. When 2dn12d\leq n-1, if we choose a generic smooth section fXf_{X} in Sym2(V)Sym^{2}(V^{\vee}), the space Zd,nXZ^{X}_{d,n} is smooth of codimension (d+22)\binom{d+2}{2}. We denote this space by Zd,nZ_{d,n}, since for generic sections fXf_{X} its topological structure is preserved. For details we refer to [1].

In classical enumerative geometry we concern the degree of this variety, but here we consider its (topological) Euler characteristic instead. When Zd,nZ_{d,n} is of dimension 0, they coincide. We denote F(d,n)=χ(Zd,n)F(d,n)=\chi(Z_{d,n}) to be the Euler characteristic.

Theorem 9.1.

When 2dn12d\leq n-1, the functions F(d,n)F(d,n) satisfy the following recursive formula.

F(d,2m)=\displaystyle F(d,2m)= F(d,2m1),\displaystyle F(d,2m-1)\/,
F(d,2m+1)=\displaystyle F(d,2m+1)= F(d,2m)+2F(d1,2m).\displaystyle F(d,2m)+2\cdot F(d-1,2m)\/.

Thus computing the initial value we have

F(d,2m1)=F(d,2m)=2d+1(mmd1).F(d,2m-1)=F(d,2m)=2^{d+1}\cdot\binom{m}{m-d-1}\/.
Proof.

First we consider the initial value d=0d=0. When d=0d=0 we can see that Z0,n=XZ_{0,n}=X and F(0,n)=χ(X)F(0,n)=\chi(X). Notice that XX is a smooth quadratic hypersurface in n\mathbb{P}^{n}, thus we have

csmX(H)=2H(1+H)n+1(1+2H)=i=0n1(1)(n+1i)(2)niHn+ lower degree terms.c_{sm}^{X}(H)=\frac{2H\cdot(1+H)^{n+1}}{(1+2H)}=\sum_{i=0}^{n-1}(-1)\binom{n+1}{i}(-2)^{n-i}H^{n}+\text{ lower degree terms}\/.

In particular we have

χ(X)=i=0n1(1)(n+1i)(2)ni=(1)n+112(n+1n)2=n+1+(1)n+112.\chi(X)=\sum_{i=0}^{n-1}(-1)\binom{n+1}{i}(-2)^{n-i}=\frac{(-1)^{n+1}-1-2\binom{n+1}{n}}{2}=n+1+\frac{(-1)^{n+1}-1}{2}\/.

This shows that F(0,2m)=F(0,2m1)=2mF(0,2m)=F(0,2m-1)=2m.

Now we prove the induction process. Choose a generic hypersurface XX such that Zd,n=Zd,nXZ_{d,n}=Z^{X}_{d,n}. Notice that the intersection of XX with a generic hyperplane HH is again a generic quadratic hypersurface Y=XHY=X\cap H in Hn1H\cong\mathbb{P}^{n-1}. We define the following two strata in Zd,nXZ^{X}_{d,n}:

AH:={Λ|ΛH}Zd,nX;BH:=Zd,nXAH.A_{H}:=\{\Lambda|\Lambda\subset H\}\subset Z^{X}_{d,n};\quad B_{H}:=Z^{X}_{d,n}\setminus A_{H}\/.

There are two canonical restriction maps

p:AHZd,n1Y;q:BHZd1,n1Yp\colon A_{H}\to Z^{Y}_{d,n-1};\quad q\colon B_{H}\to Z^{Y}_{d-1,n-1}

defined as follows. For pp, we simply send ΛX\Lambda\subset X to ΛY=XH\Lambda\subset Y=X\cap H, since ΛH\Lambda\subset H. By the definition of AHA_{H} one can observe that pp is an isomorphism, thus we have χ(AH)=χ(Zd,n1Y)=F(d,n1)\chi(A_{H})=\chi(Z^{Y}_{d,n-1})=F(d,n-1).

