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

On the topology of determinantal links

Matthias Zach
Abstract.

We study the cohomology of the generic determinantal varieties Mm,ns={φm×n:rankφ<s}M_{m,n}^{s}=\{\varphi\in\mathbb{C}^{m\times n}:\operatorname{rank}\varphi<s\}, their polar multiplicities, their sections DkMm,nsD_{k}\cap M_{m,n}^{s} by generic hyperplanes DkD_{k} of various codimension kk, and the real and complex links of the spaces (DkMm,ns,0)(D_{k}\cap M_{m,n}^{s},0). Such complex links were shown to provide the basic building blocks in a bouquet decomposition for the (determinantal) smoothings of smoothable isolated determinantal singularities. The detailed vanishing topology of such singularities was still not fully understood beyond isolated complete intersections and a few further special cases. Our results now allow to compute all distinct Betti numbers of any determinantal smoothing.

1. Introduction and results

The motivation for this paper is to understand the vanishing (co-)homology of isolated determinantal singularities (X,0)(p,0)(X,0)\subset(\mathbb{C}^{p},0) (abbreviated as IDS in the following) which admit a determinantal smoothing MM. Despite their seemingly odd definition (see below), they are quite frequently encountered; for instance any normal surface singularity in (4,0)(\mathbb{C}^{4},0) and any so-called “space curve” (C,0)(3,0)(C,0)\subset(\mathbb{C}^{3},0) is determinantal by virtue of the Hilbert-Burch theorem, see e.g. [Wah16] and [Sch77]. It was shown in [Zac20] that sections of the generic determinantal varieties Mm,ns={φm×n:rankφ<s}M_{m,n}^{s}=\{\varphi\in\mathbb{C}^{m\times n}:\operatorname{rank}\varphi<s\} by hyperplanes Dkm×nD_{k}^{\prime}\subset\mathbb{C}^{m\times n}, which are of codimension kk and in general position off the origin, provide the fundamental building blocks in a bouquet decomposition of the determinantal smoothing MM.

In this paper we aim to study these building blocks

k(Mm,ns,0)Mm,nsDk+1,\mathcal{L}^{k}(M_{m,n}^{s},0)\cong M_{m,n}^{s}\cap D^{\prime}_{k+1},

which we shall refer to as complex links of codimension kk, and discuss the implications for the vanishing topology of various singularities. Our main result in this regard is that whenever kk is sufficiently big so that the hyperplane Dk+1D^{\prime}_{k+1} misses the singular locus of Mm,nsM_{m,n}^{s} and k(Mm,ns,0)\mathcal{L}^{k}(M_{m,n}^{s},0) is smooth, then the cohomology in degrees ii below the middle degree d=dimk(Mm,ns,0)d=\dim\mathcal{L}^{k}(M_{m,n}^{s},0) is

Hi(k(Mm,ns,0))Hi(Grass(s1,m)),H^{i}(\mathcal{L}^{k}(M_{m,n}^{s},0))\cong H^{i}(\operatorname{Grass}(s-1,m)),

that is, it is isomorphic to the cohomology of a Grassmannian, see (37).

In the general case, i.e. for arbitrary codimension kk, we provide a formula (11) for the Euler characteristic of k(Mm,ns,0)\mathcal{L}^{k}(M_{m,n}^{s},0) based on the polar methods developed by Lê and Teissier. We proceed to study the polar multiplicities em,nr,ke_{m,n}^{r,k} of the generic determinantal varieties (Mm,nr+1,0)(M_{m,n}^{r+1},0) in Section 2.4 and obtain a formula (21) for these numbers as an integral of Segre classes. It should be noted that this formula does not depend on the choice of any generic hyperplane section and it can be implemented effectively in a computer algebra system. However, we were unable to prove a closed formula for the numbers em,nr,ke_{m,n}^{r,k} as a function of m,n,rm,n,r, and kk; even though the calculated Tables 1, and 37 yield several patterns.

Coming back to the case of smooth links, this allows then to also compute the middle Betti number of k(Mm,ns,0)\mathcal{L}^{k}(M_{m,n}^{s},0) (but unfortunately not the full cohomology group with integer coefficients in general, i.e. there could also be torsion).

We would like to point out that many methods developed in this article apply much more generally to compute the cohomology of the links of higher codimension for spaces (X,0)(q,0)(X,0)\subset(\mathbb{C}^{q},0) which decompose into orbits of a Lie group action G×XXG\times X\to X. The generic determinantal varieties treated here are just one particular, accessible example.

We continue with the discussion of the implications of these results for arbitrary smoothable IDS. Let

A:(p,0)(m×n,0)A\colon(\mathbb{C}^{p},0)\to(\mathbb{C}^{m\times n},0)

be a holomorphic map germ to the space of m×nm\times n-matrices. We say that (X,0)(X,0) is determinantal for AA and of type (m,n,s)(m,n,s) if we have (X,0)=(A1(Mm,ns),0)(X,0)=(A^{-1}(M_{m,n}^{s}),0) and such that codim(X,0)=codim(Mm,ns,0)=(ms+1)(ns+1)\operatorname{codim}(X,0)=\operatorname{codim}(M_{m,n}^{s},0)=(m-s+1)(n-s+1). In this case we also write (X,0)=(XAs,0)(X,0)=(X_{A}^{s},0) to emphasize the chosen determinantal structure for (X,0)(X,0).

By definition, a determinantal deformation is induced by an unfolding 𝐀(x,t)\mathbf{A}(x,t) of A(x)A(x) on parameters t=(t1,,tk)t=(t_{1},\dots,t_{k}) such that A0=𝐀(,0)A_{0}=\mathbf{A}(-,0) is equal to the original map AA and At=𝐀(,t)A_{t}=\mathbf{A}(-,t) is transverse to Mm,nsM_{m,n}^{s} in a stratified sense for t0kt\neq 0\in\mathbb{C}^{k}. Then Xt=At1(Mm,ns)X_{t}=A_{t}^{-1}(M_{m,n}^{s}) is a smoothing of (X,0)(X,0), provided that pp is strictly smaller than the codimension c=(ms2)(ns2)c=(m-s-2)(n-s-2) of the singular locus of Mm,nsM_{m,n}^{s}. In this case we shall speak of the determinantal smoothing111Note that in the case of an isolated determinantal hypersurface singularity given by (X,0)=({detA=0},0)(p,0)(X,0)=(\{\det A=0\},0)\subset(\mathbb{C}^{p},0) a smoothing of the singularity always exists, but it can be realized by a determinantal deformation only when p<4p<4. of (XAs,0)(X_{A}^{s},0) and it can be shown that, up to diffeomorphism, this smoothing depends only on the choice of the determinantal structure for (X,0)(X,0), but not on the particular unfolding of the matrix. We will therefore also write MAsM_{A}^{s} for the determinantal smoothing XtX_{t} of (X,0)(X,0).

Note that every complete intersection singularity (X,0)(p,0)(X,0)\subset(\mathbb{C}^{p},0) of codimension kk is determinantal of type (1,k,1)(1,k,1) for the matrix AA whose entries are given by a regular sequence generating the ideal of (X,0)(X,0). In this case the determinantal deformations coincide with the usual deformations of the singularity. We refer to [FKZ21] for the details on determinantal singularities and their deformations and further references for the above statements.

The aforementioned bouquet decomposition of the determinantal smoothing MAsM_{A}^{s} of an IDS (X,0)=(XAs,0)(X,0)=(X_{A}^{s},0) from [Zac20] reads

(1) MAshtmnp1(Mm,ns,0)i=1λSdim(XAs,0)M_{A}^{s}\cong_{\mathrm{ht}}\mathcal{L}^{mn-p-1}(M_{m,n}^{s},0)\vee\bigvee_{i=1}^{\lambda}S^{\dim(X_{A}^{s},0)}

where we denote by SdS^{d} the unit sphere of real dimension dd. Combining this with our results we find

Theorem 1.1.

Let MAsM_{A}^{s} be the determinantal smoothing of a dd-dimensional smoothable isolated determinantal singularity (XAs,0)(p,0)(X_{A}^{s},0)\subset(\mathbb{C}^{p},0) of type (m,n,s)(m,n,s) with smns\leq m\leq n. Then the truncated cohomology

Hd(Grass(s1,m))H(MAs)H^{\leq d}(\operatorname{Grass}(s-1,m))\subset H^{\bullet}(M_{A}^{s})

embeds into the cohomology of MAsM_{A}^{s} with a quotient concentrated in cohomological degree dd. Moreover, the left hand side is generated as an algebra by the Segre classes of the vector bundle EE on MAsM_{A}^{s} whose fiber at a point xx is presented by the perturbed matrix

nAt(x)mEx0\mathbb{C}^{n}\overset{A_{t}(x)}{\longrightarrow}\mathbb{C}^{m}\to E_{x}\to 0

which defines the smoothing MAsM_{A}^{s}.

m\km\backslash k 0 1 2 3 4 5 6 7 8 9
1 1 0 0 0 0 0 0 0 0 0
2 3 4 3 0 0 0 0 0 0 0
3 6 16 27 24 10 0 0 0 0 0
4 10 40 105 176 190 120 35 0 0 0
5 15 80 285 696 1200 1440 1155 560 126 0
6 21 140 630 2016 4760 8352 10815 10080 6426 2520
7 28 224 1218 4816 14420 33216 59143 80976 83916 63840
8 36 336 2142 10080 36540 104112 235557 424384 606564 680400
9 45 480 3510 19152 81480 276480 758205 1691920 3077838 4551840
10 55 660 5445 33792 165000 649440 2091705 5563360 12278970 22518200
m\km\backslash k 10 11 12 13 14 15 16 17 18 19
6 462 0 0 0 0 0 0 0 0 0
7 33726 11088 1716 0 0 0 0 0 0 0
8 587202 376992 169884 48048 6435 0 0 0 0 0
9 5430810 5155920 3809520 2114112 830115 205920 24310 0 0 0
10 34240800 42926400 43929600 36132096 23326875 11394240 3962530 875160 92378 0
Table 1. Polar multiplicities em,m+1m1,ke_{m,m+1}^{m-1,k} for the generic determinantal varieties (Mm,m+1m,0)(M_{m,m+1}^{m},0) appearing in the context of the Hilbert-Burch theorem

In the case of smoothable isolated Cohen-Macaulay codimension 22 singularities (X,0)(p,0)(X,0)\subset(\mathbb{C}^{p},0) we obtain some more specific results which might be of particular interest. As remarked earlier, these singularities admit a canonical determinantal structure by virtue of the Hilbert-Burch theorem (see [Hil90], [Bur68], or [Eis95] for a textbook): Let f0,,fm{x1,,xp}f_{0},\dots,f_{m}\in\mathbb{C}\{x_{1},\dots,x_{p}\} be a minimal set of generators for the ideal II defining (X,0)(X,0). Then the minimal free resolution of 𝒪X,0={x1,,xp}/I\mathcal{O}_{X,0}=\mathbb{C}\{x_{1},\dots,x_{p}\}/I over the ring 𝒪p={x1,,xp}\mathcal{O}_{p}=\mathbb{C}\{x_{1},\dots,x_{p}\} takes the form

0𝒪pmAT𝒪pm+1𝑓𝒪p𝒪X,000\to\mathcal{O}_{p}^{m}\overset{A^{T}}{\longrightarrow}\mathcal{O}_{p}^{m+1}\overset{f}{\longrightarrow}\mathcal{O}_{p}\to\mathcal{O}_{X,0}\to 0

for some matrix A𝒪m×(m+1)A\in\mathcal{O}^{m\times(m+1)} and then (X,0)=(XAm,0)(X,0)=(X_{A}^{m},0) is determinantal of type (m,m+1,m)(m,m+1,m) for the matrix AA. Moreover, any deformation of (X,0)(X,0) is also automatically determinantal (see [Sch77]), i.e. it arises from a perturbation of AA. Isolated Cohen-Macaulay codimension 22 singularities exist up to dimension d=dim(X,0)4d=\dim(X,0)\leq 4 and they are smoothable if and only if this inequality is strict.

The generic determinantal varieties appearing in the context of the Hilbert-Burch theorem are naturally limited to Mm,m+1mM_{m,m+1}^{m} and we list their polar multiplicities for values 0m100\leq m\leq 10 in Table 1. From these numbers we can then also compute the Euler characteristic of the smooth complex links of their generic hyperplane sections, listed in Table 2.

d\md\backslash m 1 2 3 4 5 6 7 8 9 10
0 1 3 6 10 15 21 28 36 45 55
1 0 -1 -10 -30 -65 -119 -196 -300 -435 -605
2 0 2 17 75 220 511 1022 1842 3075 4840
3 0 2 -7 -101 -476 -1505 -3794 -9138 -16077 -28952
Table 2. The Euler characteristic of the smooth complex links m(m+1)d3(Mm,m+1m,0)\mathcal{L}^{m(m+1)-d-3}(M_{m,m+1}^{m},0) of dimension dd of the generic determinantal varieties appearing in the context of the Hilbert-Burch theorem

As a consequence of Theorem 1.1 we obtain:

Corollary 1.2.

Let (X,0)(5,0)(X,0)\subset(\mathbb{C}^{5},0) be a Cohen-Macaulay germ of codimension 22 with an isolated singularity at the origin and MM its smoothing. Then H2(M)H^{2}(M)\cong\mathbb{Z} is free of rank one and generated by the Chern class of the canonical bundle on MM.

This has been conjectured by the author in [Zac18] where a special case of Corollary 1.2 was obtained for those singularities defined by 2×32\times 3-matrices. Our results on the Euler characteristic of the complex links also give lower bounds for the third Betti number

(2) b3χ(m(m+1)6(Mm,m+1m,0))2b_{3}\geq-\chi\left(\mathcal{L}^{m(m+1)-6}(M_{m,m+1}^{m},0)\right)-2

which can be read off directly from Table 2 depending on the size of the matrix mm. Unfortunately, we were unable to prove a closed algebraic formula for this number as a function of mm.

While in the case of threefolds the cohomology of the complex link prominently sticks out in the sense that it entirely covers the nontrivial contributions outside the middle degree, the situation is a little more hidden in dimensions d<3d<3 since then all the non-trivial (reduced) cohomology of the smoothing MM is concentrated in the middle degree. Nevertheless, using Theorem 1.1, a lower bound for the middle Betti number can be read off from Table 2 as

(3) b2χ(m(m+1)5(Mm,m+1m,0))2b_{2}\geq\chi\left(\mathcal{L}^{m(m+1)-5}(M_{m,m+1}^{m},0)\right)-2

for smoothings of isolated Cohen-Macaulay codimension 22 surfaces and

(4) b1χ(m(m+1)4(Mm,m+1m,0))1b_{1}\geq-\chi\left(\mathcal{L}^{m(m+1)-4}(M_{m,m+1}^{m},0)\right)-1

in the case of space curves in (3,0)(\mathbb{C}^{3},0). The proofs of Theorem 1.1 and Corollary 1.2 will be given in the last section.

Remark 1.3.

Wahl has shown in [Wah16] that for a normal surface singularity (X,0)(4,0)(X,0)\subset(\mathbb{C}^{4},0) which is not Gorenstein, the second Betti number μ\mu of the smoothing MM and the Tjurina number τ\tau of (X,0)(X,0) obey an inequality

μτ1\mu\geq\tau-1

with equality whenever (X,0)(X,0) is quasi-homogeneous. In particular, this applies to all generic hyperplane sections (X,0)=(DMm,m+1m,0)(D,0)(X,0)=(D\cap M_{m,m+1}^{m},0)\subset(D,0) for m>1m>1 and DD of dimension 44. It follows from the above that χ(M)=τ\chi(M)=\tau and the first few values can be read off from Table 2. Comparing Wahl’s results with the bouquet decomposition (1) we see that for non-linear quasi-homogeneous singularities, the difference of the numbers τ\tau and λ\lambda is determined by the offset

τλ=χ(m(m+1)5(Mm,m+1m,0))\tau-\lambda=\chi\left(\mathcal{L}^{m(m+1)-5}(M_{m,m+1}^{m},0)\right)

depending only on the size of the matrix m>1m>1. This is different compared to the case of isolated complete intersections (the case m=1m=1) where we have equalities τ=μ=λ\tau=\mu=\lambda; see [Wah85].

2. Real and complex links of higher codimension

2.1. Definitions

Let XUnX\hookrightarrow U\subset\mathbb{C}^{n} be the closed embedding of an equidimensional reduced complex analytic variety of dimension dd in some open set UU of n\mathbb{C}^{n} and suppose {Vα}αA\{V^{\alpha}\}_{\alpha\in A} is a complex analytic Whitney stratification for XX. Such stratifications always exist; a construction of a unique minimal stratification has been described in [TT81] and if not specified further, we will in the following always assume XX to be endowed with its minimal Whitney stratification. Moreover, we may suppose that every stratum is connected for if this was not the case, we could replace it by its distinct irreducible components.

Recall the notions of the real and complex links of XX along its strata. These can be defined as follows: Let VαV^{\alpha} be a stratum of XX and xVαXx\in V^{\alpha}\subset X an arbitrary point. For simplicity, we may assume x=0nx=0\in\mathbb{C}^{n} to be the origin. Choose a normal slice to VαV^{\alpha} through xx, i.e. a submanifold (Nx,x)(n,x)(N_{x},x)\subset(\mathbb{C}^{n},x) of complementary dimension which meets VαV^{\alpha} transversally in xx. Let Bε(x)B_{\varepsilon}(x) be the closed ball of radius ε\varepsilon centered at xx. Then the real link of XX along the stratum VαV^{\alpha} is

(5) 𝒦(X,Vα)=XNxBε(x)\mathcal{K}(X,V^{\alpha})=X\cap N_{x}\cap\partial B_{\varepsilon}(x)

for 1ε>01\gg\varepsilon>0 sufficiently small. Similarly, the complex link is

(6) (X,Vα)=XNxBε(x)l1({δ})\mathcal{L}(X,V^{\alpha})=X\cap N_{x}\cap B_{\varepsilon}(x)\cap l^{-1}(\{\delta\})

for a sufficiently general linear form ll and 1ε|δ|>01\gg\varepsilon\gg|\delta|>0. For a rigorous definition of these objects see for example [GM88, Part I, Chapter 1.4] and [GM88, Part II, Chapter 2.2]. There, one can also find a proof of the fact that the real and complex links are independent of the choices involved in their definition: They are invariants of the particular stratification of XX and unique up to non-unique homeomorphism.

In this note we shall also be concerned with the real and complex links of higher codimension for a germ (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0) at the point 0.

Definition 2.1.

The real and complex links of codimension ii of (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0) at the origin are the classical real and complex links of a section (XDi,0)(X\cap D_{i},0) of XX with a sufficiently general plane DinD_{i}\subset\mathbb{C}^{n} of codimension ii.

Note that in this definition, the classical complex link (X,0)\mathcal{L}(X,0) is the complex link of codimension 0, even though it has complex codimension 11 as an analytic subspace of XX. This choice was made in order to be compatible with the notation for the real links. The reader may think of the word “link” as an indicator to increment the codimension by one to arrive at the actual codimension of the space.

By “sufficiently general” we mean that DiD_{i} has to belong to a Zariski open subset UiGrass(ni,n)U_{i}\subset\operatorname{Grass}(n-i,n) in the Grassmannian. This set consists of planes for which variation of DiD_{i} results in a Whitney equisingular family in XDiX\cap D_{i} and we shall briefly discuss the existence of such a set. From this it will follow immediately that the real and complex links of higher codimension are invariants of the germ (X,0)(X,0) itself and unique up to non-unique homeomorphism.

Consider the Grassmann modification of XX:

GiX={(x,D)X×Grass(ni,n):xD}G_{i}X=\{(x,D)\in X\times\operatorname{Grass}(n-i,n):x\in D\}

where by Grass(ni,n)\operatorname{Grass}(n-i,n) we denote the Grassmannian of codimension ii planes in the ambient space n\mathbb{C}^{n} through the point 0. It comes naturally with two projections

(7) GiX\textstyle{G_{i}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ\scriptstyle{\rho}π\scriptstyle{\pi}Grass(ni,n)\textstyle{\operatorname{Grass}(n-i,n)}X\textstyle{X}

where by construction ρ(π1({D}))=XD\rho(\pi^{-1}(\{D\}))=X\cap D for any plane DGrass(ni,n)D\in\operatorname{Grass}(n-i,n). The Grassmannian itself has a natural embedding into GiXG_{i}X as the zero section E0E_{0} of π\pi and we may think of E0E_{0} as parametrizing the intersections of XX with planes of codimension ii. For id=dimXi\leq d=\dim X the intersection of any plane DD with (X,0)(X,0) will be of positive dimension at 0 and therefore E0E_{0} is necessarily contained in the closure of its complement.

Note that for an arbitrary point pXp\in X the fiber ρ1({p})\rho^{-1}(\{p\}) over pp in GiXG_{i}X consists of all planes DGrass(ni,n)D\in\operatorname{Grass}(n-i,n) containing both pp and the origin. It follows that the restriction

ρ:GiXE0X{0}\rho\colon G_{i}X\setminus E_{0}\to X\setminus\{0\}

is not an isomorphism, but nevertheless a holomorphic fiber bundle with smooth fibers. In particular, any given Whitney stratification on XX determines a unique Whitney stratification of GiXE0G_{i}X\setminus E_{0} by pullback of the strata.

Given such a stratification on the complement of E0E_{0}, there exists some maximal Zariski open subset UiE0Grass(ni,n)U_{i}\subset E_{0}\cong\operatorname{Grass}(n-i,n) such that all the strata of GiXE0G_{i}X\setminus E_{0} satisfy Whitney’s conditions (a) and (b) along UiU_{i}. We may extend the stratification on the open subset GiXE0G_{i}X\setminus E_{0} to a Whitney stratification of GiXG_{i}X, containing UiU_{i} as a stratum. Since the Grassmannian Grass(ni,n)\operatorname{Grass}(n-i,n) is irreducible and complex analytic subsets have real codimension 2\geq 2, this stratum is necessarily dense and connected.

A plane DD of codimension ii is sufficiently general in the sense of Definition 2.1 if it belongs to UiU_{i}. The projection π\pi provides us with a canonical choice of normal slices along UiU_{i} and by construction, we may therefore identify the real and complex links of GiXG_{i}X along UiU_{i} with the real and complex links of the germ (XD,0)(n,0)(X\cap D,0)\subset(\mathbb{C}^{n},0) for any DUiD\in U_{i}. Given that the classical real and complex links are invariants of the strata in a Whitney stratification, it follows immediately that the real and complex links of higher codimension for a germ (X,0)(X,0) are well defined invariants of the germ itself and unique up to non-unique homeomorphism.

2.2. Synopsis with polar varieties

While it was already established in the previous discussion that the real and complex links of higher codimension are invariants of the germ (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0) itself, a more specific setup will be required in the following. Let l¯=l1,l2,,ldHom(n,)\underline{l}=l_{1},l_{2},\dots,l_{d}\in\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C}) be a sequence of linear forms defining a flag of subspaces

𝒟:n=D0D1D2Dd{0}\mathcal{D}:\mathbb{C}^{n}=D_{0}\supset D_{1}\supset D_{2}\supset\dots\supset D_{d}\supset\{0\}

in n\mathbb{C}^{n} via Di+1=Dikerli+1D_{i+1}=D_{i}\cap\ker l_{i+1}.

Definition 2.2.

Let (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0) be an equidimensional reduced complex analytic germ of dimension dd and {Vα}αA\{V^{\alpha}\}_{\alpha\in A} a Whitney stratification of XX. A sequence of linear forms l1,l2,,ldHom(n,)l_{1},l_{2},\dots,l_{d}\in\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C}) is called admissible for (X,0)(X,0) if for every 0<id0<i\leq d the restriction of lil_{i} to Xi1:=XDi1X_{i-1}:=X\cap D_{i-1} does not annihilate any limiting tangent space of (Xi1,0)(X_{i-1},0) at the origin.

This particular definition is chosen with a view towards the fibration theorems and inductive arguments used in Section 4.1. However, we shall also need the results on polar varieties by Lê and Teissier from [TT81]. To this end, we recall the necessary definitions and explain how Definition 2.2 fits into the context of their paper.

The geometric setup for the treatment of limiting tangent spaces is the Nash modification. Let XUX\subset U be a suitable representative of (X,0)(X,0) in some open neighborhood UU of the origin. The Gauss map is defined on the regular locus XregX_{\mathrm{reg}} via

XregGrass(d,n),p[TpXregTpn].X_{\mathrm{reg}}\to\operatorname{Grass}(d,n),\quad p\mapsto[T_{p}X_{\mathrm{reg}}\subset T_{p}\mathbb{C}^{n}].

Then the Nash blowup X~\tilde{X} of XX is defined as the closure of the graph of the Gauss map. It comes with two morphisms

X~\textstyle{\tilde{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ν\scriptstyle{\nu}γ\scriptstyle{\gamma}Grass(d,n)\textstyle{\operatorname{Grass}(d,n)}X\textstyle{X}

where ν\nu is the projection to XX and γ\gamma the natural prolongation of the Gauss map. We denote by T~\tilde{T} the pullback of the tautological bundle on Grass(d,n)\operatorname{Grass}(d,n) along γ\gamma and by Ω~1\tilde{\Omega}^{1} the dual of T~\tilde{T}. By construction, the fiber ν1({p})\nu^{-1}(\{p\}) over a point pXp\in X consists of pairs (p,E)X~n×Grass(d,n)(p,E)\in\tilde{X}\subset\mathbb{C}^{n}\times\operatorname{Grass}(d,n) where EE is a limiting tangent space to XX at pp. In particular, such a limiting tangent space is unique at regular points pXregp\in X_{\mathrm{reg}} and ν:ν1(Xreg)Xreg\nu\colon\nu^{-1}(X_{\mathrm{reg}})\to X_{\mathrm{reg}} is an isomorphism identifying the restriction of T~\tilde{T} with the tangent bundle of XregX_{\mathrm{reg}}.

A flag 𝒟\mathcal{D} as above determines a set of degeneraci loci in the Grassmannian as follows. Every linear form lHom(n,)l\in\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C}) can be pulled back to a global section νl\nu^{*}l in the dual of the tautological bundle of Grass(d,n)\operatorname{Grass}(d,n): On a fiber EE of the tautological bundle the section νl\nu^{*}l is defined to be merely the restriction of ll to EE regarded as a linear subspace of n\mathbb{C}^{n}. This generalizes in the obvious way for the linear maps

φk:=l1l2ldk+1:ndk+1\varphi_{k}:=l_{1}\oplus l_{2}\oplus\dots\oplus l_{d-k+1}\colon\mathbb{C}^{n}\to\mathbb{C}^{d-k+1}

for every 0kd0\leq k\leq d. Adapting the notation from [TT81] we now have

ck(𝒟)\displaystyle c_{k}(\mathcal{D}) :=\displaystyle:= {EGrass(d,n):dimEDdk+1k}\displaystyle\{E\in\operatorname{Grass}(d,n):\dim E\cap D_{d-k+1}\geq k\}
=\displaystyle= {EGrass(d,n):rankνφk<dk+1}.\displaystyle\{E\in\operatorname{Grass}(d,n):\operatorname{rank}\nu^{*}\varphi_{k}<d-k+1\}.

Note that by construction c0(𝒟)=Grass(d,n)c_{0}(\mathcal{D})=\operatorname{Grass}(d,n) is the whole Grassmannian. We set γ1(ck(𝒟))X~\gamma^{-1}(c_{k}(\mathcal{D}))\subset\tilde{X} to be the corresponding degeneraci locus on the Nash modification and

Pk(𝒟):=ν(γ1(ck(𝒟)))P_{k}(\mathcal{D}):=\nu(\gamma^{-1}(c_{k}(\mathcal{D})))

its image in XX. By construction, a point pXp\in X belongs to Pk(𝒟)P_{k}(\mathcal{D}) if and only if there exists a limiting tangent space EE to XregX_{\mathrm{reg}} at pp such that the restriction of φk\varphi_{k} to EE does not have full rank.

The kk-th polar multiplicity of (X,0)(X,0) is then defined for 0k<d0\leq k<d to be

(8) mk(X,0):=m0(Pk(𝒟)),m_{k}(X,0):=m_{0}(P_{k}(\mathcal{D})),

the multiplicity of the kk-th polar variety. We shall see below that for k=dk=d the variety Pd(𝒟)P_{d}(\mathcal{D}) is empty for a generic flag so that a definition of md(X,0)m_{d}(X,0) does not make sense. In the other extreme case where k=0k=0, the polar multiplicity m0(X,0)m_{0}(X,0) is simply the multiplicity of (X,0)(X,0) itself.

Lemma 2.3.

Let XX be a sufficiently small representative of (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0). A sequence of linear forms l1,,ldl_{1},\dots,l_{d} is admissible for (X,0)(X,0) if and only if for every 0i<d0\leq i<d one has

(9) γ1(cdi(𝒟))X^i=\gamma^{-1}(c_{d-i}(\mathcal{D}))\cap\widehat{X}_{i}=\emptyset

for the associated flag 𝒟\mathcal{D}. Here Xi:=XDiX_{i}:=X\cap D_{i} and X^i\widehat{X}_{i} denotes the strict transform of XiX_{i} in the Nash modification.

Proof.

The proof will proceed by induction on ii. For i=0i=0 we have X^0=X~\widehat{X}_{0}=\tilde{X} and the statement is about the choice of l1l_{1}. Consider the projectivized analytic set of degenerate covectors

{([l],(p,E))Hom(n,)×ν1({0}):l|E=0}\{([l],(p,E))\in\mathbb{P}\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C})\times\nu^{-1}(\{0\}):l|_{E}=0\}

along the central fiber ν1({0})\nu^{-1}(\{0\}) of the Nash modification. Since this fiber has strictly smaller dimension than XX, the set of degenerate covectors has dimension <n1<n-1 and therefore the discriminant, i.e. the image of its projection to Hom(n,)\mathbb{P}\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C}), is a closed analytic set of positive codimension. Now (9) is satisfied if and only if l1l_{1} belongs to the complement of the affine cone of the discriminant.

Such a choice for l1l_{1} determines X1=XD1X_{1}=X\cap D_{1}. Note that by construction this intersection is transversal and therefore X1X_{1} inherits a Whitney stratification from the one on XX. Moreover, at a regular point pX1p\in X_{1} the tangent space

TpX1=kerνl1TpXT_{p}X_{1}=\ker\nu^{*}l_{1}\subset T_{p}X

is naturally contained in the tangent space of XX to that point. Let X^1X~\widehat{X}_{1}\subset\tilde{X} be the strict transform of X1X_{1} in the Nash modification. Taking limits of appropriate (sub-)sequences of regular points, it is easy to see that every limiting tangent space EE^{\prime} of X1X_{1} at 0 is contained in a limiting tangent space EE of XX along X1X_{1}. Consequently, a second linear form l2l_{2} annihilates the limiting tangent space EE^{\prime} of X1X_{1} if and only if l1l2l_{1}\oplus l_{2} is degenerate on EE. But this means nothing else than

(0,E)X^1γ1(cd1(𝒟))(0,E)\in\widehat{X}_{1}\cap\gamma^{-1}(c_{d-1}(\mathcal{D}))\neq\emptyset

which establishes the claim for i=1i=1. The remainder of the induction is a repetition of the previous steps and left to the reader. ∎

The previous lemma provides the link of Definition 2.2 with the “Théorème de Bertini idéaliste” by Lê and Teissier [TT81, Théorème 4.1.3]. They establish the existence of Zariski open subsets UiGrass(ni,n)U^{\prime}_{i}\subset\operatorname{Grass}(n-i,n) with certain good properties concerning the variety γ1(cdi+1(Di))\gamma^{-1}(c_{d-i+1}(D_{i})) for DiUiD_{i}\in U^{\prime}_{i}. A posteriori, they discuss in [TT81, Proposition 4.1.5] that if the whole flag 𝒟\mathcal{D} has been chosen such that DiUiD_{i}\subset U^{\prime}_{i} for all ii, then also (9) is in fact satisfied for all ii. Thus we obtain the following

Corollary 2.4.

For every equidimensional reduced analytic germ (X,0)(X,0) there exists a Zariski open and dense subset of admissible sequences of linear forms l1,l2,ldl_{1},l_{2}\dots,l_{d}. Moreover, this sequence can be chosen such that for the associated flag 𝒟\mathcal{D} the space Di={l1==li=0}D_{i}=\{l_{1}=\dots=l_{i}=0\} is sufficiently general in the sense of Definition 2.1 so that the real and complex links of codimension ii are given by

𝒦i(X,0)=XDiBε(0)\mathcal{K}^{i}(X,0)=X\cap D_{i}\cap\partial B_{\varepsilon}(0)

and

i(X,0)=XDiBε(0)li+11({δ})\mathcal{L}^{i}(X,0)=X\cap D_{i}\cap B_{\varepsilon}(0)\cap l_{i+1}^{-1}(\{\delta\})

for 1ε|δ|>01\gg\varepsilon\gg|\delta|>0, respectively.

Proof.

Consider the sets UiUiU^{\prime}_{i}\cap U_{i} with UiU^{\prime}_{i} from Lê’s and Teissier’s “Théorème de Bertini idéaliste” and UiU_{i} the set of Whitney equisingular sections from the discussion of Definition 2.1. Since the intersection of Zariski open sets is again Zariski open, we may choose l1,l2,,ldl_{1},l_{2},\dots,l_{d} to be any sequence of linear forms such that

Di={l1==li=0}UiUiD_{i}=\{l_{1}=\dots=l_{i}=0\}\subset U^{\prime}_{i}\cap U_{i}

for all ii. ∎

We will henceforth assume that the sequence of linear forms lil_{i} has been chosen such that the associated flag 𝒟\mathcal{D} has DiUiUiD_{i}\in U^{\prime}_{i}\cap U_{i} where UiU^{\prime}_{i} is the Zariski open subset of Lê’s and Teissiers’ “Théorème de Bertini idéaliste” and UiU_{i} the Zariski open subset of Whitney equisingular sections of codimension ii from the discussion of Definition 2.1.

2.3. The Euler characteristic of complex links

Lê and Teissier have described a method to compute the Euler characteristic of complex links from the polar multiplicities in [TT81, Proposition 6.1.8]. We briefly sketch how to use their results inductively in our setup from Definition 2.2.

As before, let (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0) be a reduced, equidimensional complex analytic germ of dimension dd, endowed with a Whitney stratification {Vα}αA\{V^{\alpha}\}_{\alpha\in A}. We will assume that V0={0}V^{0}=\{0\} is a stratum and write Xα=Vα¯X^{\alpha}=\overline{V^{\alpha}} for the closure of any other stratum VαV^{\alpha} of XX. Throughout this section, we will assume that an admissible sequence of linear forms l1,,ldl_{1},\dots,l_{d} and the associated flag 𝒟\mathcal{D} have been chosen as in Corollary 2.4 for all germs (Xα,0)(X^{\alpha},0) at once. This flag being fixed, we will in the following suppress it from our notation and simply write Pk(Xα,0)P_{k}(X^{\alpha},0) for the polar varieties of the germ (Xα,0)(X^{\alpha},0) with 𝒟\mathcal{D} being understood.

Denote by (X,Vα)\mathcal{L}(X,V^{\alpha}) the classical complex links of XX along the stratum VαV^{\alpha} and by i\mathcal{L}^{i} the complex link of codimension ii of XX at the origin. Then by [TT81, Théorème 6.1.9]

(10) χ(0)χ(1)=α0m0(Pd(α)1(Xα,0))(1)d(α)1(1χ((X,Vα))).\chi\left(\mathcal{L}^{0}\right)-\chi\left(\mathcal{L}^{1}\right)=\sum_{\alpha\neq 0}m_{0}\left(P_{d(\alpha)-1}\left(X^{\alpha},0\right)\right)\cdot(-1)^{d(\alpha)-1}\left(1-\chi(\mathcal{L}(X,V^{\alpha}))\right).

For our specific setup we may interpret this formula in the context of stratified Morse theory. To this end, note that the restriction of l2l_{2} to the complex link 0\mathcal{L}^{0}

l2:XBε(0)l11({δ1})l_{2}\colon X\cap B_{\varepsilon}(0)\cap l_{1}^{-1}(\{\delta_{1}\})\to\mathbb{C}

is a stratified Morse function with critical points on the interior of the strata Vα0V^{\alpha}\cap\mathcal{L}^{0} of 0\mathcal{L}^{0} precisely at the intersection points

0Pd(α)1(Xα,0)={q1α,,qm0(P(α)d1(Xα,0))α},\mathcal{L}^{0}\cap P_{d(\alpha)-1}(X^{\alpha},0)=\{q^{\alpha}_{1},\dots,q^{\alpha}_{m_{0}(P(\alpha)_{d-1}(X^{\alpha},0))}\},

cf. [TT81, Corollaire 4.1.6] and [TT81, Corollaire 4.1.9]. The complex link of codimension 22 can be identified with the general fiber

1l21({δ2})0\mathcal{L}^{1}\cong l_{2}^{-1}(\{\delta_{2}\})\cap\mathcal{L}^{0}

for some regular value δ2\delta_{2} off the discriminant and we can use the function λ=|l2δ2|2\lambda=|l_{2}-\delta_{2}|^{2} as a Morse function in order to reconstruct 0\mathcal{L}^{0} from 1\mathcal{L}^{1}. It can be shown that for suitable choices of the represenatives involved, the critical points of λ\lambda on the boundary of 0\mathcal{L}^{0} are “outward pointing” and hence do not contribute to changes in topology; see for instance [Zac17] or [PnZ18] for a discussion. For an interior critical point qjαVα0q^{\alpha}_{j}\in V^{\alpha}\cap\mathcal{L}^{0} we have the product of the tangential and the normal Morse data

(Dd(α)1,Dd(α)1)×(C((X,Vα)),(X,Vα))\left(D^{d(\alpha)-1},\partial D^{d(\alpha)-1}\right)\times\left(C(\mathcal{L}(X,V^{\alpha})),\mathcal{L}(X,V^{\alpha})\right)

where Dd(α)D^{d(\alpha)} is the disc of real dimension d(α)=dimVαd(\alpha)=\dim_{\mathbb{C}}V^{\alpha} and C((X,Vα))C(\mathcal{L}(X,V^{\alpha})) the cone over the complex link of XX along VαV^{\alpha}. It is a straighforward calculation that the Euler characteristic changes precisely by (1)d(α)1(1χ((X,Vα)))(-1)^{d(\alpha)-1}(1-\chi(\mathcal{L}(X,V^{\alpha}))) for the attachement of this cell at any of the critical points qjαq_{j}^{\alpha}. Summation over all these points on all relevant strata therefore gives us back the Formula (10) by Lê and Teissier.

It is evident that the above procedure can be applied inductively, cf. [TT81, Remarque 6.1.10]. This allows to reconstruct the codimension ii complex link i\mathcal{L}^{i} of (X,0)(X,0) from its hyperplane sections

ii+1d1={x1,,xm0(X,0)}\mathcal{L}^{i}\supset\mathcal{L}^{i+1}\supset\dots\supset\mathcal{L}^{d-1}=\{x_{1},\dots,x_{m_{0}(X,0)}\}

starting with d1\mathcal{L}^{d-1} which is just a set of points whose number is equal to the multiplicity m0(X,0)m_{0}(X,0) of (X,0)(X,0) at the origin. We leave it to the reader to verify the formula

(11) χ(i)=αA(j=i+1d(α)(1)d(α)jm0(Pd(α)j(Xα,0)))(1χ((X,Vα)).\chi(\mathcal{L}^{i})=\sum_{\alpha\in A}\left(\sum_{j=i+1}^{d(\alpha)}(-1)^{d(\alpha)-j}m_{0}\left(P_{d(\alpha)-j}(X^{\alpha},0)\right)\right)\cdot\left(1-\chi(\mathcal{L}(X,V^{\alpha})\right).

The coefficients appearing in this formula are nothing but the local Euler obstruction of (XαDi1,0)(X^{\alpha}\cap D_{i-1},0) at the origin:

(12) Eu(XαDi1,0)=j=id(α)(1)d(α)jm0(Pd(α)j(Xα,0))\mathrm{Eu}(X^{\alpha}\cap D_{i-1},0)=\sum_{j=i}^{d(\alpha)}(-1)^{d(\alpha)-j}m_{0}\left(P_{d(\alpha)-j}(X^{\alpha},0)\right)

cf. [TT81, Corollaire 5.1.2].