We define qq by sending a dd-plane Λ\Lambda to the d1d-1-plane ΛH\Lambda\cap H. This is well-defined since ΛBH\Lambda\in B_{H} iff ΛH\Lambda\not\subset H. The map qq is dominant for generic XX and HH, and for any ΓZd1,n1Y\Gamma\in Z^{Y}_{d-1,n-1} the fiber q1(Γ)q^{-1}(\Gamma) is

{ΛG(d+1,n+1)|Γ(Λ)X}{ΛG(d+1,n)|Γ(Λ)XH=Y}.\{\Lambda\in G(d+1,n+1)|\Gamma\subset\mathbb{P}(\Lambda)\subset X\}\setminus\{\Lambda\in G(d+1,n)|\Gamma\subset\mathbb{P}(\Lambda)\subset X\cap H=Y\}\/.

We denote the two spaces by CXC_{X} and CYC_{Y} respectively. Notice that {ΛG(d+1,n+1)|Γ(Λ)}\{\Lambda\in G(d+1,n+1)|\Gamma\subset\mathbb{P}(\Lambda)\} can be identified with G(1,nd)G(1,n-d). For any LG(1,nd)L\in G(1,n-d), to force a the (projectivization of) the span of LL with Γ\Gamma to live in XX, we need d1d-1 linear equations and a degree 22 equation. Thus we can identify CXC_{X} with a generic quadratic hypersurface in n2d\mathbb{P}^{n-2d}, and similarly CYC_{Y} is identified with a generic quadratic hypersurface in n2d1\mathbb{P}^{n-2d-1}. This shows that the fibers q1(Γ)q^{-1}(\Gamma) are all isomorphic, and we then have

χ(BH)=χ(q1(Γ))F(d1,n1).\chi(B_{H})=\chi(q^{-1}(\Gamma))\cdot F(d-1,n-1)\/.

Moreover, the generic assumption shows that CXC_{X} and CYC_{Y} are both smooth. Thus similar argument of the initial case shows

χ(q1(Γ))=\displaystyle\chi(q^{-1}(\Gamma))= χ(CX)χ(CY)\displaystyle\chi(C_{X})-\chi(C_{Y})
=\displaystyle= (n2d+1+(1)n2d+112)(n2d+(1)n2d12)\displaystyle\left(n-2d+1+\frac{(-1)^{n-2d+1}-1}{2}\right)-\left(n-2d+\frac{(-1)^{n-2d}-1}{2}\right)
=\displaystyle= 1+(1)n1\displaystyle 1+(-1)^{n-1}

The proof is then closed by

F(d,n)=χ(Zd,nX)=χ(AH)+χ(BH)=F(d,n1)+(1+(1)nd)F(d1,n1).F(d,n)=\chi(Z^{X}_{d,n})=\chi(A_{H})+\chi(B_{H})=F(d,n-1)+(1+(-1)^{n-d})\cdot F(d-1,n-1)\/.

When d=2rd=2r and n=3r+1n=3r+1, the space Zk,mZ_{k,m} is of dimension 0. Thus we have a direct Corollary

Corollary 3.

For a generic quadratic hypersurface XX in 3r+1\mathbb{P}^{3r+1}, there are exactly 2d+1(AAd1)2^{d+1}\binom{A}{A-d-1} projective 2r2r-planes living in XX. Here A:=2n+12A:=2\cdot\lfloor\frac{n+1}{2}\rfloor.

Now we can finish the proof of Proposition 6.5.

Corollary 4.
G(d+1,2m+1)c(SQ)c(d+22)(Sym2(S))c(Sym2(S))=\displaystyle\int_{G(d+1,2m+1)}\frac{c(S^{\vee}\otimes Q)c_{\binom{d+2}{2}}(Sym^{2}(S^{\vee}))}{c(Sym^{2}(S^{\vee}))}= 2d+1(mmd1)\displaystyle 2^{d+1}\cdot\binom{m}{m-d-1}
G(d+1,2m)c(SQ)c(d+22)(Sym2(S))c(Sym2(S))=\displaystyle\int_{G(d+1,2m)}\frac{c(S^{\vee}\otimes Q)c_{\binom{d+2}{2}}(Sym^{2}(S^{\vee}))}{c(Sym^{2}(S^{\vee}))}= 2d+1(mmd1).\displaystyle 2^{d+1}\cdot\binom{m}{m-d-1}\/.
Proof.