2.4. Polar multiplicities of generic determinantal varieties

We now turn towards the study of the generic determinantal varieties Mm,nsm×nM_{m,n}^{s}\subset\mathbb{C}^{m\times n}. These are equipped with the rank stratification, i.e. the decomposition

Mm,ns=r<sVm,nr,Vm,nr={φm×n:rankφ=r}.M_{m,n}^{s}=\bigcup_{r<s}V_{m,n}^{r},\quad V_{m,n}^{r}=\{\varphi\in\mathbb{C}^{m\times n}:\operatorname{rank}\varphi=r\}.

Due to its local analytic triviality, this stratification is easily seen to satisfy both Whitney’s conditions (a) and (b).

The reduced Euler characteristics of the classical complex links have been computed by Ebeling and Gusein-Zade in [EGZ09, Proposition 3]:

(13) 1χ((Mm,ns,0))=(1)s1(m1s1),1-\chi(\mathcal{L}(M_{m,n}^{s},0))=(-1)^{s-1}{m-1\choose s-1},

where, without loss of generality, it is assumed that mnm\leq n. The generic determinantal varieties admit a recursive pattern in the following sense. For r<smnr<s\leq m\leq n, a normal slice to the stratum Vm,nrMm,nsV_{m,n}^{r}\subset M_{m,n}^{s} through the point 𝟏m,nr\mathbf{1}_{m,n}^{r} is given by the set of matrices of the form

Nm,nr=𝟏r(mr)×(nr)=(𝟏r00(mr)×(nr))m×n.N_{m,n}^{r}=\mathbf{1}^{r}\oplus\mathbb{C}^{(m-r)\times(n-r)}=\begin{pmatrix}\mathbf{1}^{r}&0\\ 0&\mathbb{C}^{(m-r)\times(n-r)}\end{pmatrix}\subset\mathbb{C}^{m\times n}.

It follows immediately that (Mm,ns,Vm,nr)(Mmr,nrsr,0)\mathcal{L}(M_{m,n}^{s},V_{m,n}^{r})\cong\mathcal{L}(M_{m-r,n-r}^{s-r},0) and hence

(14) 1χ((Mm,ns,Vm,nr))=(1)sr1(mr1sr1).1-\chi(\mathcal{L}(M_{m,n}^{s},V_{m,n}^{r}))=(-1)^{s-r-1}{m-r-1\choose s-r-1}.

In order to determine the topological Euler characteristic of the complex links of higher codimension i(Mm,ns,0)\mathcal{L}^{i}(M_{m,n}^{s},0) of the generic determinantal varieties by means of the previous section, we need to know all the relevant polar multiplicities

(15) em,nr,k:=mk(Mm,nr+1,0)=m0(Pk(Mm,nr+1,0)).e_{m,n}^{r,k}:=m_{k}(M_{m,n}^{r+1},0)=m_{0}\left(P_{k}(M_{m,n}^{r+1},0)\right).

There are several methods to achieve this. For instance, one could simply choose random linear forms lil_{i} and compute the resulting multiplicity with the aid of a computer algebra system using e.g. Serre’s intersection formula. However, this approach provides very little insight and the results a priori depend on the choice of the linear forms. Recently, X. Zhang has computed the polar multiplicities in [Zha17] using Chern class calculus which is an exact computation not depending on any particular choices. His formulas, however, are very complicated since they appear as byproducts of the study of the Chern-Mather classes of determinantal varieties. In this section we will follow a more direct approach to the computation of the polar multiplicities using Chern classes.

In [TT81, Théorème 5.1.1], Lê and Teissier give the following formula for the polar multiplicities of a germ (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0):

(16) m0(Pk(𝒟))(1)d1𝔜ck(T~)c1(𝒪(1))dk1.m_{0}\left(P_{k}(\mathcal{D})\right)(-1)^{d-1}\int_{\mathfrak{Y}}c_{k}(\tilde{T})\cdot c_{1}(\mathcal{O}(1))^{d-k-1}.

Here the integral is taken over the exceptional divisor 𝔜\mathfrak{Y} of the blowup 𝔛\mathfrak{X} of the Nash modification X~\tilde{X} along the pullback of the maximal ideal at the origin for (X,0)(n,0)(X,0)\subset(\mathbb{C}^{n},0) and 𝒪(1)\mathcal{O}(1) denotes the dual of the tautological bundle for that blowup. By construction, these spaces can be arranged in a commutative diagram

(17) 𝔜\textstyle{\mathfrak{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ν1({0})\textstyle{\nu^{-1}(\{0\})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π1({0})\textstyle{\pi^{-1}(\{0\})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔛\textstyle{\mathfrak{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X~\textstyle{\tilde{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ν\scriptstyle{\nu}Bl0X\textstyle{\mathrm{Bl}_{0}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}X.\textstyle{X.}

where Bl0X\mathrm{Bl}_{0}X denotes the usual blowup of XX at the origin. We will now describe this diagram for the particular case where (X,0)=(Mm,ns,0)(m×n,0)(X,0)=(M_{m,n}^{s},0)\subset(\mathbb{C}^{m\times n},0) is a generic determinantal variety and deduce our particular formula (21) from that.

The Nash blowup of (Mm,ns,0)(m×n,0)(M_{m,n}^{s},0)\subset(\mathbb{C}^{m\times n},0) has been studied by Ebeling and Gusein-Zade in [EGZ09] and we briefly review their discussion. Let r=s1r=s-1 be the rank of the matrices in the open stratum Vm,nrMm,nsV_{m,n}^{r}\subset M_{m,n}^{s}. The tangent space to Vm,nrV_{m,n}^{r} at a point φ\varphi is known to be

(18) TφVm,nr={ψHom(n,m):ψ(kerφ)imφ}.T_{\varphi}V_{m,n}^{r}=\{\psi\in\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C}^{m}):\psi(\ker\varphi)\subset\mathrm{im}\,\varphi\}.

This fact can be exploited to replace the Grassmannian for the Nash blowup of Mm,nsM_{m,n}^{s} by a product: Let Grass(r,m)\operatorname{Grass}(r,m) be the Grassmannian of rr-planes in m\mathbb{C}^{m} and Grass(r,n)\operatorname{Grass}(r,n) the Grassmannian of rr-planes in (n)(\mathbb{C}^{n})^{\vee}, the dual of n\mathbb{C}^{n}. Then the Gauss map factors through

γ^:Vm,nrGrass(r,n)×Grass(r,m),φ((kerφ),imφ),\hat{\gamma}:V_{m,n}^{r}\to\operatorname{Grass}(r,n)\times\operatorname{Grass}(r,m),\quad\varphi\mapsto\left((\ker\varphi)^{\perp},\mathrm{im}\,\varphi\right),

where by (kerφ)(\ker\varphi)^{\perp} we mean the linear forms in (n)(\mathbb{C}^{n})^{\vee} vanishing on kerMn\ker M\subset\mathbb{C}^{n}. We denote this double Grassmannian by G=Grass(r,n)×Grass(r,m)G=\operatorname{Grass}(r,n)\times\operatorname{Grass}(r,m). On GG we have the two tautological exact sequences 0S1i1𝒪nπ1Q100\to S_{1}\overset{i_{1}}{\longrightarrow}\mathcal{O}^{n}\overset{\pi_{1}}{\longrightarrow}Q_{1}\to 0 and 0S2i2𝒪mπ2Q200\to S_{2}\overset{i_{2}}{\longrightarrow}\mathcal{O}^{m}\overset{\pi_{2}}{\longrightarrow}Q_{2}\to 0 coming from the two factors with S2S_{2} corresponding to the images and Q1Q_{1}^{\vee} to the kernels of the matrices φ\varphi under the modified Gauss map γ^\hat{\gamma}. With this notation, the space of matrices

Hom(n,m)(n)m\operatorname{Hom}(\mathbb{C}^{n},\mathbb{C}^{m})\cong(\mathbb{C}^{n})^{\vee}\otimes\mathbb{C}^{m}

pulls back to the trivial bundle 𝒪n𝒪m\mathcal{O}^{n}\otimes\mathcal{O}^{m} from which we may project to the product

π=π1π2:𝒪n𝒪mQ1Q2.\pi=\pi_{1}\otimes\pi_{2}\colon\mathcal{O}^{n}\otimes\mathcal{O}^{m}\to Q_{1}\otimes Q_{2}.

Then the condition on ψ\psi in (18) becomes π(ψ)=0\pi(\psi)=0 and consequently the Nash bundle on M~m,ns\tilde{M}_{m,n}^{s} is given by

(19) T~=γ^(kerπ).\tilde{T}=\hat{\gamma}^{*}\left(\ker\pi\right).

The Nash transform M~m,ns\tilde{M}_{m,n}^{s} itself can easily be seen to be isomorphic to the total space of the vector bundle

(20) M~m,ns|Hom((S1),S2)|=|S1S2|\tilde{M}_{m,n}^{s}\cong\left|\operatorname{Hom}\left((S_{1})^{\vee},S_{2}\right)\right|=\left|S_{1}\otimes S_{2}\right|

on GG. In particular, M~m,ns\tilde{M}_{m,n}^{s} is smooth and the maximal ideal of the origin in m×n\mathbb{C}^{m\times n} pulls back to the ideal sheaf of the zero section in M~m,ns\tilde{M}_{m,n}^{s}. Thus the exceptional divisor 𝔜\mathfrak{Y} in (17), i.e. the domain of integration in (16), is nothing but the projectivized bundle

𝔜=M~m,ns=(S1S2).\mathfrak{Y}=\mathbb{P}\tilde{M}_{m,n}^{s}=\mathbb{P}\left(S_{1}\otimes S_{2}\right).
Proposition 2.5.

The kk-th polar multiplicity of (Mm,nr+1,0)(m×n,0)(M_{m,n}^{r+1},0)\subset(\mathbb{C}^{m\times n},0) is given by

(21) em,nr,k=(1)(m+n)rr21Gsk(Q1Q2)s(m+n)r2r2k(S1S2)e_{m,n}^{r,k}=(-1)^{(m+n)\cdot r-r^{2}-1}\int_{G}s_{k}\left(Q_{1}\otimes Q_{2}\right)s_{(m+n)r-2r^{2}-k}(S_{1}\otimes S_{2})

where G=Grass(r,n)×Grass(r,m)G=\operatorname{Grass}(r,n)\times\operatorname{Grass}(r,m), SiS_{i} and QiQ_{i} are the tautological sub- and quotient bundles coming from either one of the two factors, and sks_{k} denotes the kk-th Segre class.

Proof.

Starting from the Lê-Teissier formula (16) we substitute the different terms according to the above identifications for the generic determinantal varieties. From (19) we see that for every k0k\geq 0

ck(T~)=sk(Q1Q2)c_{k}(\tilde{T})=s_{k}(Q_{1}\otimes Q_{2})

is the kk-th Segre class of the complement Q1Q2Q_{1}\otimes Q_{2} of T~\tilde{T} in 𝒪n𝒪m\mathcal{O}^{n}\otimes\mathcal{O}^{m}. The integral in question becomes

(S1S2)sk(Q1Q2)c1(𝒪(1))(m+n)rr21k\int_{\mathbb{P}(S_{1}\otimes S_{2})}s_{k}(Q_{1}\otimes Q_{2})\cdot c_{1}(\mathcal{O}(1))^{(m+n)r-r^{2}-1-k}

and we may perform this integration in two steps with the first one being integration along the fibers of the projection (S1S2)G\mathbb{P}(S_{1}\otimes S_{2})\to G. Since 𝒪(1)\mathcal{O}(1) denotes the dual of the tautological bundle for the projectivization of the underlying vector bundle S1S2S_{1}\otimes S_{2}, the result now follows from the projection formula, cf. [Ful98, Chapter 3.1]. ∎

Formula (21) can be implemented in computer algebra systems such as Singular [Dec+19]. For instance, using the library schubert.lib, the computations can easily be carried out for m,n5m,n\leq 5 on a desktop computer. Series such as, for example, e7,84,ke_{7,8}^{4,k} are also still feasible, but take up to 9 minutes to finish.

We have computed several polar multiplicities of the generic determinantal varieties using this formula. The results are listed in Tables 37.

Remark 2.6.

The reader will also note a symmetry that first appears in Table 4: For 0rmn0\leq r\leq m\leq n the polar multiplicities satisfy

(22) em,nr,k=em,nmr,2(mr)rk.e_{m,n}^{r,k}=e_{m,n}^{m-r,2(m-r)r-k}.

This phenomenon is based on a duality of the (projectivized) conormal modifications of the generic determinantal varieties, as has been explained to the author by Terence Gaffney in an oral communication. Formula (22) then follows from [Ura81, Theorem 3.3]. Interestingly, we have not succeeded to derive this symmetry from the formula in (21), but we nevertheless use it in the following tables in order to not duplicate the statements.

Remark 2.7.

The computation of Chern- and Segre classes of tensor products of vector bundles is a surprisingly expensive task from a computational point of view. Different formulas and algorithms have, for instance, been implemented in the Singular libraries chern.lib; see also [Szi19] for a further discussion.

The kk-th Segre class of the tensor product S1S2S_{1}\otimes S_{2} of the two tautological subbundles on G=Grass(r,n)×Grass(r,m)G=\operatorname{Grass}(r,n)\times\operatorname{Grass}(r,m) is the restriction of a universal polynomial

Pk[c1(S1),,cr(S1),c1(S2),,cr(S2)]P_{k}\in\mathbb{Z}[c_{1}(S_{1}),\dots,c_{r}(S_{1}),c_{1}(S_{2}),\dots,c_{r}(S_{2})]

in the Chern classes of the tautological bundles on the product of infinite Grassmannians m,nGrass(r,n)×Grass(r,m)\bigcup_{m,n\in\mathbb{N}}\operatorname{Grass}(r,n)\times\operatorname{Grass}(r,m). However, these polynomials are not sparse and their degree in the Chern roots is bounded only by r2r^{2}. Given that we have 2r2r variables, the number of coefficients of the polynomials PkP_{k} can roughly be estimated by (r2+2r12r1){r^{2}+2r-1\choose 2r-1}. Already for values for r10r\geq 10 we can therefore expect to have flooded the full RAM of any modern desktop computer.

Compared to that, the explicit model for the cohomology of Grass(r,n)\operatorname{Grass}(r,n) introduced above leads to an algebra of dimension (mr)(nr){m\choose r}{n\choose r} for the cohomology of GG. This number is in general strictly smaller than the number of coefficients of PkP_{k}. Since we shall only need the results for fixed values of mm and nn, it seems likely that a manual implementation of a modular approach, using the ideals Jm,rJ_{m,r} introduced above, could produce some further results for the polar multiplicities em,nr,ke_{m,n}^{r,k} which can not be reached using the methods provided by schubert.lib and chern.lib.

Other than that, it would, of course, be even more appealing to find a closed formula for the polar multiplicities as a function of m,n,rm,n,r, and kk.

k:k: 0 11 22
2×22\times 2 2 2 2
2×32\times 3 3 4 3
2×42\times 4 4 6 4
2×52\times 5 5 8 5
2×62\times 6 6 10 6
2×72\times 7 7 12 7
Table 3. The Polar multiplicities e2,n1,ke_{2,n}^{1,k} of 2×n2\times n-matrices for n7n\leq 7; values for kk which are not explicitly listed, are zero.
e3,n1,ke_{3,n}^{1,k} e3,n2,ke_{3,n}^{2,k}
k:k: 0 11 22 33 44 0 11 22 33 44
3×33\times 3 6 12 12 6 3 3 6 12 12 3
3×43\times 4 10 24 27 16 6 6 16 27 24 10
3×53\times 5 15 40 48 30 10 10 30 48 40 15
3×63\times 6 21 60 75 48 15 15 48 75 60 21
3×73\times 7 28 84 108 70 21 21 70 108 84 28
3×83\times 8 36 112 147 96 28 28 96 147 112 36
3×93\times 9 45 144 192 126 36 36 126 192 144 45
3×103\times 10 55 180 243 160 45 45 160 243 180 55
3×113\times 11 66 220 300 198 55 55 198 300 220 66
3×123\times 12 78 264 363 240 66 66 240 363 264 78
3×133\times 13 91 312 432 286 78 78 286 432 312 91
3×143\times 14 105 364 507 336 91 91 336 507 364 105
3×153\times 15 120 420 588 390 105 105 390 588 420 120
3×163\times 16 136 480 675 448 120 120 448 675 480 136
3×173\times 17 153 544 768 510 136 136 510 768 554 153
3×183\times 18 171 612 867 576 153 153 576 876 612 171
3×193\times 19 190 684 972 646 171 171 646 972 684 190
3×203\times 20 210 760 1083 720 190 190 720 1038 760 210
Table 4. Polar multiplicities for 3×n3\times n-matrices for n20n\leq 20; all values for kk which are not explicitly listed are zero.
e4,n2,ke_{4,n}^{2,k} with kk running from the left to the right e4,n1,ke_{4,n}^{1,k} with kk running from the left to the right
k:k: 0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6
4×44\times 4 20 80 176 256 286 256 176 80 20 20 60 84 68 36 12 4
4×54\times 5 50 240 595 960 1116 960 595 240 50 35 120 190 176 105 40 10
4×64\times 6 105 560 1488 2520 2980 2520 1488 560 105 56 210 360 360 228 90 20
4×74\times 7 196 1120 3115 5432 6488 5432 3115 1120 196 84 336 609 640 420 168 35
4×84\times 8 336 2016 5792 10304 12390 10304 5792 2016 336 120 504 952 1036 696 280 56
4×94\times 9 540 3360 9891 17856 21576 17856 9891 3360 540 165 720 1404 1568 1071 432 84
4×104\times 10 825 5280 15840 28920 35076 28920 15840 5280 825 220 990 1980 2256 1560 630 120
4×114\times 11 1210 7920 24123 44440 54060 44440 24123 7920 1210 286 1320 2695 3120 2178 880 165
4×124\times 12 1716 11440 35280 65472 79838 65472 35280 11440 1716 364 1716 3564 4180 2940 1188 220
4×134\times 13 - - - - - - - - - 455 2184 4602 5456 3861 1560 286
4×144\times 14 - - - - - - - - - 560 2730 5824 6968 4956 2002 364
4×154\times 15 - - - - - - - - - 680 3360 7245 8736 6240 2520 455
4×164\times 16 - - - - - - - - - 816 4080 8880 10780 7728 3120 560
4×174\times 17 - - - - - - - - - 969 4896 10744 13120 9435 3808 680
4×184\times 18 - - - - - - - - - 1140 5814 12852 15776 11376 4590 816
4×194\times 19 - - - - - - - - - 1330 6840 15219 18768 13566 5472 969
4×204\times 20 - - - - - - - - - 1540 7980 17860 22116 16020 6460 1140
4×214\times 21 - - - - - - - - - 1771 9240 20790 25840 18753 7560 1330
4×224\times 22 - - - - - - - - - 2024 10626 24024 29960 21780 8778 1540
4×234\times 23 - - - - - - - - - 2300 12144 27577 34496 25116 10120 1771
4×244\times 24 - - - - - - - - - 2600 13800 31464 39468 28776 11592 2024
l:l: 6 5 4 3 2 1 0
e4,n3,le_{4,n}^{3,l} with ll running from the right to the left
Table 5. Polar multiplicities for 4×n4\times n-matrices. A - indicates that this value has not been computed; all other entries for kk and ll which are not explicitly listed, are equal to zero.
e5,n2,ke_{5,n}^{2,k} with kk running from the left to the right
k:k: 0 1 2 3 4 5 6 7 8 9 10 11 12
5×55\times 5 175 1050 3180 6320 9180 10320 9360 7080 4545 2430 1020 300 50
5×65\times 6 490 3360 11445 25396 40890 50520 49495 39120 24981 12640 4830 1260 175
5×75\times 7 1176 8820 32480 77280 132300 172074 175080 141120 89880 44310 16128 3920 490
5×85\times 8 2520 20160 78498 196080 349860 470400 489930 399504 253980 123200 43470 10080 1176
5×95\times 9 4950 41580 168840 437220 803916 1106640 1171360 962640 611100 293076 101160 22680 2520
5×105\times 10 9075 79200 332310 884840 1664685 2332440 2498535 2064960 1309290 622560 211365 46200 4950
5×115\times 11 15730 141570 609840 1660296 3180705 4518690 4885440 4055040 2568456 1213080 406560 87120 9075
5×125\times 12 26026 240240 1057485 2931760 5699760 8188224 8918470 7427640 4700460 2207920 732303 154440 15730
l:l: 12 11 10 9 8 7 6 5 4 3 2 1 0
e5,n3,le_{5,n}^{3,l} with ll running from the right to the left
Table 6. Polar multiplicities for 5×n5\times n-matrices of rank 22 and 33; all entries for kk and ll which are not explicitly listed, are equal to zero.
e5,n1,ke_{5,n}^{1,k} with kk running from the left to the right
k:k: 0 1 2 3 4 5 6 7 8
5×55\times 5 70 280 520 580 430 220 80 20 5
5×65\times 6 126 560 1155 1440 1200 696 285 80 15
5×75\times 7 210 1008 2240 3010 2700 1680 728 210 35
5×85\times 8 330 1680 3948 5600 5285 3440 1540 448 70
5×95\times 9 495 2640 6480 9576 9380 6300 2880 840 126
5×105\times 10 715 3960 10065 15360 15480 10640 4935 1440 210
5×115\times 11 1001 5720 14960 23430 24150 16896 7920 2310 330
5×125\times 12 1365 8008 21450 34320 36025 25560 12078 3520 495
5×135\times 13 1820 10920 29848 48620 51810 37180 17680 5148 715
5×145\times 14 2380 14560 40495 66976 72280 52360 25025 7280 1001
5×155\times 15 3060 19040 53760 90090 98280 71760 34440 10010 1365
5×165\times 16 3876 24480 70040 118720 130725 96096 46280 13440 1820
5×175\times 17 4845 31008 89760 153680 170600 126140 60928 17680 2380
5×185\times 18 5985 38760 113373 195840 218960 162720 78795 22848 3060
5×195\times 19 7315 47880 141360 246126 276930 206720 100320 29070 3876
5×205\times 20 8855 58520 174230 305520 345705 259080 125970 36480 4845
l:l: 8 7 6 5 4 3 2 1 0
e5,n4,le_{5,n}^{4,l} with ll running from the right to the left
Table 7. Polar multiplicities for 5×n5\times n-matrices of rank 11 and 44; all entries for kk and ll which are not explicitly listed, are equal to zero.
Example 2.8.