When 2dn12d\leq n-1, recall that the scheme Zd,nXZ^{X}_{d,n} is the zero locus of a section sX:G(d+1,n+1)Sym2(S)s_{X}\colon G(d+1,n+1)\to Sym^{2}(S^{\vee}). For generic XX it is smooth of codimension (k+22)=rkSym2(S)\binom{k+2}{2}=\operatorname{rk}Sym^{2}(S^{\vee}). Thus we have

csmZd,nX=c(SQ)c(k+22)(Sym2(S))c(Sym2(S))A(G(d+1,n+1)).c_{sm}^{Z^{X}_{d,n}}=\frac{c(S^{\vee}\otimes Q)c_{\binom{k+2}{2}}(Sym^{2}(S^{\vee}))}{c(Sym^{2}(S^{\vee}))}\in A_{*}(G(d+1,n+1))\/.

And the formula follows directly from the theorem.

It remains to prove the following: for 2dn2d\geq n we have

G(d+1,n+1)c(SQ)c(d+22)(Sym2(S))c(Sym2(S))=0.\int_{G(d+1,n+1)}\frac{c(S^{\vee}\otimes Q)c_{\binom{d+2}{2}}(Sym^{2}(S^{\vee}))}{c(Sym^{2}(S^{\vee}))}=0\/.

This follows from by standard Schubert calculus. First, from [30] we have

c(d+22)(Sym2(S))=2d+1Δ[d+1](c(S));c_{\binom{d+2}{2}}(Sym^{2}(S^{\vee}))=2^{d+1}\Delta_{[d+1]}(c(S^{\vee}))\/;

where [d+1][d+1] denotes the partition (d+1,d,d1,,1,0,,0)(d+1,d,d-1,\cdots,1,0,\cdots,0). Recall [16, Lemma 14.5.1], which says that if there exists some 0rk10\leq r\leq k-1 such that si(E)=0s_{i}(E)=0 for i>ri>r and λr+1>0\lambda_{r+1}>0, then

Δ(λ1,,λk)(c(E))=0\Delta_{(\lambda_{1},\cdots,\lambda_{k})}(c(E))=0

Notice that when 2dn2d\geq n, rkQ=ndd<d+1=rkS\operatorname{rk}Q=n-d\leq d<d+1=\operatorname{rk}S. Thus si(S)=ci(Q)=0s_{i}(S^{\vee})=c_{i}(Q^{\vee})=0 for i>ndi>n-d. Take λ=[d+1]\lambda=[d+1], one sees that λnd+1>0\lambda_{n-d+1}>0, and then

Δ[d+1](c(S))=0.\Delta_{[d+1]}(c(S^{\vee}))=0\/.

This completes the proof. ∎

Remark 6.

Although we work with \mathbb{C} in this section, notice that Corollary 4 is purely about binomial identities, thus it naturally holds for any characteristic 0 field.

Remark 7.

Despite the similarity in forms, the author doesn’t know a good direct proof to the Schubert identity appeared in the computation in the skew-symmetric case:

G(2k,2n)c(SQ)c(2n2k2)(2Q)c(2Q)[G(2k,2n)]=\displaystyle\int_{G(2k,2n)}\frac{c(S^{\vee}\otimes Q)c_{\binom{2n-2k}{2}}(\wedge^{2}Q)}{c(\wedge^{2}Q)}\cap[G(2k,2n)]= r=kn(1)nk(2r2k)(nr)\displaystyle\sum_{r=k}^{n}(-1)^{n-k}\binom{2r}{2k}\binom{n}{r}
G(2k+1,2n+1)c(SQ)c(2n2k2)(2Q)c(2Q)[G(2k+1,2n+1)]=\displaystyle\int_{G(2k+1,2n+1)}\frac{c(S^{\vee}\otimes Q)c_{\binom{2n-2k}{2}}(\wedge^{2}Q)}{c(\wedge^{2}Q)}\cap[G(2k+1,2n+1)]= r=kn(1)nk(2r+12k+1)(nr)\displaystyle\sum_{r=k}^{n}(-1)^{n-k}\binom{2r+1}{2k+1}\binom{n}{r}

It should be interesting to find out what enumerative problems correspond to them.

Remark 8.

In fact, the author does not know a direct Schubert calculus proof of the Schubert integration equalities for the skew-symmetric and the symmetric case. Based on the computation for small Grassmannian we believe that the clean forms come from the symmetry and certain vanishing property from Schubert calculus. We will discuss more about the formulas in another paper.

References

  • [1] Allen B. Altman and Steven L. Kleiman. Foundations of the theory of Fano schemes. Compositio Math., 34(1):3–47, 1977.
  • [2] P. Aluffi. Projective duality and a Chern-Mather involution. ArXiv e-prints 1601.05427, January 2016.
  • [3] Paolo Aluffi. Shadows of blow-up algebras. Tohoku Math. J. (2), 56(4):593–619, 2004.
  • [4] Tom Braden and Mikhail Grinberg. Perverse sheaves on rank stratifications. Duke Math. J., 96(2):317–362, 1999.
  • [5] J. P. Brasselet. Local Euler obstruction, old and new. In XI Brazilian Topology Meeting (Rio Claro, 1998), pages 140–147. World Sci. Publ., River Edge, NJ, 2000.
  • [6] J.-P. Brasselet and M.-H. Schwartz. Sur les classes de Chern d’un ensemble analytique complexe. In The Euler-Poincaré characteristic (French), volume 82 of Astérisque, pages 93–147. Soc. Math. France, Paris, 1981.
  • [7] J.-P. Brasselet, Lê D ung Tráng, and J. Seade. Euler obstruction and indices of vector fields. Topology, 39(6):1193–1208, 2000.
  • [8] Jean-Paul Brasselet and Nivaldo G. Grulha, Jr. Local Euler obstruction, old and new, II. In Real and complex singularities, volume 380 of London Math. Soc. Lecture Note Ser., pages 23–45. Cambridge Univ. Press, Cambridge, 2010.
  • [9] P. Bressler, M. Finkelberg, and V. Lunts. Vanishing cycles on Grassmannians. Duke Math. J., 61(3):763–777, 1990.
  • [10] Jean-Luc Brylinski, Alberto S. Dubson, and Masaki Kashiwara. Formule de l’indice pour modules holonomes et obstruction d’Euler locale. C. R. Acad. Sci. Paris Sér. I Math., 293(12):573–576, 1981.
  • [11] Alexandru Dimca. Sheaves in topology. Universitext. Springer-Verlag, Berlin, 2004.
  • [12] Alexandru Dimca and Stefan Papadima. Hypersurface complements, Milnor fibers and higher homotopy groups of arrangments. Ann. of Math. (2), 158(2):473–507, 2003.
  • [13] Alberto S. Dubson. Formule pour l’indice des complexes constructibles et des Modules holonomes. C. R. Acad. Sci. Paris Sér. I Math., 298(6):113–116, 1984.
  • [14] David Eisenbud. Linear sections of determinantal varieties. Amer. J. Math., 110(3):541–575, 1988.
  • [15] Lars Ernström. Topological Radon transforms and the local Euler obstruction. Duke Math. J., 76(1):1–21, 1994.
  • [16] William Fulton. Intersection theory. Springer-Verlag, Berlin, 1984.
  • [17] Terence Gaffney, Nivaldo G. Grulha, Jr., and Maria A. S. Ruas. The local Euler obstruction and topology of the stabilization of associated determinantal varieties. Math. Z., 291(3-4):905–930, 2019.
  • [18] Gerardo González-Sprinberg. L’obstruction locale d’Euler et le théorème de MacPherson. 83:7–32, 1981.
  • [19] Mark Goresky and Robert MacPherson. Intersection homology theory. Topology, 19(2):135–162, 1980.
  • [20] Mark Goresky and Robert MacPherson. Intersection homology. II. Invent. Math., 72(1):77–129, 1983.