We may use the above tables together with formula (11) to compute the Euler characteristics of complex links of higher codimension for the generic determinantal varieties. For instance

χ(6(M3,43,0))=e3,42,0e3,42,1+e3,42,2e3,42,3=616+2724=7.\chi\left(\mathcal{L}^{6}(M_{3,4}^{3},0)\right)=e_{3,4}^{2,0}-e_{3,4}^{2,1}+e_{3,4}^{2,2}-e_{3,4}^{2,3}=6-16+27-24=-7.

Note that since 6(M3,43,0)\mathcal{L}^{6}(M_{3,4}^{3},0) is smooth of complex dimension 33, the summation over αA\alpha\in A degenerates and only the smooth stratum V3,42V_{3,4}^{2} is relevant. Moreover, the complex link of M3,43M_{3,4}^{3} along this stratum is empty, so that the factor (1χ((M3,43,V3,42)))(1-\chi(\mathcal{L}(M_{3,4}^{3},V_{3,4}^{2}))) simply reduces to 11.

This computation confirms the results in an earlier paper [FKZ15] where it was shown that the Betti numbers of 6(M3,43,0)\mathcal{L}^{6}(M_{3,4}^{3},0) are

(b0,b1,b2,b3)=(1,0,1,9).(b_{0},b_{1},b_{2},b_{3})=(1,0,1,9).

In [FKZ15], this was a very particular example. We will discuss in Section 4 how the distinct Betti numbers can be computed for all smooth complex links of generic determinantal varieties.

To also give an example for a singular complex link, consider 5(M3,43,0)\mathcal{L}^{5}(M_{3,4}^{3},0): This space is of complex dimension 44 and has isolated singularities which are themselves determinantal of the form (M2,32,0)(M_{2,3}^{2},0). If D6D^{\prime}_{6} is a plane of codimension 66 in 3×4\mathbb{C}^{3\times 4} in general position off the origin such that 5(M3,43,0)=D6M3,43\mathcal{L}^{5}(M_{3,4}^{3},0)=D^{\prime}_{6}\cap M_{3,4}^{3}, then these singular points are precisely the intersection points of D6D^{\prime}_{6} with M3,42M_{3,4}^{2} and their number is equal to the multiplicity e3,41,0=10e_{3,4}^{1,0}=10.

If we let ll be a further, sufficiently general linear form on 3×4\mathbb{C}^{3\times 4}, then the generic fiber of its restriction to M3,43D6M_{3,4}^{3}\cap D^{\prime}_{6} is the previous space 6(M3,43,0)\mathcal{L}^{6}(M_{3,4}^{3},0) whose topology we already know. According to Table (4), ll has 1010 classical Morse critical points on V3,42D6V_{3,4}^{2}\cap D^{\prime}_{6} and 1010 further stratified Morse critical points on V3,41D6=M3,42D6V_{3,4}^{1}\cap D^{\prime}_{6}=M_{3,4}^{2}\cap D^{\prime}_{6}.

For the first set of points, 1010 more cells of real dimension 44 are added which changes the Euler characteristic by (1)106101=10(-1)^{10-6}\cdot 10\cdot 1=10 in Formula (10) (resp. (11)).

The second set of critical points on the lower dimensional stratum V3,41V_{3,4}^{1} have a nontrivial complex link (M3,43,V3,41)0(M2,32,0)\mathcal{L}(M_{3,4}^{3},V_{3,4}^{1})\cong\mathcal{L}^{0}(M_{2,3}^{2},0) appearing in the normal Morse datum. This complex link is nothing but the Milnor fiber of the A0+A_{0}^{+}-singularity in [FKZ15]: Despite being a space of complex dimension 33, it is homotopy equivalent to a 22-sphere and its Betti numbers are (b0,b1,b2,b3)=(1,0,1,0)(b_{0},b_{1},b_{2},b_{3})=(1,0,1,0) in accordance with the Formula by Ebeling and Gusein-Zade (13). This means that we attach real 33-cells rather than 44-cells and the Euler characteristic changes by 10-10 rather than +10+10, as one might have expected. The overall outcome therefore is

χ(5(M3,43,0))=χ(6(M3,43,0))+10(1(10+1))α=1+10(10)α=2=7.\chi\left(\mathcal{L}^{5}(M_{3,4}^{3},0)\right)=\chi(\mathcal{L}^{6}(M_{3,4}^{3},0))+\underbrace{10\cdot(1-(1-0+1))}_{\alpha=1}+\underbrace{10\cdot(1-0)}_{\alpha=2}=-7.

It is interesting to see the cancellation of the two contributions to the Euler characteristic given that the equality of the two relevant multiplicities is not a coincidence, but due to the duality noted by Gaffney.

We shall see later on that the first four Betti numbers of the open stratum V3,42D6V_{3,4}^{2}\cap D^{\prime}_{6} of 5(M3,43,0)\mathcal{L}^{5}(M_{3,4}^{3},0) are

(b0,b1,b2,b3)=(1,0,1,0).(b_{0},b_{1},b_{2},b_{3})=(1,0,1,0).

It can be shown that the attachements of the 33-cells at the points of V3,41D6V_{3,4}^{1}\cap D^{\prime}_{6} glue their boundaries all to the very same generator of the second homology group. From the long exact sequence of the pair (X3,42D6,V3,42D6)(X_{3,4}^{2}\cap D^{\prime}_{6},V_{3,4}^{2}\cap D^{\prime}_{6}) and the previous computation of the Euler characteristic one can then deduce that the Betti numbers of 5(M3,43,0)\mathcal{L}^{5}(M_{3,4}^{3},0) must be

(b0,b1,b2,b3,b4)=(1,0,0,9,1).(b_{0},b_{1},b_{2},b_{3},b_{4})=(1,0,0,9,1).

In particular we see that the cells attached at the classical critical points of ll on the smooth stratum kill off all the cycles in the top homology group of 6(M3,43,0)\mathcal{L}^{6}(M_{3,4}^{3},0). Those coming from the stratified Morse critical points on the lower dimensional stratum survive and lead to new cycles. Details for the computation of the Betti numbers in this example will appear in a forthcoming note.

3. Determinantal strata as homogeneous spaces

Let GG be a Lie group and :G×XX*\colon G\times X\to X a smooth action on a manifold XX. Then for every point xXx\in X the orbit GxXG*x\subset X is a locally closed submanifold which is diffeomorphic to the quotient G/GxG/G_{x} of GG by the stabilizer GxG_{x} of xx. The next lemma shows that up to homotopy we can always find a compact model for this orbit by choosing an appropriate maximal compact subgroup of GG.

Lemma 3.1.

Let GG be a Lie group, GGG^{\prime}\subset G a closed subgroup, and UGU\subset G a maximal compact subgroup such that U=UGU^{\prime}=U\cap G^{\prime} is again a maximal compact subgroup of GG^{\prime}. Then the inclusion U/UG/GU/U^{\prime}\hookrightarrow G/G^{\prime} is a weak homotopy equivalence.

Proof.

The projection GG/GG\to G/G^{\prime} is a fiber bundle with fiber GG^{\prime} and the same holds for UU/UU\mapsto U/U^{\prime} with fiber UU^{\prime}. Hence, there is a commutative diagram of long exact sequences of homotopy groups

\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk(G)\textstyle{\pi_{k}(G^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk(G)\textstyle{\pi_{k}(G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk(G/G)\textstyle{\pi_{k}(G/G^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk1(G)\textstyle{\pi_{k-1}(G^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk(U)\textstyle{\pi_{k}(U^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk(U)\textstyle{\pi_{k}(U)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk(U/U)\textstyle{\pi_{k}(U/U^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk1(U)\textstyle{\pi_{k-1}(U^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}

and it is well known that for any Lie group the inclusion of its maximal compact subgroup is a homotopy equivalence. The assertion therefore follows from the five-lemma. ∎

3.1. The Lie group action on m×n\mathbb{C}^{m\times n}

We now turn to the discussion of the strata in the rank stratification as homogeneous spaces. Fix integers 0<mn0<m\leq n. The space m×n\mathbb{C}^{m\times n} of complex m×nm\times n-matrices has a natural left action by the complex Lie group

Gm,n:=GL(m;)×GL(n;)G_{m,n}:=\operatorname{GL}(m;\mathbb{C})\times\operatorname{GL}(n;\mathbb{C})

via multiplication:

:Gm,n×m×nm×n,((P,Q),A)(P,Q)A=PAQ1.*\colon G_{m,n}\times\mathbb{C}^{m\times n}\to\mathbb{C}^{m\times n},\quad((P,Q),A)\mapsto(P,Q)*A=P\cdot A\cdot Q^{-1}.

For two matrices Am×nA\in\mathbb{C}^{m\times n} and Bm×nB\in\mathbb{C}^{m^{\prime}\times n^{\prime}} we will denote by ABA\oplus B the (m+m)×(n+n)(m+m^{\prime})\times(n+n^{\prime}) block matrix

(A00B).\begin{pmatrix}A&0\\ 0&B\end{pmatrix}.

Let 0m,nm×n0_{m,n}\in\mathbb{C}^{m\times n} be the zero matrix. For any number 0rm0\leq r\leq m we will write 𝟏m,nr=𝟏r0mr,nr\mathbf{1}_{m,n}^{r}=\mathbf{1}^{r}\oplus 0_{m-r,n-r} for the m×nm\times n-matrix with a unit matrix 𝟏r\mathbf{1}^{r} of rank rr in the upper left corner and zeroes in all other entries. Then clearly

Vm,nr=G𝟏m,nrGm,n/Gm,nr,V_{m,n}^{r}=G*\mathbf{1}_{m,n}^{r}\cong G_{m,n}/G_{m,n}^{r},

where Gm,nrG_{m,n}^{r} is the stabilizer of 𝟏m,nr\mathbf{1}_{m,n}^{r} in GG. A direct computation yields that Gm,nrG_{m,n}^{r} consists of pairs of block matrices of the form

((AB0C),(A0DE))\left(\begin{pmatrix}A&B\\ 0&C\end{pmatrix},\begin{pmatrix}A&0\\ D&E\end{pmatrix}\right)

with AGL(r,)A\in\operatorname{GL}(r,\mathbb{C}), CGL(mr;)C\in\operatorname{GL}(m-r;\mathbb{C}) and EGL(nr;)E\in\operatorname{GL}(n-r;\mathbb{C}) invertible, and BB and DD arbitrary of appropriate sizes.

As a compact subgroup Um,nGm,nU_{m,n}\subset G_{m,n} we may choose the unitary matrices U(m)×U(n)U(m)\times U(n). It is easily verified that its intersection Um,nrU_{m,n}^{r} with the subgroup Gm,nrG_{m,n}^{r} consists of pairs of matrices

((S00P),(S00Q))=(SP,SQ)\left(\begin{pmatrix}S&0\\ 0&P\end{pmatrix},\begin{pmatrix}S&0\\ 0&Q\end{pmatrix}\right)=\left(S\oplus P,S\oplus Q\right)

with SU(r)S\in U(r), PU(mr)P\in U(m-r), and QU(nr)Q\in U(n-r), and that this is in fact a maximal compact subgroup of Gm,nrG_{m,n}^{r}. Note that due to the fact that in contrast to Gm,nrG_{m,n}^{r} the off-diagonal blocks in the subgroup Um,nrU_{m,n}^{r} are all zero, we find that

(23) Um,nrU(r)×U(mr)×U(nr)U_{m,n}^{r}\cong U(r)\times U(m-r)\times U(n-r)

is again isomorphic to a product of unitary groups.

Let us, for the moment, consider only the first factor U(m)U(m) of Um,nU_{m,n} which we may consider as a subgroup via the inclusion U(m)×{𝟏r}Um,nU(m)\times\{\mathbf{1}^{r}\}\subset U_{m,n}. The stabilizer of 𝟏m,nr\mathbf{1}_{m,n}^{r} of the restriction of the action to U(m)U(m) is simply the subgroup 𝟏rU(mr)\mathbf{1}^{r}\oplus U(m-r). The U(m)U(m)-orbit can easily be identified with the Stiefel manifold Stief(r,m)\operatorname{Stief}(r,m) of orthonormal rr-frames in m\mathbb{C}^{m}:

(24) Stief(r,m)U(m)/(𝟏rU(mr))U(m)𝟏m,nr\operatorname{Stief}(r,m)\cong U(m)/\left(\mathbf{1}^{r}\oplus U(m-r)\right)\cong U(m)*\mathbf{1}_{m,n}^{r}

so that an rr-frame v¯=(v1,,vr)\underline{v}=(v_{1},\dots,v_{r}) in m\mathbb{C}^{m} is given by the first rr columns of the matrix (v¯)=A𝟏m,nr(\underline{v})=A\cdot\mathbf{1}_{m,n}^{r} for some AU(m)A\in U(m). The group U(r)U(r) operates naturally on the Stiefel manifold via the left action

U(r)×Stief(r,m)Stief(r,m),(S,A𝟏m,rr)A𝟏m,rrS1.U(r)\times\operatorname{Stief}(r,m)\to\operatorname{Stief}(r,m),\quad(S,A\cdot\mathbf{1}^{r}_{m,r})\mapsto A\cdot\mathbf{1}^{r}_{m,r}\cdot S^{-1}.

The quotient of this action is the Grassmannian of rr-planes Grass(r,m)\operatorname{Grass}(r,m) since either two rr-frames span the same subspace if and only if they lay in the same orbit under this U(r)U(r)-action.

It is easy to see with the above identifications (24) that two matrices AA and AA^{\prime} in U(m)U(m) represent the same element in Grass(r,m)\operatorname{Grass}(r,m) if and only if A1AU(r)U(mr)A^{-1}\cdot A^{\prime}\in U(r)\oplus U(m-r):

A𝟏m,rr\displaystyle A\cdot\mathbf{1}_{m,r}^{r} =\displaystyle= A𝟏m,rrS1\displaystyle A^{\prime}\cdot\mathbf{1}_{m,r}^{r}\cdot S^{-1}
𝟏m,rr\displaystyle\Leftrightarrow\qquad\mathbf{1}_{m,r}^{r} =\displaystyle= A1A(S1𝟏mr)𝟏m,rr\displaystyle A^{-1}\cdot A^{\prime}\cdot(S^{-1}\oplus\mathbf{1}^{m-r})\cdot\mathbf{1}_{m,r}^{r}
(𝟏rP)\displaystyle\Leftrightarrow\quad(\mathbf{1}^{r}\oplus P) =\displaystyle= A1A(S1𝟏mr)\displaystyle A^{-1}\cdot A^{\prime}\cdot(S^{-1}\oplus\mathbf{1}^{m-r})
(SP)\displaystyle\Leftrightarrow\quad(S\oplus P) =\displaystyle= A1A\displaystyle A^{-1}\cdot A^{\prime}

for some PU(mr)P\in U(m-r). In other words, the above U(r)U(r) action is compatible with the natural inclusion of subgroups

U(r)𝟏mrU(r)U(mr)U(m)U(r)\oplus\mathbf{1}^{m-r}\hookrightarrow U(r)\oplus U(m-r)\hookrightarrow U(m)

and accordingly

(25) Grass(r,m)Stief(r,m)/U(r)U(m)/U(r)U(mr).\operatorname{Grass}(r,m)\cong\operatorname{Stief}(r,m)/U(r)\cong U(m)/U(r)\oplus U(m-r).

We can repeat these considerations for the second factor U(n)U(n) embedded into Um,nU_{m,n} as {𝟏m}×U(n)\{\mathbf{1}^{m}\}\times U(n). Then

Stief(r,n)U(n)/(𝟏rU(nr))U(n)𝟏m,nr\operatorname{Stief}(r,n)\cong U(n)/(\mathbf{1}^{r}\oplus U(n-r))\cong U(n)*\mathbf{1}_{m,n}^{r}

with any rr-frame w¯Stief(r,n)\underline{w}\in\operatorname{Stief}(r,n) given by the first rr rows of the matrix (w¯)=𝟏m,nrB1(\underline{w})=\mathbf{1}^{r}_{m,n}\cdot B^{-1} for some BU(n)B\in U(n). Accordingly, we will write the left action by U(r)U(r) on Stief(r,n)\operatorname{Stief}(r,n) as

U(r)×Stief(r,n)Stief(r,n),(S,𝟏m,nrB1)S𝟏m,nrB1U(r)\times\operatorname{Stief}(r,n)\to\operatorname{Stief}(r,n),\quad(S,\mathbf{1}_{m,n}^{r}\cdot B^{-1})\mapsto S\cdot\mathbf{1}_{m,n}^{r}\cdot B^{-1}

in this case.

Note that, on the one hand, the subgroup Um,nrU_{m,n}^{r} intersects the subgroups U(m)×{𝟏n}U(m)\times\{\mathbf{1}^{n}\} and {𝟏m}×U(n)\{\mathbf{1}^{m}\}\times U(n) in 𝟏rU(mr)\mathbf{1}^{r}\oplus U(m-r) and 𝟏rU(nr)\mathbf{1}^{r}\oplus U(n-r), respectively, and the action of the latter subgroups affects only either one of the two factors. The U(r)U(r)-action, on the other hand, is “diagonal” and we may exploit these facts by observing that the quotient

(26) Um,n/U(mr)×U(nr)Stief(r,m)×Stief(r,n)U_{m,n}/U(m-r)\times U(n-r)\cong\operatorname{Stief}(r,m)\times\operatorname{Stief}(r,n)

is a product of Stiefel manifolds, equipped with a free, diagonal U(r)U(r)-action. The quotient Stief(r,m)×Stief(r,n)/U(r)\operatorname{Stief}(r,m)\times\operatorname{Stief}(r,n)/U(r) is then naturally isomorphic to Um,n/(U(r)×U(mr)×U(nr))U_{m,n}/\left(U(r)\times U(m-r)\times U(n-r)\right) and, via the particular choice of the matrix 𝟏m,nr\mathbf{1}_{m,n}^{r}, this manifold can be identified with the orbit Um,n𝟏m,nrU_{m,n}*\mathbf{1}_{m,n}^{r}. We will in the following denote this orbit by Om,nrVm,nrm×nO_{m,n}^{r}\subset V_{m,n}^{r}\subset\mathbb{C}^{m\times n} and refer to it as the compact orbit model for the stratum Vm,nrV_{m,n}^{r}.

Lemma 3.2.

The two natural projections

Stief(r,n)\textstyle{\operatorname{Stief}(r,n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Stief(r,m)\textstyle{\operatorname{Stief}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Om,nr\textstyle{O_{m,n}^{r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ2\scriptstyle{\lambda_{2}}λ1\scriptstyle{\lambda_{1}}Grass(r,n)\textstyle{\operatorname{Grass}(r,n)}Grass(r,m)\textstyle{\operatorname{Grass}(r,m)}

equip the space Um,n/Um,nrUm,n𝟏m,nrU_{m,n}/U_{m,n}^{r}\cong U_{m,n}*\mathbf{1}_{m,n}^{r} with two structures as a fiber bundle over the respective Grassmannian with Stiefel manifolds as fibers.

Proof.

It suffices to establish this claim for the first projection λ1\lambda_{1} to Grass(r,m)\operatorname{Grass}(r,m). Consider the commutative diagram

Stief(r,m)×Stief(r,n)\textstyle{\operatorname{Stief}(r,m)\times\operatorname{Stief}(r,n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ1\scriptstyle{\lambda^{\prime}_{1}}α\scriptstyle{\alpha}Om,nr\textstyle{O_{m,n}^{r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ1\scriptstyle{\lambda_{1}}Stief(r,m)\textstyle{\operatorname{Stief}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}Grass(r,m)\textstyle{\operatorname{Grass}(r,m)}

where λ1\lambda^{\prime}_{1} takes a pair of rr-frames (v¯,w¯)(\underline{v},\underline{w}) to v¯\underline{v}, β\beta is the quotient map v¯spanv¯\underline{v}\mapsto\operatorname{span}\underline{v} from (25) and α\alpha the one from the discussion of (26). We need to describe the fiber of an arbitrary point WGrass(r,m)W\in\operatorname{Grass}(r,m). To this end, consider its preimages

(βλ1)1({W})=β1({W})×Stief(r,n)λ1β1({W})=U(r)v¯(\beta\circ\lambda^{\prime}_{1})^{-1}(\{W\})=\beta^{-1}(\{W\})\times\operatorname{Stief}(r,n)\overset{\lambda^{\prime}_{1}}{\longrightarrow}\beta^{-1}(\{W\})=U(r)*\underline{v}

with v¯=(v1,,vr)Stief(r,m)\underline{v}=(v_{1},\dots,v_{r})\in\operatorname{Stief}(r,m) some rr-frame in m\mathbb{C}^{m} with spanv¯=W\operatorname{span}\underline{v}=W. The fiber of λ1\lambda^{\prime}_{1} over v¯\underline{v} is simply the Stiefel manifold Stief(r,n)\operatorname{Stief}(r,n). Now if v¯\underline{v}^{\prime} is any other rr-frame spanning WW, then there exists a unique matrix SU(r)S\in U(r) such that v¯=v¯S1\underline{v}^{\prime}=\underline{v}\cdot S^{-1}. The free diagonal action on Stief(r,m)×Stief(r,n)\operatorname{Stief}(r,m)\times\operatorname{Stief}(r,n) gives a natural identification of the fibers of λ1\lambda^{\prime}_{1} over v¯\underline{v} and v¯\underline{v}^{\prime} via

(v¯,w¯)(v¯,Sw¯)(\underline{v},\underline{w})\mapsto(\underline{v}^{\prime},S\cdot\underline{w})

and this furnishes an obvious notion of parallel sections of λ1\lambda^{\prime}_{1} over the orbit U(r)v¯U(r)*\underline{v}. These parallel sections can then be identified with either one of the Stiefel manifolds Stief(r,n)×{v¯}\operatorname{Stief}(r,n)\times\{\underline{v}\} over a point v¯\underline{v} in the orbit. ∎

Remark 3.3.

The structures of the manifolds Om,nrO_{m,n}^{r} as fiber bundle is in general not trivial. For instance

S2m1Om,11λ1Grass(1,m)m1S^{2m-1}\cong O_{m,1}^{1}\overset{\lambda_{1}}{\longrightarrow}\operatorname{Grass}(1,m)\cong\mathbb{P}^{m-1}

is the Hopf fibration which is known not to be a product.

3.2. The Cartan model

The cohomology of homogeneous spaces can be computed via the Cartan model as outlined by Borel in [Bor53, Théorème 25.2]. Let GG be a compact, connected Lie group and UGU\subset G a closed subgroup thereof. Then [Bor53, Théorème 25.2] allows for the description of the cohomology of the quotient G/UG/U as the cohomology of an explicit complex

(27) H(G/U)H(SUF)H^{\bullet}(G/U)\cong H\left(S_{U}\otimes_{\mathbb{Z}}\bigwedge F\right)

under certain favourable assumptions on GG and UU. The objects on the right hand side are the following.

  • The ring SUS_{U} is the cohomology ring of a classifying space for the group UU, see for instance [Hus94, Part I, Chapter 4.11]. Such a classifying space BUBU for a compact Lie group UU is given by the quotient of any weakly contractible space EUEU with a free UU-action. Then the projection EUBUEU\to BU turns EUEU into a universal bundle in the sense that every principal UU-bundle PP over a paracompact Haussdorff space XX can be written as P=fEUP=f^{*}EU for some continuous map f:XBUf\colon X\to BU.

    In particular, this property can be used to show that the cohomology ring SU=H(BU)S_{U}=H^{\bullet}(BU) is in fact unique up to unique isomorphism and independent of the choice of the space EUEU, see e.g. [Bor53, Section 18]222The approach by Borel might seem unnecessarily technical given the Milnor construction of universal bundles in [Mil56] three years later.. In most practical cases (cf. [Bor53, Théorème 19.1]) the cohomology rings of classifying spaces are weighted homogeneous polynomial rings SU[c1,,cr]S_{U}\cong\mathbb{Z}[c_{1},\dots,c_{r}] in variables of even degree and therefore in particular commutative.

    Furthermore, we note that the total space EGEG of a universal GG bundle is naturally equipped with a free UU action, as well. It can therefore also be used to construct a classifying space BUBU as the intermediate quotient

    EGBU=EG/UBG=EG/G.EG\to BU=EG/U\to BG=EG/G.

    The pullback in cohomology of the projection BUBGBU\to BG is called the characteristic homomorphism ρ:SGSU\rho\colon S_{G}\to S_{U} for the inclusion of the subgroup UGU\subset G, cf. [Bor53, Théorème 22.2].

  • The module FF is a free, graded \mathbb{Z}-module

    F=ε1εrF=\mathbb{Z}\varepsilon_{1}\oplus\dots\oplus\mathbb{Z}\varepsilon_{r}

    in generators εi\varepsilon_{i} of odd degree. Hopf has shown in [Hop41, Satz 1] that the rational homology of a compact Lie group UU is graded isomorphic to the homology of a product of odd-dimensional spheres. This result has been strengthened to also include cohomology with integer coefficients in the absence of torsion in H(U)H^{\bullet}(U), see e.g. [Bor53, Proposition 7.3]. Then the generators εi\varepsilon_{i} can be thought of as the volume forms of the spheres and the cup product turns the cohomology of the group into an exterior algebra on these generators

    H(U)(ε1εr)H^{\bullet}(U)\cong\bigwedge\left(\mathbb{Z}\varepsilon_{1}\oplus\dots\oplus\mathbb{Z}\varepsilon_{r}\right)

    which appears as F\bigwedge F in (27).

  • The differential DD on SUFS_{U}\otimes_{\mathbb{Z}}\bigwedge F is given by linear extension of the map

    D(a1)=0,D(1εi)=ρ(ci)1D(a\otimes 1)=0,\quad D(1\otimes\varepsilon_{i})=\rho(c_{i})\otimes 1

    for all aSUa\in S_{U}, where cic_{i} is a transgression element of εi\varepsilon_{i} in a universal GG-bundle and ρ:SGSU\rho\colon S_{G}\to S_{U} the characteristic homomorphism from before. A transgression can be defined in a universal UU-bundle π:EUBU\pi\colon EU\to BU as above: The element εi\varepsilon_{i} is called universally transgressive if there exists a cochain ωi\omega_{i} on EUEU which restricts to the cohomology class εi\varepsilon_{i} in every fiber and for which there exists another cochain aia_{i} on BGBG with πai=dωi\pi^{*}a_{i}=\operatorname{d}\!\omega_{i}. Then cic_{i} is taken to be the cohomology class of aia_{i} and it is said to “correspond to εi\varepsilon_{i} under transgression”. For a more detailed account see [Bor53]. We will discuss the particular form of this transgression below in the cases of interest for this article.

Note that (27) is an isomorphism of graded \mathbb{Z}-modules with grading given by the degree of the cohomology classes on either side. But we can also think of the algebra SUFS_{U}\otimes_{\mathbb{Z}}\bigwedge F as a Koszul algebra in the generators 1εi1\otimes_{\mathbb{Z}}\varepsilon_{i} over the ring SUS_{U}:

SUFp=0r(p(i=1rSUεi)).S_{U}\otimes_{\mathbb{Z}}\bigwedge F\cong\bigoplus_{p=0}^{r}\left(\bigwedge^{p}\left(\bigoplus_{i=1}^{r}S_{U}\otimes_{\mathbb{Z}}\varepsilon_{i}\right)\right).

This gives another grading on the right hand side of (27) by the degrees pp of the distinct exterior powers. In the following we will refer to the two gradings as the cohomological degree and the Koszul degree respectively.

3.3. Maximal tori and the Weyl group

As explained in [Bor53, Section 29], the characteristic homomorphism ρ:SGSU\rho\colon S_{G}\to S_{U} associated to an inclusion of a subgroup UGU\subset G is best understood in terms of inclusions of maximal tori SUS\subset U and TGT\subset G. This will be an essential ingredient for the computation of the cohomology of the orbits Om,nrO_{m,n}^{r} in the determinantal strata.

For the group U(1)S1U(1)\cong S^{1}\subset\mathbb{C} a classifying space BU(1)BU(1) is given by the infinite projective space

BU(1)k=1kBU(1)\cong\bigcup_{k=1}^{\infty}\mathbb{P}^{k}

which can be understood as a direct limit with k\mathbb{P}^{k} included in k+1\mathbb{P}^{k+1} as the hyperplane section at infinity. A universal U(1)U(1)-bundle is then given by 𝒪(1)\mathcal{O}(-1)^{*}, the tautological bundle with its zero section removed or, equivalently, by the direct limit of unit spheres S2k+1k+1{0}S^{2k+1}\subset\mathbb{C}^{k+1}\setminus\{0\} which are projected to k\mathbb{P}^{k} via the Hopf fibration. Then the cohomology ring SU(1)[α]S_{U(1)}\cong\mathbb{Z}[\alpha] of BU(1)BU(1) is a free polynomial ring in α\alpha, the first Chern class of 𝒪(1)\mathcal{O}(1), and the generator ε\varepsilon of H1(U(1))H^{1}(U(1)) corresponds to α\alpha under transgression.

Now it is easy to see that for a torus T=(U(1))rT=(U(1))^{r} the classifying spaces and universal bundles can be chosen to be merely products of the one just described for U(1)U(1). It follows that STS_{T} is a polynomial ring

ST[α1,,αr]S_{T}\cong\mathbb{Z}[\alpha_{1},\dots,\alpha_{r}]

with all αj\alpha_{j} of degree 22.

If TT is a maximal torus in a compact connected Lie group GG, then SGS_{G} is contained in STS_{T} as the invariant ring under the action of the Weyl group. Moreover, if UGU\subset G is a subgroup and the tori SUS\subset U and TGT\subset G have been chosen such that STS\subset T, then the characteristic homomorphism ρ=ρ(U,G)\rho=\rho(U,G) for the inclusion UGU\subset G is completely determined by the one for the inclusion STS\subset T so that one has a commutative diagram

(28) SG\textstyle{S_{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(U,G)\scriptstyle{\rho(U,G)}SU\textstyle{S_{U}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ST\textstyle{S_{T}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(S,T)\scriptstyle{\rho(S,T)}SS\textstyle{S_{S}}

where the vertical arrows denote the inclusions of invariant subgroups. We note that in general one needs real coefficients in cohomology as in [Bor53, Proposition 29.2]. For the particular cases that we shall need below, however, the calculations have been carried out for integer coefficients as well.

3.4. Cohomology of Stiefel manifolds and Grassmannians

As discussed earlier the Stiefel manifolds and Grassmannians can be considered as homogeneous spaces of U(n)U(n) modulo various subgroups of block matrices with unitary blocks. Also the orbit varieties Om,nrO_{m,n}^{r} can be decomposed into these building blocks. Therefore we briefly review the theory for the Lie group U(n)U(n) as it can be found in [Bor53] or [Hus94] and illustrate the formula (27) for the classical cases, thereby fixing notation for the description of the cohomology of the orbit models Om,nrO_{m,n}^{r} that we are really aiming for.

The cohomology of the unitary group U(n)U(n) is known to be

(29) H(U(n))(ε1ε2εn1εn)H^{\bullet}(U(n))\cong\bigwedge\left(\mathbb{Z}\varepsilon_{1}\oplus\mathbb{Z}\varepsilon_{2}\oplus\dots\mathbb{Z}\varepsilon_{n-1}\oplus\mathbb{Z}\varepsilon_{n}\right)

with generators εi\varepsilon_{i} of degree 2i12i-1, see for example [Bor53, Proposition 9.1] or [Hus94, Part I, Chapter 7]. We will write

Fn=[1][3][32n][12n]F_{n}=\mathbb{Z}[-1]\oplus\mathbb{Z}[-3]\oplus\dots\oplus\mathbb{Z}[3-2n]\oplus\mathbb{Z}[1-2n]

for the free graded module whose direct summands are shifted by 12i1-2i for 1in1\leq i\leq n so that with this notation H(U(n))FnH^{\bullet}(U(n))\cong\bigwedge F_{n}. A maximal torus in U(n)U(n) is given by the diagonal matrices

T\displaystyle T =\displaystyle= {diag(λ1,,λn):λiU(1)}\displaystyle\{\mathrm{diag}(\lambda_{1},\dots,\lambda_{n}):\lambda_{i}\in U(1)\}

and the Weyl group is the symmetric group 𝔖n\mathfrak{S}_{n} of permutations of nn elements in this case. Writing α1,,αn\alpha_{1},\dots,\alpha_{n} for the generators of the cohomology of STS_{T} as above we find that

SGSTS_{G}\hookrightarrow S_{T}

is the inclusion of the invariant subring SG=[α1,,αn]𝔖nS_{G}=\mathbb{Z}[\alpha_{1},\dots,\alpha_{n}]^{\mathfrak{S}_{n}}. According to the fundamental theorem of symmetric functions, SGS_{G} is itself a polynomial ring in the elementary symmetric functions

σkn(α1,,αn):=0<i1<<iknj=1nαij.\sigma_{k}^{n}(\alpha_{1},\dots,\alpha_{n}):=\sum_{0<i_{1}<\dots<i_{k}\leq n}\prod_{j=1}^{n}\alpha_{i_{j}}.

Moreover, the generators εk\varepsilon_{k} correspond to these σk\sigma_{k} under transgression in a universal bundle EU(n)BU(n)EU(n)\to BU(n), see for example [Bor53, Section 19].

The cohomology of Stiefel manifolds Stief(r,n)\operatorname{Stief}(r,n) turns out to be a truncated version of the cohomology of U(n)U(n):

(30) H(Stief(r,n))Fr[2r2n].H^{\bullet}(\operatorname{Stief}(r,n))\cong\bigwedge F_{r}[2r-2n].

In order to make the connection with Formula (27) recall that Stief(r,n)U(n)/(𝟏rU(nr))\operatorname{Stief}(r,n)\cong U(n)/(\mathbf{1}^{r}\oplus U(n-r)). As maximal tori in the subgroup U:=𝟏rU(nr)U^{\prime}:=\mathbf{1}^{r}\oplus U(n-r) we may choose

T\displaystyle T^{\prime} =\displaystyle= {𝟏rdiag(μ1,,μnr):μjU(1)}\displaystyle\{\mathbf{1}^{r}\oplus\mathrm{diag}(\mu_{1},\dots,\mu_{n-r}):\mu_{j}\in U(1)\}

so that TTT^{\prime}\subset T with TU(n)T\subset U(n) as before. Writing ST=[β1,,βnr]S_{T^{\prime}}=\mathbb{Z}[\beta_{1},\dots,\beta_{n-r}] for the cohomology of the classifying space we find that the diagram (28) becomes

[α1,,αn]𝔖n\textstyle{\mathbb{Z}[\alpha_{1},\dots,\alpha_{n}]^{\mathfrak{S}_{n}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(U,U)\scriptstyle{\rho(U^{\prime},U)}[β1,,βnr]𝔖nr\textstyle{\mathbb{Z}[\beta_{1},\dots,\beta_{n-r}]^{\mathfrak{S}_{n-r}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[α1,,αn]\textstyle{\mathbb{Z}[\alpha_{1},\dots,\alpha_{n}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(T,T)\scriptstyle{\rho(T^{\prime},T)}[β1,,βnr]\textstyle{\mathbb{Z}[\beta_{1},\dots,\beta_{n-r}]}

where the map ρ(T,T)\rho(T^{\prime},T) is given by

ρ(T,T):αj{βjr if j>r,0 otherwise. \rho(T^{\prime},T)\colon\alpha_{j}\mapsto\begin{cases}\beta_{j-r}&\textnormal{ if }j>r,\\ 0&\textnormal{ otherwise. }\end{cases}

It follows that ρ(U,U)\rho(U^{\prime},U) is merely a substitution of the variables in the symmetric functions given by ρ(T,T)\rho(T^{\prime},T) so that

ρ(U,U):σkn(α1,,αn){σknr(β1,,βnr) if knr,0 otherwise.\rho(U^{\prime},U)\colon\sigma_{k}^{n}(\alpha_{1},\dots,\alpha_{n})\mapsto\begin{cases}\sigma_{k}^{n-r}(\beta_{1},\dots,\beta_{n-r})&\textnormal{ if }k\leq n-r,\\ 0&\textnormal{ otherwise.}\end{cases}

With these considerations at hand we can now investigate the homology of the complex SUFnS_{U^{\prime}}\otimes_{\mathbb{Z}}\bigwedge F_{n}, with its transgression differential DD from (27). To this end we shall apply the following well known reduction lemma.

Lemma 3.4.

Let RR be a ring, MM an RR-module and x,y1,,ynRx,y_{1},\dots,y_{n}\in R be elements. When xx is a non-zerodivisor on MM, then there is a canonical isomorphism

H(Kosz(x,y1,,yn;M))H(Kosz(y1,,yn;M/xM)H(\operatorname{Kosz}(x,y_{1},\dots,y_{n};M))\cong H(\operatorname{Kosz}(y_{1},\dots,y_{n};M/xM)

for the entire cohomology of the Koszul complexes.

Proof.

See [BH93, Corollary 1.6.13 (b)]. ∎

The ring SU[c1,,cnr]S_{U^{\prime}}\cong\mathbb{Z}[c_{1},\dots,c_{n-r}] is freely generated in the elementary symmetric functions in the β1,,βnr\beta_{1},\dots,\beta_{n-r}. Now the transgression differential DD takes the form

D:1εi{ci for inr0 otherwise. D\colon 1\otimes\varepsilon_{i}\mapsto\begin{cases}c_{i}&\textnormal{ for }i\leq n-r\\ 0&\textnormal{ otherwise. }\end{cases}

Since the cic_{i} form a regular sequence on the module M=SUM=S_{U^{\prime}} with quotient SU/c1,,cnrS_{U^{\prime}}/\langle c_{1},\dots,c_{n-r}\rangle\cong\mathbb{Z}, it follows inductively from Lemma 3.4 that

H(Stief(r,n))H(SUFn)H(Kosz(0,,0;)Fr[2r2n]H(\operatorname{Stief}(r,n))\cong H\left(S_{U^{\prime}}\otimes_{\mathbb{Z}}\bigwedge F_{n}\right)\cong H\left(\operatorname{Kosz}(0,\dots,0;\mathbb{Z}\right)\cong\bigwedge F_{r}[2r-2n]

as anticipated.