  • [21] S. M. Guseĭ n Zade and V. Èbeling. On the indices of 1-forms on determinantal singularities. Tr. Mat. Inst. Steklova, 267(Osobennosti i Prilozheniya):119–131, 2009.
  • [22] Joe Harris and Loring W. Tu. On symmetric and skew-symmetric determinantal varieties. Topology, 23(1):71–84, 1984.
  • [23] June Huh. Milnor numbers of projective hypersurfaces and the chromatic polynomial of graphs. J. Amer. Math. Soc., 25(3):907–927, 2012.
  • [24] Jose Israel Rodriguez and Botong Wang. Computing Euler obstruction functions using maximum likelihood degrees. arXiv e-prints, page arXiv:1710.04310, October 2017.
  • [25] Benjamin F. Jones. Singular Chern classes of Schubert varieties via small resolution. Int. Math. Res. Not. IMRN, (8):1371–1416, 2010.
  • [26] Masaki Kashiwara. Index theorem for a maximally overdetermined system of linear differential equations. Proc. Japan Acad., 49:803–804, 1973.
  • [27] Masaki Kashiwara. Systems of microdifferential equations, volume 34 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1983. Based on lecture notes by Teresa Monteiro Fernandes translated from the French, With an introduction by Jean-Luc Brylinski.
  • [28] Masaki Kashiwara and Pierre Schapira. Sheaves on manifolds, volume 292 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1990. With a chapter in French by Christian Houzel.
  • [29] Gary Kennedy. MacPherson’s Chern classes of singular algebraic varieties. Comm. Algebra, 18(9):2821–2839, 1990.
  • [30] Alain Lascoux. Classes of C hern of a tensor product. CR Acad. Sci. Paris S  ’e r. AB, 286(8):A385 – A387, 1978.
  • [31] R. D. MacPherson. Chern classes for singular algebraic varieties. Ann. of Math. (2), 100:423–432, 1974.
  • [32] Leonardo C. Mihalcea and Rahul Singh. Mather classes and conormal spaces of Schubert varieties in cominuscule spaces. arXiv e-prints, page arXiv:2006.04842, June 2020.
  • [33] Bernt Ivar Utstø l Nø dland. Local Euler obstructions of toric varieties. J. Pure Appl. Algebra, 222(3):508–533, 2018.
  • [34] Sutipoj Promtapan. Chern-Schwartz-MacPherson Classes of Symmetric and Skew-symmetric Determinantal Varieties. PhD thesis, University of North Carolina at Chapel Hill, 2019.
  • [35] Sutipoj Promtapan and Richard Rimanyi. Characteristic classes of symmetric and skew-symmetric degeneracy loci. Facets of Algebraic Geometry: A Volume in Honour of William Fulton’s 80th birthday, Cambridge Univ. Press LMS Lecture Note Series, 2020.
  • [36] Joerg Schuermann. A short proof of a formula of Brasselet, Le and Seade for the Euler obstruction. arXiv Mathematics e-prints, page math/0201316, January 2002.
  • [37] E. Tevelev. Projectively Dual Varieties. ArXiv Mathematics e-prints, December 2001.
  • [38] Lê D ung Tráng and Bernard Teissier. Variétés polaires locales et classes de Chern des variétés singulières. Ann. of Math. (2), 114(3):457–491, 1981.
  • [39] Xiping Zhang. Local Euler Obstruction and Chern-Mather classes of Determinantal Varieties. arXiv e-prints, page arXiv:1706.02032, June 2017.
  • [40] Xiping Zhang. Chern classes and characteristic cycles of determinantal varieties. J. Algebra, 497:55–91, 2018.