Let 0rn0\leq r\leq n be integers. The case of a Grassmannian

Grass(r,n)U(n)/(U(r)U(nr))\operatorname{Grass}(r,n)\cong U(n)/(U(r)\oplus U(n-r))

is similar, only that the maximal tori for U(n)U(n) and its subgroup U=U(r)U(nr)U^{\prime}=U(r)\oplus U(n-r) can be chosen to be the same so that ST=ST=[α1,,αn]S_{T}=S_{T^{\prime}}=\mathbb{Z}[\alpha_{1},\dots,\alpha_{n}]. The difference comes from the Weyl groups. For U(n)U(n) we again have the full symmetric group 𝔖n\mathfrak{S}_{n}, but for the subgroup UU^{\prime} we find

𝔖r×𝔖nr𝔖n\mathfrak{S}_{r}\times\mathfrak{S}_{n-r}\subset\mathfrak{S}_{n}

whose action on [α1,,αn]\mathbb{Z}[\alpha_{1},\dots,\alpha_{n}] respects the partion of the variables αj\alpha_{j} into subsets {α1,,αr}\{\alpha_{1},\dots,\alpha_{r}\} and {αnr+1,,αn}\{\alpha_{n-r+1},\dots,\alpha_{n}\}. We write

xj=σjr(α1,,αr),yk=σknr(αnr+1,,αn)x_{j}=\sigma^{r}_{j}(\alpha_{1},\dots,\alpha_{r}),\quad y_{k}=\sigma^{n-r}_{k}(\alpha_{n-r+1},\dots,\alpha_{n})

for the elementary symmetric polynomials in the respective set of variables. Then SUS_{U^{\prime}} is a free polynomial subring in these variables

SU[xj,yk:0<jr, 0<knr][α1,,αn]S_{U^{\prime}}\cong\mathbb{Z}[x_{j},y_{k}:0<j\leq r,\,0<k\leq n-r]\subset\mathbb{Z}[\alpha_{1},\dots,\alpha_{n}]

containing SUS_{U} as the invariant subring under arbitrary permutations, i.e. forgetting about the particular partition. It is an elementary exercise in the theory of symmetric functions to verify that in this situation

(31) σdn(α¯)=j+k=dσjr(α1,,αr)σknr(αnr+1,,αn)=j+k=dxjyk.\sigma_{d}^{n}(\underline{\alpha})=\sum_{j+k=d}\sigma_{j}^{r}(\alpha_{1},\dots,\alpha_{r})\cdot\sigma_{k}^{n-r}(\alpha_{n-r+1},\dots,\alpha_{n})=\sum_{j+k=d}x_{j}\cdot y_{k}.

Now the complex SUFnS_{U^{\prime}}\otimes_{\mathbb{Z}}\bigwedge F_{n} in (27) looks as follows, cf. [Bor53, Proposition 31.1]: The differential DD takes each one of the generators εk\varepsilon_{k} to D(1εk)=σkn(α1,,αn)D(1\otimes\varepsilon_{k})=\sigma_{k}^{n}(\alpha_{1},\dots,\alpha_{n}). These are nn weighted homogeneous relations in a free graded polynomial ring with nn variables x1,,xr,y1,,ynrx_{1},\dots,x_{r},y_{1},\dots,y_{n-r}. From the fact that the cohomology of the associated Koszul complex is finite dimensional, we see that the elements (31) must form a regular sequence on SUS_{U^{\prime}} so that the complex is exact except at Koszul degree zero where we find

(32) H(Grass(r,n))SU/Ir,nH^{\bullet}(\operatorname{Grass}(r,n))\cong S_{U^{\prime}}/I_{r,n}

with Ir,nI_{r,n} the weighted homogeneoeus ideal generated by the elements (31).

Remark 3.5.

This model for the cohomology is linked to the geometry of the Grassmannians as follows. Let

0𝒮𝒪n𝒬00\to\mathcal{S}\to\mathcal{O}^{n}\to\mathcal{Q}\to 0

be the tautological sequence on Grass(r,n)\operatorname{Grass}(r,n). Then modulo InI_{n} we find that cj=cj(𝒮)c_{j}=c_{j}(\mathcal{S}) is the jj-th Chern class of the tautological bundle and sk=ck(𝒬)s_{k}=c_{k}(\mathcal{Q}) the kk-th Chern class of the tautological quotient bundle. The latter are nothing but the kk-th Segre classes of 𝒮\mathcal{S} which gives us precisely the relations (31) by expansion of the product of total Chern classes

ct(𝒮)ct(𝒬)=c(𝒪n)=1c_{t}(\mathcal{S})\cdot c_{t}(\mathcal{Q})=c(\mathcal{O}^{n})=1

in all degrees.

This cohomological model can be simplified further. Let x0=1,x1,,xrx_{0}=1,x_{1},\dots,x_{r} be the Chern classes of the tautological subbundle and y0=1,y1,,ynry_{0}=1,y_{1},\dots,y_{n-r} those of the tautological quotient bundle. The relations given by the images D(1εi)D(1\otimes\varepsilon_{i}), i.e. the generators of the ideal Ir,nI_{r,n} are

0=x1+y10=x2+x1y1+y20=xr+xr1y1++yr0=xry1+xr1y2++yr+10=xryn2r+xr1yn2r+1++ynr0=xryn2r1+xr1yn2r++x1ynr0=xryn2r+xr1yn2r+1++x2ynr0=xrynr2+xr1ynr1+xr2ynr0=xrynr1+xr1ynr0=xrynr.\begin{matrix}0=&x_{1}&+&y_{1}\\ 0=&x_{2}&+&x_{1}\cdot y_{1}&+&y_{2}\\ \vdots&\vdots&&\vdots&&&\ddots\\ 0=&x_{r}&+&x_{r-1}\cdot y_{1}&+&\dots&+&y_{r}\\ 0=&x_{r}\cdot y_{1}&+&x_{r-1}\cdot y_{2}&+&\dots&+&y_{r+1}\\ \vdots&\vdots&&\vdots&&&&\vdots\\ 0=&x_{r}\cdot y_{n-2r}&+&x_{r-1}\cdot y_{n-2r+1}&+&\dots&+&y_{n-r}\\ 0=&x_{r}\cdot y_{n-2r-1}&+&x_{r-1}\cdot y_{n-2r}&+&\dots&+&x_{1}\cdot y_{n-r}\\ 0=&x_{r}\cdot y_{n-2r}&+&x_{r-1}\cdot y_{n-2r+1}&+&\dots&+&x_{2}\cdot y_{n-r}\\ \vdots&&\ddots&&\ddots&&&\vdots\\ 0=&&&x_{r}\cdot y_{n-r-2}&+&x_{r-1}\cdot y_{n-r-1}&+&x_{r-2}\cdot y_{n-r}\\ 0=&&&&&x_{r}\cdot y_{n-r-1}&+&x_{r-1}\cdot y_{n-r}\\ 0=&&&&&&&x_{r}\cdot y_{n-r}.\end{matrix}

The first nrn-r equations can be used to eliminate all of the yy-variables and express them in terms of xx. Substituting these into the last rr equations we obtain polynomials in xx which we denote by

h1(n),h2(n),,hr(n).h_{1}^{(n)},h_{2}^{(n)},\dots,h_{r}^{(n)}.

If we let Jr,n=h1(n),h2(n),,hr(n)[x1,,xr]J_{r,n}=\left\langle h_{1}^{(n)},h_{2}^{(n)},\dots,h_{r}^{(n)}\right\rangle\subset\mathbb{Z}[x_{1},\dots,x_{r}] be the ideal generated by these elements then

(33) H(Grass(r,n))[x1,,xr]/Jr,n.H^{\bullet}(\operatorname{Grass}(r,n))\cong\mathbb{Z}[x_{1},\dots,x_{r}]/J_{r,n}.

Moreover, the polynomials hk(n)h_{k}^{(n)} satisfy a recurrence relation that can easily be derived from their explicit construction. Writing them in a vector we have

(h1(0)h2(0)hr(0))T=(x1x2xr)T\begin{pmatrix}h_{1}^{(0)}&h_{2}^{(0)}&\dots&h_{r}^{(0)}\end{pmatrix}^{T}=\begin{pmatrix}x_{1}&x_{2}&\dots&x_{r}\end{pmatrix}^{T}

and

(34) (h1(n+1)h2(n+1)hr1(n+1)hr(n+1))=(x1100x2010xr1001xr000)(h1(n)h2(n)hr1(n)hr(n))\begin{pmatrix}h_{1}^{(n+1)}\\ h_{2}^{(n+1)}\\ \vdots\\ h_{r-1}^{(n+1)}\\ h_{r}^{(n+1)}\end{pmatrix}=\begin{pmatrix}-x_{1}&1&0&\cdots&0\\ -x_{2}&0&1&\ddots&\vdots\\ \vdots&\vdots&\ddots&\ddots&0\\ -x_{r-1}&0&\cdots&0&1\\ -x_{r}&0&0&\cdots&0\end{pmatrix}\cdot\begin{pmatrix}h_{1}^{(n)}\\ h_{2}^{(n)}\\ \vdots\\ h_{r-1}^{(n)}\\ h_{r}^{(n)}\end{pmatrix}

Comparing this with the construction of the infinite Grassmannian Grass(r,)=n=rGrass(r,n)\operatorname{Grass}(r,\infty)=\bigcup_{n=r}^{\infty}\operatorname{Grass}(r,n) we see the following: Since the cohomology ring of Grass(r,n)\operatorname{Grass}(r,n) is generated by the Chern classes x1,,xrx_{1},\dots,x_{r} of the tautological bundle for every nn and the tautological bundle on Grass(r,n+1)\operatorname{Grass}(r,n+1) restricts to the one on Grass(r,n)\operatorname{Grass}(r,n), the pullback in cohomology for the inclusion Grass(r,n)Grass(r,n+1)\operatorname{Grass}(r,n)\hookrightarrow\operatorname{Grass}(r,n+1) is given by

[x1,,xr]/Jr,n+1[x1,,xr]/Jr,n\mathbb{Z}[x_{1},\dots,x_{r}]/J_{r,n+1}\to\mathbb{Z}[x_{1},\dots,x_{r}]/J_{r,n}

where the containment of ideals Jr,n+1Jr,nJ_{r,n+1}\subset J_{r,n} is confirmed by the recurrence relation (34) above.

3.5. Cohomology of the matrix orbits

This section will entirely consist of the proof of the following:

Proposition 3.6.

Let 0<rmn0<r\leq m\leq n be integers. The cohomology of the orbit variety Om,nrO_{m,n}^{r} is graded isomorphic to the Koszul algebra

H(Om,nr)(Rmrη1Rmrηr)H^{\bullet}\left(O_{m,n}^{r}\right)\cong\bigwedge\left(R_{m}^{r}\cdot\eta_{1}\oplus\dots\oplus R_{m}^{r}\cdot\eta_{r}\right)

over the ring Rmr=H(Grass(r,m))R_{m}^{r}=H^{\bullet}(\operatorname{Grass}(r,m)) with each ηj\eta_{j} a free generator of degree 2n2j+12n-2j+1.

We use the Cartan model (27) for Om,nrO_{m,n}^{r}. The group in question is Um,n=U(m)×U(n)U_{m,n}=U(m)\times U(n) with the subgroup

Um,nr\displaystyle U_{m,n}^{r} :=\displaystyle:= Stab(𝟏m,nr)\displaystyle\operatorname{Stab}(\mathbf{1}_{m,n}^{r})
=\displaystyle= {(SP,SQ)U(m)×U(n)}\displaystyle\left\{(S\oplus P,\,S\oplus Q)\in U(m)\times U(n)\right\}
\displaystyle\cong U(r)×U(mr)×U(nr).\displaystyle U(r)\times U(m-r)\times U(n-r).

As a maximal torus of this subgroup we choose pairs of diagonal block matrices

(diag(λ¯)diag(μ¯),diag(λ¯)diag(ν¯))\left(\operatorname{diag}(\underline{\lambda})\oplus\operatorname{diag}(\underline{\mu}),\quad\operatorname{diag}(\underline{\lambda})\oplus\operatorname{diag}(\underline{\nu})\right)

with all nontrivial entries λ1,,λr,μ1,,μmr,ν1,,νnrU(1)\lambda_{1},\dots,\lambda_{r},\mu_{1},\dots,\mu_{m-r},\nu_{1},\dots,\nu_{n-r}\in U(1). This is contained in the maximal torus TT of Um,nU_{m,n} in the obvious way.

We will write Sm,nrS_{m,n}^{r} for the ring H(BUm,nr)H^{\bullet}(BU_{m,n}^{r}). If we let {αj}j=1r\{\alpha_{j}\}_{j=1}^{r} be the set of Chern roots associated to the subgroup U(r)U(r), {βj}j=1mr\{\beta_{j}\}_{j=1}^{m-r} the ones for U(mr)U(m-r) and {γj}j=1nr\{\gamma_{j}\}_{j=1}^{n-r} those for U(nr)U(n-r), then similar to the case of flag varieties, the ring Sm,nrS_{m,n}^{r} is the invariant subring of [α¯,β¯,γ¯]\mathbb{Z}[\underline{\alpha},\underline{\beta},\underline{\gamma}] by the action of the group

𝔖m,nr:=𝔖r×𝔖mr×𝔖nr\mathfrak{S}_{m,n}^{r}:=\mathfrak{S}_{r}\times\mathfrak{S}_{m-r}\times\mathfrak{S}_{n-r}

acting by the permutation of the distinct sets of variables. We set

xj\displaystyle x_{j} =\displaystyle= σjr(α1,,αr) for j=1,,r,\displaystyle\sigma_{j}^{r}(\alpha_{1},\dots,\alpha_{r})\quad\textnormal{ for }j=1,\dots,r,
yk\displaystyle y_{k} =\displaystyle= σkmr(β1,,βmr) for k=1,,mr,\displaystyle\sigma_{k}^{m-r}(\beta_{1},\dots,\beta_{m-r})\quad\textnormal{ for }k=1,\dots,m-r,
zl\displaystyle z_{l} =\displaystyle= σlnr(γ1,,γnr) for l=1,,nr\displaystyle\sigma_{l}^{n-r}(\gamma_{1},\dots,\gamma_{n-r})\quad\textnormal{ for }l=1,\dots,n-r

so that Sm,ni1,,ip[c¯,y¯,z¯]S_{m,n}^{i_{1},\dots,i_{p}}\cong\mathbb{Z}[\underline{c},\underline{y},\underline{z}] is a free polynomial ring containing [x¯,y¯,z¯]\mathbb{Z}[\underline{x},\underline{y},\underline{z}] as another free polynomial subring.

Since the group Um,n=U(m)×U(n)U_{m,n}=U(m)\times U(n) is a product, also the exterior algebra in (27) takes the form of a product:

(FmFn)(Fm)(Fn).\bigwedge(F_{m}\oplus F_{n})\cong\left(\bigwedge F_{m}\right)\otimes_{\mathbb{Z}}\left(\bigwedge F_{n}\right).

We let {εj}j=1m\{\varepsilon_{j}\}_{j=1}^{m} be the generators of FmF_{m} and {εk}k=1n\{\varepsilon^{\prime}_{k}\}_{k=1}^{n} those of FnF_{n}. With the notation above it is easy to see from the inclusion of maximal tori that the differential DD of (27) takes these generators to the elements

D(1εj)\displaystyle D(1\otimes\varepsilon_{j}) =\displaystyle= r+s=jxrys,\displaystyle\sum_{r+s=j}x_{r}\cdot y_{s},
D(1εk)\displaystyle D(1\otimes\varepsilon^{\prime}_{k}) =\displaystyle= r+s=kxrzs\displaystyle\sum_{r+s=k}x_{r}\cdot z_{s}

in Sm,ni1,,ipS_{m,n}^{i_{1},\dots,i_{p}} and, as already discussed in Remark 3.5, these are precisely the relations between the Chern classes of the tautological sub- and quotient bundles.

We now consider Koszul complexes on Sm,ni1,,ipS_{m,n}^{i_{1},\dots,i_{p}} associated to various subsets of the generators D(1εj)D(1\otimes\varepsilon_{j}) and D(1εk)D(1\otimes\varepsilon^{\prime}_{k}). As discussed earlier in Remark 3.5 the elements

D(1ε1),D(1εmr),D(1ε1),D(1εnr)D(1\otimes\varepsilon_{1}),\dots D(1\otimes\varepsilon_{m-r}),D(1\otimes\varepsilon^{\prime}_{1}),\dots D(1\otimes\varepsilon^{\prime}_{n-r})

form a regular sequence on [x¯,y¯,z¯]\mathbb{Z}[\underline{x},\underline{y},\underline{z}] that can be used to eliminate the variables y¯\underline{y} and z¯\underline{z}. Using the reduction lemma for the homology of Koszul complexes, Lemma 3.4, this reduces the problem to a Koszul complex on the ring

[x1,,xr,y1,,ymr,z1,,znr]D(1ε1),,D(1εmr),D(1ε1),,D(1εnr)[x1,,xr]\frac{\mathbb{Z}[x_{1},\dots,x_{r},y_{1},\dots,y_{m-r},z_{1},\dots,z_{n-r}]}{\langle D(1\otimes\varepsilon_{1}),\dots,D(1\otimes\varepsilon_{m-r}),D(1\otimes\varepsilon^{\prime}_{1}),\dots,D(1\otimes\varepsilon^{\prime}_{n-r})\rangle}\cong\mathbb{Z}[x_{1},\dots,x_{r}]

In this quotient, the next rr relations reduce to

D(1εmr+1)¯=h1(m),,D(1εm)¯=hr(m)\overline{D(1\otimes\varepsilon_{m-r+1})}=h^{(m)}_{1},\quad\dots,\quad\overline{D(1\otimes\varepsilon_{m})}=h^{(m)}_{r}

and they form another regular sequence with successive quotient H(Grass(r,n))=[x1,,xr]/Jr,nH^{\bullet}(\operatorname{Grass}(r,n))=\mathbb{Z}[x_{1},\dots,x_{r}]/J_{r,n} as in Remark 3.5. Consequently, the last rr relations

D(1εnr+1)¯=h1(n),,D(1εn)¯=hr(n)\overline{D(1\otimes\varepsilon^{\prime}_{n-r+1})}=h^{(n)}_{1},\quad\dots,\quad\overline{D(1\otimes\varepsilon^{\prime}_{n})}=h^{(n)}_{r}

all reduce to zero in H(Grass(r,m))H^{\bullet}(\operatorname{Grass}(r,m)) due to (34) since by assumption mnm\leq n. In terms of the identification of the homology of Koszul complexes from Lemma 3.4 this means

H([x,y,z](FmFn))\displaystyle H\left(\mathbb{Z}[x,y,z]\otimes_{\mathbb{Z}}\bigwedge\left(F_{m}\oplus F_{n}\right)\right)
\displaystyle\cong H([x](εmr+1εmεnr+1εn))\displaystyle H\left(\mathbb{Z}[x]\otimes_{\mathbb{Z}}\bigwedge\left(\mathbb{Z}\varepsilon_{m-r+1}\oplus\dots\oplus\mathbb{Z}\varepsilon_{m}\oplus\mathbb{Z}\varepsilon^{\prime}_{n-r+1}\oplus\dots\oplus\mathbb{Z}\varepsilon^{\prime}_{n}\right)\right)
\displaystyle\cong H([x]/Jm,r(εnr+1εn))\displaystyle H\left(\mathbb{Z}[x]/J_{m,r}\otimes_{\mathbb{Z}}\bigwedge\left(\mathbb{Z}\varepsilon_{n-r+1}^{\prime}\oplus\dots\oplus\mathbb{Z}\varepsilon^{\prime}_{n}\right)\right)
\displaystyle\cong [x]/Jm,r(εnr+1εn)\displaystyle\mathbb{Z}[x]/J_{m,r}\otimes_{\mathbb{Z}}\bigwedge\left(\mathbb{Z}\varepsilon_{n-r+1}^{\prime}\oplus\dots\oplus\mathbb{Z}\varepsilon^{\prime}_{n}\right)

where the differentials of the complex in the second last line are all zero so that we may drop the homology functor H()H(-). To finish the proof of Proposition 3.6 we now set ηj\eta_{j} to be equal to 1εnr+j1\otimes\varepsilon^{\prime}_{n-r+j} for j=1,,rj=1,\dots,r in the last line.

Remark 3.7.

We review the fiber bundle structure

Stief(r,n)Om,nrλ1Grass(r,m)\operatorname{Stief}(r,n)\hookrightarrow O_{m,n}^{r}\overset{\lambda_{1}}{\longrightarrow}\operatorname{Grass}(r,m)

described in Lemma 3.2. Given the explicit descriptions of the cohomology of both the Stiefel manifold and the Grassmannian, we may infer from the proof of Proposition 3.6 that, similar to the case the Hopf’s theorem, the cohomology of Om,nrO_{m,n}^{r} is isomorphic to that of a product

H(Om,nr)H(Stief(r,n)×Grass(r,m)),H^{\bullet}(O_{m,n}^{r})\cong H^{\bullet}(\operatorname{Stief}(r,n)\times\operatorname{Grass}(r,m)),

despite the fact that the structure of Om,nrO_{m,n}^{r} as a fiber bundle is in general non-trivial. By construction, the cohomology classes ηj\eta_{j} restrict to the generators of the cohomology of Stief(r,n)\operatorname{Stief}(r,n) in every fiber. Yet it seems difficult to write down an explicit lift of these elements to the original complex [x,y,z](FmFn)\mathbb{Z}[x,y,z]\otimes\bigwedge\left(F_{m}\oplus F_{n}\right).

Remark 3.8.

It would be nice to have a more geometric understanding of the elements ηj\eta_{j} that might, for example, be derived from the tautological sequence

(35) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ2Q2\textstyle{\lambda_{2}^{*}Q_{2}^{\vee}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪n\textstyle{\mathcal{O}^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}𝒪m\textstyle{\mathcal{O}^{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ1Q1\textstyle{\lambda_{1}^{*}Q_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

on Om,nrO_{m,n}^{r} where the QiQ_{i} denote the tautological quotient bundles on respective Grassmannians for the projections λ1\lambda_{1} and λ2\lambda_{2} as in Lemma 3.2 and φ\varphi denotes the tautological section

Om,nrHom(𝒪n,𝒪m),φφO_{m,n}^{r}\to\operatorname{Hom}(\mathcal{O}^{n},\mathcal{O}^{m}),\quad\varphi\mapsto\varphi

on Om,nrO_{m,n}^{r}, identifying the subbundles λ2S2\lambda_{2}^{*}S_{2}^{\vee} with λ1S1\lambda_{1}^{*}S_{1}.

4. Betti numbers of smooth links

A hyperplane (Di,0)(m×n,0)(D_{i},0)\subset(\mathbb{C}^{m\times n},0) of codimension ii in general position will intersect Mm,nr+1M_{m,n}^{r+1} in a non-trivial way whenever idimMm,nr+1i\leq\dim M_{m,n}^{r+1}. This intersection (Mm,nr+1Di,0)(M_{m,n}^{r+1}\cap D_{i},0) will have isolated singularity as long as DiD_{i} meets the singular locus Mm,nrM_{m,n}^{r} of Mm,nr+1M_{m,n}^{r+1} only at the origin, i.e. idimMm,nri\geq\dim M_{m,n}^{r}. The real and complex links of codimension ii of (Mm,nr+1,0)(M_{m,n}^{r+1},0) will therefore be smooth within the range

(m+n)(r1)(r1)2i<(m+n)rr2.(m+n)(r-1)-(r-1)^{2}\leq i<(m+n)r-r^{2}.

In this section we will be concerned with the cohomology of 𝒦i(Mm,nr+1,0)\mathcal{K}^{i}(M_{m,n}^{r+1},0) and i(Mm,nr+1,0)\mathcal{L}^{i}(M_{m,n}^{r+1},0) for ii in this range. More precisely, we will show for that for ii in the above range

(36) Hk(Grass(r,m))\displaystyle H^{k}(\operatorname{Grass}(r,m)) \displaystyle\cong Hk(𝒦i(Mm,nr+1,0)) for k<d(i),\displaystyle H^{k}(\mathcal{K}^{i}(M_{m,n}^{r+1},0))\quad\textnormal{ for }\quad k<d(i),
(37) Hk(Grass(r,m))\displaystyle H^{k}(\operatorname{Grass}(r,m)) \displaystyle\cong Hk(i(Mm,nr+1,0)) for k<d(i),\displaystyle H^{k}(\mathcal{L}^{i}(M_{m,n}^{r+1},0))\quad\textnormal{ for }\quad k<d(i),
(38) Hd(i)(Grass(r,m))\displaystyle H^{d(i)}(\operatorname{Grass}(r,m)) \displaystyle\subset Hd(i)(𝒦i(Mm,nr+1,0))Hd(i)(i(Mm,nr+1,0))\displaystyle H^{d(i)}(\mathcal{K}^{i}(M_{m,n}^{r+1},0))\subset H^{d(i)}(\mathcal{L}^{i}(M_{m,n}^{r+1},0))

where d(i)=dimi(Mm,nr+1,0)=(m+n)rr2i1d(i)=\dim_{\mathbb{C}}\mathcal{L}^{i}(M_{m,n}^{r+1},0)=(m+n)r-r^{2}-i-1. The left hand sides all agree with the cohomology of Vm,nrV_{m,n}^{r} in this range for kk and the above maps are given by the pullback in cohomology for the natural inclusions of the real and complex links into that stratum.

For the complex links the middle Betti number can then be computed from the polar multiplicities using Formula (11). In case r=1r=1 this gives a complete description of the classical real and complex links of (Mm,n2,0)(M_{m,n}^{2},0), see Remark 4.5.

However, besides the case i=0i=0 and r=1r=1, we are unable to determine whether or not Hd(i)(i(Mm,nr+1,0))H^{d(i)}(\mathcal{L}^{i}(M_{m,n}^{r+1},0)) has torsion. Also, the two middle cohomology groups of 𝒦i(Mm,nr+1,0)\mathcal{K}^{i}(M_{m,n}^{r+1},0) can not be computed by our methods for i>0i>0.

4.1. The variation sequence on stratified spaces

Throughout this section, let (X,0)(q,0)(X,0)\subset(\mathbb{C}^{q},0) be an equidimensional reduced complex analytic germ of complex dimension dd, endowed with a complex analytic Whitney stratification {Vα}αA\{V^{\alpha}\}_{\alpha\in A}. We will denote by f:(n,0)(,0)f\colon(\mathbb{C}^{n},0)\to(\mathbb{C},0) the germ of a holomorphic function with an isolated singularity on (X,0)(X,0) in the stratified sense. For a Milnor ball BεB_{\varepsilon} of sufficiently small radius ε>0\varepsilon>0 and representatives XX and ff of the space and the function we let

X:=BεX\partial X:=\partial B_{\varepsilon}\cap X

be the real link of (X,0)(X,0) and

M=XBεf1({δ})M=X\cap B_{\varepsilon}\cap f^{-1}(\{\delta\})

the Milnor fiber of ff for some εδ>0\varepsilon\gg\delta>0. Denote the regular part of XX by XregX_{\mathrm{reg}}. Then clearly by construction of X\partial X and MM

Xreg:=BεXreg,Mreg:=MXreg\partial X_{\mathrm{reg}}:=\partial B_{\varepsilon}\cap X_{\mathrm{reg}},\quad M_{\mathrm{reg}}:=M\cap X_{\mathrm{reg}}

are the regular loci of X\partial X and MM, respectively.

Lemma 4.1.

The natural maps in cohomology

Hk(Xreg)Hk(Mreg)H^{k}(\partial X_{\mathrm{reg}})\to H^{k}(M_{\mathrm{reg}})

are injective for k=d1k=d-1 and an isomorphism for k<d1k<d-1.

Proof.

It is well known that we may “inflate” the Milnor-Lê fibration for ff and identify its total space with an open subset of X\partial X itself:

XBεf1(Dδ)\textstyle{X\cap B_{\varepsilon}\cap f^{-1}\left(\partial D_{\delta}\right)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}f\scriptstyle{f}Xf1(Dδ)\textstyle{\partial X\setminus f^{-1}(D_{\delta})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}argf\scriptstyle{\arg f}Dδ\textstyle{\partial D_{\delta}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}S1.\textstyle{S^{1}.}

See for example [GM88, Part II, Chapter 2.A, Proposition 2.A.3], or [GM88, Part II, Chapter 2.4] for the particular case where ff is linear.

Since ff has isolated singularity on (X,0)(X,0) it is a stratified submersion on XX near the boundary K:=Xf1({0})K:=\partial X\cap f^{-1}(\{0\}) of the central fiber. We may therefore identify

Xf1(Dδ)K×Dδ\partial X\cap f^{-1}(D_{\delta})\cong K\times D_{\delta}

and extend the fibration by argf\arg f over K×DδK\times D_{\delta}^{*} to the whole complement of KXK\subset\partial X.

The assertion is now a consequence of the variation sequence for the regular loci. It is evident from Thom’s first isotopy lemma that

argf:XKS1\arg f\colon\partial X\setminus K\to S^{1}

respects the stratification of MXKM\subset\partial X\setminus K induced from XX. Therefore, the common chain of isomorphisms for the variation sequence

Hk(Xreg,Mreg)\displaystyle H^{k}\left(\partial X_{\mathrm{reg}},M_{\mathrm{reg}}\right) \displaystyle\cong Hk(Xreg,Mreg(Xregf1(D¯δ)))\displaystyle H^{k}\left(\partial X_{\mathrm{reg}},M_{\mathrm{reg}}\cup\left(\partial X_{\mathrm{reg}}\cap f^{-1}(\overline{D}_{\delta})\right)\right)
\displaystyle\cong Hk(Mreg×[0,1],Kreg×[0,1]Mreg×{0,1})\displaystyle H^{k}\left(M_{\mathrm{reg}}\times[0,1],K_{\mathrm{reg}}\times[0,1]\cup M_{\mathrm{reg}}\times\{0,1\}\right)
\displaystyle\cong Hk((Mreg,Kreg)×([0,1],[0,1]))\displaystyle H^{k}\left((M_{\mathrm{reg}},K_{\mathrm{reg}})\times([0,1],\partial[0,1])\right)
\displaystyle\cong Hk1(Mreg,Kreg)\displaystyle H^{k-1}(M_{\mathrm{reg}},K_{\mathrm{reg}})

restricts to the regular loci for every kk. Here, we deliberately identified KregK_{\mathrm{reg}} with the boundary part BεMreg\partial B_{\varepsilon}\cap M_{\mathrm{reg}} of the regular locus of MM. In the last step we used the Künneth formula for pairs of spaces, see e.g. [Hat02, Theorem 3.21]. For a more thorough treatment see also [GM88, Part II, Chapter 2.5].

With these identifications, the long exact sequence of the pair (Xreg,Mreg)(\partial X_{\mathrm{reg}},M_{\mathrm{reg}}) reads

Hk1(Mreg,Kreg)Hk(Xreg)Hk(Mreg)Hk(Mreg,Kreg)\dots\to H^{k-1}(M_{\mathrm{reg}},K_{\mathrm{reg}})\to H^{k}(\partial X_{\mathrm{reg}})\to H^{k}(M_{\mathrm{reg}})\to H^{k}(M_{\mathrm{reg}},K_{\mathrm{reg}})\to\dots

The assertion now follows from the fact that Hk(Mreg,Kreg)=0H^{k}(M_{\mathrm{reg}},K_{\mathrm{reg}})=0 for all k<d1=dimMregk<d-1=\dim_{\mathbb{C}}M_{\mathrm{reg}} which is due to complex stratified Morse theory for non-proper Morse functions, see [GM88, Introduction, Chapter 2.2]. ∎

Remark 4.2.

The cited theorem [GM88, Introduction, Chapter 2.2] is a statement on the relative homology for the smooth locus of a projective algebraic variety and its intersection with a generic hyperplane. The statement can be generalized to germs of complex analytic sets and their intersections with a suitable ball, cf. [GM88, Introduction, Chapter 2.4]. Then one can use complex stratified Morse theory for non-proper Morse functions to study the connectivity of the pair (Mreg,Kreg)(M_{\mathrm{reg}},K_{\mathrm{reg}}) via a Morsification ρ\rho of the squared distance function to the origin. The key observation is that for the regular locus MregM_{\mathrm{reg}} the local Morse datum at a critical point is always (d2)(d-2)-connected. The last statement can easily be derived by induction on dimension from [GM88, Part II, Chapter 3.3, Corollary 1].

Now suppose we are given (X,0)(q,0)(X,0)\subset(\mathbb{C}^{q},0) as above and an admissible sequence of linear forms l1,,ldl_{1},\dots,l_{d} in the sense of Definition 2.2 and Corollary 2.4. Then we can apply Lemma 4.1 inductively to f=li+1f=l_{i+1} on (XDi,0)(X\cap D_{i},0):

Proposition 4.3.

Let (X,0)(q,0)(X,0)\subset(\mathbb{C}^{q},0) be an equidimensional reduced complex analytic germ of dimension dd, and l1,l2,,ldl_{1},l_{2},\dots,l_{d} an admissible sequence of linear forms as in Corollary 2.4. Let 𝒦regk\mathcal{K}^{k}_{\mathrm{reg}} and regk\mathcal{L}^{k}_{\mathrm{reg}} be the regular loci of the real and complex links of codimension kk of (X,0)(X,0). Then for arbitrary 0k<d0\leq k<d the natural maps

Hi(𝒦reg0)Hi(𝒦regk)Hi(regk)H^{i}(\mathcal{K}^{0}_{\mathrm{reg}})\to H^{i}(\mathcal{K}^{k}_{\mathrm{reg}})\to H^{i}(\mathcal{L}^{k}_{\mathrm{reg}})

are isomorphisms for every i<dk1=dimki<d-k-1=\dim\mathcal{L}^{k} and injective for i=dk1i=d-k-1.

Proof.

We proceed by induction on the codimension kk of the complex links. For k=0k=0 this follows directly from Lemma 4.1.

Now suppose the statement holds up to codimension k1k-1. As already discussed in the proof of Lemma 4.1, the real link 𝒦k\mathcal{K}^{k} can be identified with the boundary of k1\mathcal{L}^{k-1}. But the pair (regk1,𝒦regk)(\mathcal{L}^{k-1}_{\mathrm{reg}},\mathcal{K}^{k}_{\mathrm{reg}}) is cohomologically dk1d-k-1-connected due to the LHT and the long exact sequence yields

0Hi(regk1)Hi(𝒦regk)00\to H^{i}(\mathcal{L}^{k-1}_{\mathrm{reg}})\overset{\cong}{\longrightarrow}H^{i}(\mathcal{K}^{k}_{\mathrm{reg}})\to 0

for i<dk1i<d-k-1 and

0Hdk1(regk1)Hdk1(𝒦regk)Hdk(regk1,𝒦regk)0\to H^{d-k-1}(\mathcal{L}^{k-1}_{\mathrm{reg}})\to H^{d-k-1}(\mathcal{K}^{k}_{\mathrm{reg}})\to H^{d-k}(\mathcal{L}^{k-1}_{\mathrm{reg}},\mathcal{K}^{k}_{\mathrm{reg}})\to\cdots

for the middle part. Since by our induction hypothesis the cohomology groups Hi(regk1)Hi(𝒦reg0)H^{i}(\mathcal{L}^{k-1}_{\mathrm{reg}})\cong H^{i}(\mathcal{K}^{0}_{\mathrm{reg}}) are all isomorphic for i<dki<d-k, the first part of the assertion on Hi(𝒦reg0)Hi(𝒦regk)H^{i}(\mathcal{K}^{0}_{\mathrm{reg}})\to H^{i}(\mathcal{K}^{k}_{\mathrm{reg}}) follows.

For the second part on Hi(𝒦reg0)Hi(regk)H^{i}(\mathcal{K}^{0}_{\mathrm{reg}})\to H^{i}(\mathcal{L}^{k}_{\mathrm{reg}}) consider

lk+1:(XDk,0)(,0).l_{k+1}\colon(X\cap D_{k},0)\to(\mathbb{C},0).

By construction the Milnor fiber of this function is k\mathcal{L}^{k} while 𝒦k\mathcal{K}^{k} is the real link of (XDk,0)(X\cap D_{k},0). All the relevant cohomology groups of 𝒦k\mathcal{K}^{k} have already been determined by our previous considerations so that the remaining statements follow readily from Lemma 4.1 applied to f=lk+1f=l_{k+1}.

4.2. Proof of Formulas (36) – (38)

As remarked earlier, the relevant range for the codimension ii for the real and complex links 𝒦i(Mm,nr+1,0)\mathcal{K}^{i}(M_{m,n}^{r+1},0) and i(Mm,nr+1,0)\mathcal{L}^{i}(M_{m,n}^{r+1},0) of the generic determinantal variety (Mm,nr+1,0)(M_{m,n}^{r+1},0) is

dimMm,nr=(m+n)(r1)(r1)2i<(m+n)rr2=dimMm,nr+1.\dim M_{m,n}^{r}=(m+n)(r-1)-(r-1)^{2}\leq i<(m+n)r-r^{2}=\dim M_{m,n}^{r+1}.

We may assume mnm\leq n. Then for these values of ii the dimension of the complex links does not exceed

d(i):=dimi(Mm,nr+1,0)=(m+n)rr2i1<m+n2r+12n2r+1.d(i):=\dim_{\mathbb{C}}\mathcal{L}^{i}(M_{m,n}^{r+1},0)=(m+n)r-r^{2}-i-1<m+n-2r+1\leq 2n-2r+1.

Recall that due to the homogeneity of the singularity, the cohomology of the regular part of the classical real link of (Mm,nr+1,0)(M_{m,n}^{r+1},0) is given by

H(𝒦reg0(Mm,nr+1,0))\displaystyle H^{\bullet}(\mathcal{K}^{0}_{\mathrm{reg}}(M_{m,n}^{r+1},0)) \displaystyle\cong H(Vm,nr)H(Om,nr)\displaystyle H^{\bullet}(V_{m,n}^{r})\cong H^{\bullet}(O_{m,n}^{r})
\displaystyle\cong (Rmrη1Rmrηr)\displaystyle\bigwedge\left(R_{m}^{r}\cdot\eta_{1}\oplus\dots\oplus R_{m}^{r}\cdot\eta_{r}\right)

according to Proposition 3.6. Now note that the bound on d(i)d(i) assures that each and every generator ηjH2n2j+1(Om,nr)\eta_{j}\in H^{2n-2j+1}(O_{m,n}^{r}) is taken to a vanishing cohomology group of a smooth complex link i(Mm,nr+1,0)\mathcal{L}^{i}(M_{m,n}^{r+1},0). The formulas (36), (37), and (38) now follow directly from Proposition 4.3.

4.3. The middle cohomology groups

The results of the previous section allow to determine the cohomology below the middle degrees of the real and complex links 𝒦i\mathcal{K}^{i} and i\mathcal{L}^{i} of codimension ii for a purely dd-dimensional germ (X,0)(X,0), provided idimXsingi\geq\dim X_{\mathrm{sing}} so that (XDi,0)(X\cap D_{i},0) has isolated singularity and the links are smooth. We shall now discuss how to obtain information on the remaining part of the cohomology and to which extend this is possible.

When idimXsingi\geq\dim X_{\mathrm{sing}} the complex link i=regi\mathcal{L}^{i}=\mathcal{L}^{i}_{\mathrm{reg}} is Stein and the higher cohomology groups vanish due to the LHT. We can use Lefschetz duality (see e.g. [Hat02, Theorem 3.43]) to identify homology with cohomology:

Hk(i,i)H2(di1)k(i)andHk(i)H2(di1)k(i,i).H^{k}(\mathcal{L}^{i},\partial\mathcal{L}^{i})\cong H_{2(d-i-1)-k}(\mathcal{L}^{i})\quad\textnormal{and}\quad H^{k}(\mathcal{L}^{i})\cong H_{2(d-i-1)-k}(\mathcal{L}^{i},\partial\mathcal{L}^{i}).

Note that the middle homology group Hdi1(i)H_{d-i-1}(\mathcal{L}^{i}) is known to always be free; this is the case even for arbitrary smooth Stein manifolds. The real link 𝒦i\mathcal{K}^{i} is an oriented smooth compact manifold of real dimension 2d2i12d-2i-1 and we have Poincaré duality (cf. [Hat02, Theorem 3.30]):

Hk(𝒦i)H2d2i1k(𝒦i)H^{k}(\mathcal{K}^{i})\cong H_{2d-2i-1-k}(\mathcal{K}^{i})

for all kk. All these cohomology groups sit in the classical variation sequence with its middle part being of particular interest:

0Hdi1(𝒦i)Hdi1(i)VARHdi1(i,i)Hdi(𝒦i)0.0\to H^{d-i-1}(\mathcal{K}^{i})\to H^{d-i-1}(\mathcal{L}^{i})\overset{\mathrm{VAR}}{\longrightarrow}H^{d-i-1}(\mathcal{L}^{i},\partial\mathcal{L}^{i})\to H^{d-i}(\mathcal{K}^{i})\to 0.

We return to the particular case of the generic determinantal singularity (X,0)=(Mm,nr+1,0)(m×n,0)(X,0)=(M_{m,n}^{r+1},0)\subset(\mathbb{C}^{m\times n},0). For the smooth complex links i=i(Mm,nr+1,0)\mathcal{L}^{i}=\mathcal{L}^{i}(M_{m,n}^{r+1},0) we can use the knowledge of the Euler characteristic from the polar multiplicities in Section 2.3, Formula (11) to also determine the rank of the middle cohomology group:

(39) bdi1(i)=(j=id(1)jm0(Pdj(X,0)))(1)di1χ(i)+(j=0di2(1)dijbj(Om,nr)).(1)di1j=0di2(1)jbj(i)b_{d-i-1}(\mathcal{L}^{i})=\underbrace{\left(\sum_{j=i}^{d}(-1)^{j}m_{0}\left(P_{d-j}(X,0)\right)\right)}_{(-1)^{d-i-1}\chi(\mathcal{L}^{i})}+\underbrace{\left(\sum_{j=0}^{d-i-2}(-1)^{d-i-j}b_{j}(O_{m,n}^{r})\right).}_{-(-1)^{d-i-1}\sum_{j=0}^{d-i-2}(-1)^{j}b_{j}(\mathcal{L}^{i})}

Switching from integer to rational coefficients, this allows us to fully compute the rational cohomology, but unfortunately this method does not allow to detect torsion in the middle cohomology group with integer coefficients. However, so far no example of a smooth complex link of a generic determinantal variety is known for which the middle cohomology group Hdi1(i(Mm,nr+1,0))H^{d-i-1}(\mathcal{L}^{i}(M_{m,n}^{r+1},0)) does have torsion.

Being an oriented smooth compact manifold of odd real dimension, χ(𝒦i)\chi(\mathcal{K}^{i}) is always zero and therefore computation of the Euler characteristic does not help in this case; not even the full rational cohomology of 𝒦i\mathcal{K}^{i} can be computed by our methods. It should be noted, however, that there are examples for which torsion appears:

Example 4.4.

Consider the singularity (Y,0):=(M2,32D2,0)(2×3,0)(Y,0):=(M_{2,3}^{2}\cap D_{2},0)\subset(\mathbb{C}^{2\times 3},0) for a generic plane D2D_{2} of codimension 22. This is an isolated normal surface singularity, the simplest one of the “rational triple points” discussed by Tjurina [Tju68]. It was shown in [Zac20] that the so-called Tjurina modification π:YY\pi\colon Y^{\prime}\to Y of (Y,0)(Y,0) is smooth and hence a resolution of singularities for (Y,0)(Y,0). The space YY^{\prime} is isomorphic to the total space of the bundle

Y|𝒪1(3)|Y^{\prime}\cong|\mathcal{O}_{\mathbb{P}^{1}}(-3)|

and hence the complex link 𝒦2=𝒦2(M2,32,0)\mathcal{K}^{2}=\mathcal{K}^{2}(M_{2,3}^{2},0) can be identified with the sphere bundle of 𝒪1(3)\mathcal{O}_{\mathbb{P}^{1}}(-3). Then a part of the Euler sequence for this bundle reads

H0(1)eH2(1)H2(𝒦2)0.\dots H^{0}(\mathbb{P}^{1})\overset{\cup e}{\longrightarrow}H^{2}(\mathbb{P}^{1})\to H^{2}(\mathcal{K}^{2})\to 0.

But with the canonical generators for H(1)H^{\bullet}(\mathbb{P}^{1}), the cup product with the Euler class ee is simply multiplication by 3-3 and we find that

(40) H2(𝒦2(M2,32,0))/3.H^{2}(\mathcal{K}^{2}(M_{2,3}^{2},0))\cong\mathbb{Z}/3\mathbb{Z}.

The complex link 2=2(M2,32,0)\mathcal{L}^{2}=\mathcal{L}^{2}(M_{2,3}^{2},0), i.e. the “Milnor fiber” in the variation sequence, is known to have middle cohomology H1(2)2H^{1}(\mathcal{L}^{2})\cong\mathbb{Z}^{2}; see e.g. [Zac20]. Since H1(2,2)H1(2)2H^{1}(\mathcal{L}^{2},\partial\mathcal{L}^{2})\cong H_{1}(\mathcal{L}^{2})\cong\mathbb{Z}^{2} is also free and the map VAR\mathrm{VAR} necessarily has rank 22, it follows that

H1(𝒦2(M2,32,0))=0.H^{1}(\mathcal{K}^{2}(M_{2,3}^{2},0))=0.

To conclude this section, we mention one last case that might be of particular interest:

Remark 4.5.

For r=1r=1 the generic determinantal varieties (Mm,n2,0)(M_{m,n}^{2},0) all have isolated singularity and therefore the classical real and complex links 𝒦0=𝒦0(Mm,n2,0)\mathcal{K}^{0}=\mathcal{K}^{0}(M_{m,n}^{2},0) and 0=0(Mm,n2,0)\mathcal{L}^{0}=\mathcal{L}^{0}(M_{m,n}^{2},0) are already smooth. In this case we know all the cohomology groups of 𝒦0\mathcal{K}^{0}

H(𝒦0(Mm,n2,0))H(m1×S2n1)H(m1)ε2n1H^{\bullet}(\mathcal{K}^{0}(M_{m,n}^{2},0))\cong H^{\bullet}(\mathbb{P}^{m-1}\times S^{2n-1})\cong H^{\bullet}(\mathbb{P}^{m-1})\otimes_{\mathbb{Z}}\bigwedge\mathbb{Z}\varepsilon_{2n-1}

which is free Abelian in all degrees. Using the variation sequence we infer that the middle cohomology of 0\mathcal{L}^{0} can not have torsion: By Poincaré duality also the relative cohomology group to the left is isomorphic to the middle homology group of 0\mathcal{L}^{0}. Now the latter is known to always be free Abelian for Stein manifolds. Therefore we have

(41) Hk(0(Mm,n2,0)){ if k(mr)r is even 0 otherwise.H^{k}(\mathcal{L}^{0}(M_{m,n}^{2},0))\cong\begin{cases}\mathbb{Z}&\textnormal{ if $k\leq(m-r)r$ is even }\\ 0&\textnormal{ otherwise.}\end{cases}

We would like to mention one particular consequence of the previous remark. It is well known that on a locally complete intersection XX the constant sheaf X[dimX]\mathbb{Z}_{X}[\dim X] is perverse for the middle perversity; see e.g. [Dim04, Theorem 5.1.20]. Most of the generic determinantal varieties Mm,nsM_{m,n}^{s} are not complete intersections and we can now show that this algebraic property is also reflected in their topology:

Corollary 4.6.

For 1<sm<n1<s\leq m<n, i.e. for non-square matrices, the constant sheaf Mm,ns[dimMm,ns]\mathbb{Z}_{M_{m,n}^{s}}[\dim M_{m,n}^{s}] on the generic determinantal variety, shifted by its dimension, is never a perverse sheaf for the middle perversity.

Proof.

For these values of s,ms,m, and nn the rank stratification is the minimal Whitney stratification of Mm,nsM_{m,n}^{s}. Now the constant sheaf can only be perverse (for the middle perversity) if the variety has a certain rectified homological depth, meaning that along every stratum Vm,nrMm,nsV_{m,n}^{r}\subset M_{m,n}^{s} the real link 𝒦(Mm,ns,Vm,nr)\mathcal{K}(M_{m,n}^{s},V_{m,n}^{r}) along Vm,nrV_{m,n}^{r} satisfies

Hk(𝒦(Mm,ns,Vm,nr))=0 for all k<codim(Vm,nr,Mm,ns)1.H_{k}(\mathcal{K}(M_{m,n}^{s},V_{m,n}^{r}))=0\textnormal{ for all }k<\operatorname{codim}(V_{m,n}^{r},M_{m,n}^{s})-1.

See [HL91, Section 1.1.3], [HL91, Theorem 1.4] for a discussion on rectified homological depth and e.g. the proof of [Dim04, Theorem 5.1.20] for its relation to perversity. For r=s2r=s-2 the codimension on the right hand side is

c=(ms+2)(ns+2)(ms+1)(ns+1)=m+n2s+34c=(m-s+2)(n-s+2)-(m-s+1)(n-s+1)=m+n-2s+3\geq 4

for all values 1<sm<n1<s\leq m<n. But according to Remark 4.5 we have

H2(𝒦0(Mm,ns,Vm,ns2))H2(𝒦0(Mms+2,ns+22))H2(ms+1)H^{2}(\mathcal{K}^{0}(M_{m,n}^{s},V_{m,n}^{s-2}))\cong H^{2}(\mathcal{K}^{0}(M_{m-s+2,n-s+2}^{2}))\cong H^{2}(\mathbb{P}^{m-s+1})\cong\mathbb{Z}

so that the criterion for the perversity given above is violated. ∎

5. Implications for the vanishing topology of smoothable IDS

5.1. Proof of Theorem 1.1

We consider a GL\mathrm{GL}-miniversal unfolding

𝐀:(p,0)×(k,0)(m×n,0)\mathbf{A}\colon(\mathbb{C}^{p},0)\times(\mathbb{C}^{k},0)\to(\mathbb{C}^{m\times n},0)

of the defining matrix AA for (XAs,0)(X_{A}^{s},0) on kk parameters t1,,tkt_{1},\dots,t_{k}. See, for instance, [FKZ21] for the underlying notion of GL\operatorname{GL}-equivalence and the existence and construction of such miniversal unfoldings. The induced deformation

XAs\textstyle{X_{A}^{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳𝐀s\textstyle{\mathcal{X}_{\mathbf{A}}^{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}{0}\textstyle{\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\textstyle{\mathbb{C}^{k}}

given by the projection of 𝒳𝐀s=𝐀1(Mm,ns)p×k\mathcal{X}_{\mathbf{A}}^{s}=\mathbf{A}^{-1}(M_{m,n}^{s})\subset\mathbb{C}^{p}\times\mathbb{C}^{k} to the parameter space k\mathbb{C}^{k} is versal in the sense that it covers all possible determinantal deformations of (XAs,0)(X_{A}^{s},0) coming from this particular choice of a determinantal structure.

Let the discriminant (Δ,0)(k,0)(\Delta,0)\subset(\mathbb{C}^{k},0) be the set of parameters t=(t1,,tk)t=(t_{1},\dots,t_{k}) with singular fibers X𝐀s(t)X_{\mathbf{A}}^{s}(t). For a suitable representative of the unfolding 𝐀\mathbf{A} we may choose a Milnor ball BεpB_{\varepsilon}\subset\mathbb{C}^{p} around the origin and a polydisc 0Dδk0\in D_{\delta}\subset\mathbb{C}^{k} in the parameter space such that

π:𝒳𝐀sBε(DδΔ)(DδΔ)\pi\colon\mathcal{X}_{\mathbf{A}}^{s}\cap B_{\varepsilon}\cap(D_{\delta}\setminus\Delta)\to(D_{\delta}\setminus\Delta)

is a smooth fiber bundle of manifolds with boundary with fiber MAsM_{A}^{s}, the smoothing of (XAs,0)(X_{A}^{s},0).

Since for tΔt\notin\Delta the map

At=𝐀(,t):Bεm×nA_{t}=\mathbf{A}(-,t)\colon B_{\varepsilon}\to\mathbb{C}^{m\times n}

is transverse to the rank stratification implies that AtA_{t} does not degenerate on the fiber X𝐀s(t)X_{\mathbf{A}}^{s}(t) in the sense that it has constant rank r=s1r=s-1 at all points xX𝐀s(t)x\in X_{\mathbf{A}}^{s}(t). Therefore the cokernel

nAtmE0\mathbb{C}^{n}\overset{A_{t}}{\longrightarrow}\mathbb{C}^{m}\to E\to 0

presented by AtA_{t} is a well defined vector bundle of rank mrm-r on the fiber over tt.

Variation of tt in DδΔD_{\delta}\setminus\Delta does not change the fiber X𝐀s(t)X_{\mathbf{A}}^{s}(t) up to diffeomorphism. In the same sense, it varies the vector bundle EE in a CC^{\infty} way on these fibers, but does not change its isomorphism class as a smooth complex vector bundle.

Due to the versality of 𝐀\mathbf{A}, the various deformations of (XAs,0)(X_{A}^{s},0) and its smoothing MAsM_{A}^{s} in the proof of the bouquet decomposition (1) in [Zac20] can be realized using piecewise smooth paths in the parameter space k\mathbb{C}^{k}. In particular, this leads to a representative X𝐀s(t)X_{\mathbf{A}}^{s}(t) of MAsM_{A}^{s} which has an embedding of the complex link mnp1(Mm,ns,0)X𝐀(t)\mathcal{L}^{mn-p-1}(M_{m,n}^{s},0)\hookrightarrow X_{\mathbf{A}}(t). Now it follows from the results in Section 4 that the truncated cohomology of the Grassmannian Hd(Grass(r,m))H^{\leq d}(\operatorname{Grass}(r,m)) is generated by the algebra of Segre classes of the tautological quotient bundle. Since by construction the pullback of this bundle to both X𝐀(t)X_{\mathbf{A}}(t) and mnp1(Mm,ns,0)\mathcal{L}^{mn-p-1}(M_{m,n}^{s},0) is the vector bundle EE in question, the result follows.

5.2. Proof of Corollary 1.2

Let MU5M\subset U\subset\mathbb{C}^{5} be a smoothing of the isolated Cohen-Macaulay threefold (X,0)5(X,0)\subset\mathbb{C}^{5} on some open set UU on which a perturbation

At:Um×(m+1)A_{t}\colon U\to\mathbb{C}^{m\times(m+1)}

of the matrix AA is defined. By definition the canonical sheaf on MM is given by ωM=xt𝒪U2(𝒪M,𝒪U)\omega_{M}=\mathcal{E}xt^{2}_{\mathcal{O}_{U}}(\mathcal{O}_{M},\mathcal{O}_{U}) which can be computed from a free resolution of 𝒪M\mathcal{O}_{M} as an 𝒪U\mathcal{O}_{U}-module. Such a free resolution is provided by the deformation of the resolution of 𝒪X,0\mathcal{O}_{X,0} from the Hilbert-Burch theorem. Now it is easy to see from the proof of Theorem 1.1 that the line bundle ωM\omega_{M} is presented by the restriction of AtA_{t} to MAsUM_{A}^{s}\subset U, i.e. it is the vector bundle EE in Theorem 1.1 and the claim follows.

Acknowlegdements

The author wishes to thank Duco van Straten for introducing him to the work of Borel on the cohomology of homogeneous spaces and Sam Hagh Shenas Noshari for further helpful conversations on the topic, Terence Gaffney for discussions on polar varieties and –multiplicities, Xiping Zhang for an exchange on their computation, and James Damon for conversations on the “characteristic cohomology” of determinantal singularities.

References

  • [BH93] W. Bruns and J. Herzog “Cohen-Macaulay rings” Cambridge University Press, 1993
  • [Bor53] A. Borel “Sur la cohomologie des espaces fibres principaux et des espace homogenes de groupes de Lie compacts” In Annals of Mathematics 57.1, 1953
  • [Bur68] L. Burch “On ideals of finite homological dimension in local rings” In Proc. Cambridge Philos. Soc. 64, 1968, pp. 941–948 DOI: 10.1017/s0305004100043620
  • [Dec+19] Wolfram Decker, Gert-Martin Greuel, Gerhard Pfister and Hans Schönemann Singular 4-1-2 — A computer algebra system for polynomial computations”, http://www.singular.uni-kl.de, 2019
  • [Dim04] Alexandru Dimca “Sheaves in topology”, Universitext Springer-Verlag, Berlin, 2004, pp. xvi+236 DOI: 10.1007/978-3-642-18868-8
  • [EGZ09] W. Ebeling and S. Gusein-Zade “On indices of 1-forms on determinantal singularities” 267, Proc. Steklov Inst. Math, 2009, pp. 113–124
  • [Eis95] D. Eisenbud “Commutative Algebra with a View Toward Algebraic Geometry” Springer-Verlag New York, 1995
  • [FKZ15] A. Frühbis-Krüger and M. Zach “On the Vanishing Topology of Isolated Cohen-Macaulay codimension 2 Singularities”, 2015 eprint: arXiv:1501.01915
  • [FKZ21] A. Frühbis-Krüger and M. Zach “Determinantal Singularities”, 2021 eprint: arXiv:2106.04855v1
  • [Ful98] William Fulton “Intersection theory” 2, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics] Springer-Verlag, Berlin, 1998, pp. xiv+470 DOI: 10.1007/978-1-4612-1700-8
  • [GM88] M. Goresky and R. MacPherson “Stratified Morse Theory” Springer Verlag, 1988
  • [Hat02] A. Hatcher “Algebraic Topology” Cambridge University Press, 2002
  • [Hil90] D. Hilbert “Ueber die Theorie der algebraischen Formen” In Math. Ann. 36.4, 1890, pp. 473–534 DOI: 10.1007/BF01208503
  • [HL91] H.A. Hamm and D.T. Lê “Rectified Homotopical Depth and Grothendieck Conjectures” In The Grothendieck Festschrift II Birkhäuser, 1991
  • [Hop41] H. Hopf “Über die Topologie der Gruppen-Mannigfaltigkeiten und ihrer Verallgemeinerungen” In Ann. of Math. 42, 1941, pp. 22–52
  • [Hus94] D. Husemoeller “Fiber Bundles, third edition” Springer, 1994
  • [Mil56] John Milnor “Construction of universal bundles. II” In Ann. of Math. (2) 63, 1956, pp. 430–436 DOI: 10.2307/1970012
  • [PnZ18] G. Peñafort and M. Zach “Kato-Matsumoto type results for disentanglements”, 2018 eprint: arXiv:1807.07183
  • [Sch77] M. Schaps “Deformations of Cohen-Macaulay schemes of codimension 2 and nonsingular deformations of space curves” In Amer. J. Math. 99, 1977, pp. 669 –684
  • [Szi19] Z. Szilágyi “On Chern classes of tensor products of vector bundles”, 2019 eprint: arXiv:1909.13278v1
  • [Tju68] G. N. Tjurina “Absolute isolatedness of rational singularities and triple rational points” In Func. Anal. Appl. 2, 1968, pp. 324–333
  • [TT81] Lê Dung Tràng and B. Teissier “Variétés polaires locales et classes de Chern des variétés singulières” In Ann. of Math. (2) 114.3, 1981, pp. 457–491 DOI: 10.2307/1971299
  • [Ura81] T. Urabe “Duality of Numerical Characters of Polar Loci” In Publ. RIMS Kyoto Univ., 1981, pp. 331–345
  • [Wah16] J. Wahl “Milnor and Tjurina numbers for smoothings of surface singularities”, 2016 eprint: arXiv:1307.6491v2
  • [Wah85] Jonathan M. Wahl “A characterization of quasihomogeneous Gorenstein surface singularities” In Compositio Math. 55.3, 1985, pp. 269–288 URL: http://www.numdam.org/item?id=CM_1985__55_3_269_0
  • [Zac17] M. Zach “Topology of isolated determinantal singularities”, 2017
  • [Zac18] M. Zach “An observation concerning the vanishing topology of certain isolated determinantal singularities” In Mathematische Zeitschrift, 2018 DOI: 10.1007/s00209-018-2143-9
  • [Zac20] M. Zach “Bouquet decomposition for Determinantal Milnor fibers” In Journal of Singularities 22, 2020, pp. 190 –204 DOI: 10.5427/jsing.2020.22m
  • [Zha17] X. Zhang “Local Euler Obstruction and Chern-Mather classes of Determinantal Varieties”, 2017 eprint: ArXiv:1706.02032