Skip to content
Licensed Unlicensed Requires Authentication Published by De Gruyter May 3, 2018

Probabilistic Schubert calculus

  • Peter Bürgisser and Antonio Lerario ORCID logo

Abstract

We initiate the study of average intersection theory in real Grassmannians. We define the expected degreeedegG(k,n) of the real Grassmannian G(k,n) as the average number of real k-planes meeting nontrivially k(n-k) random subspaces of n, all of dimension n-k, where these subspaces are sampled uniformly and independently from G(n-k,n). We express edegG(k,n) in terms of the volume of an invariant convex body in the tangent space to the Grassmannian, and prove that for fixed k2 and n,

edegG(k,n)=degG(k,n)12εk+o(1),

where degG(k,n) denotes the degree of the corresponding complex Grassmannian and εk is monotonically decreasing with limkεk=1. In the case of the Grassmannian of lines, we prove the finer asymptotic

edegG(2,n+1)=83π5/2n(π24)n(1+𝒪(n-1)).

The expected degree turns out to be the key quantity governing questions of the random enumerative geometry of flats. We associate with a semialgebraic set XPn-1 of dimension n-k-1 its Chow hypersurface Z(X)G(k,n), consisting of the k-planes A in n whose projectivization intersects X. Denoting N:=k(n-k), we show that

𝔼#(g1Z(X1)gNZ(XN))=edegG(k,n)i=1N|Xi||Pm|,

where each Xi is of dimension m=n-k-1, the expectation is taken with respect to independent uniformly distributed g1,,gmO(n) and |Xi| denotes the m-dimensional volume of Xi.

Award Identifier / Grant number: BU 1371/2-2

Funding statement: The first author was partially supported by DFG grant BU 1371/2-2.

A Appendix

A.1 Proof of Theorem 3.2

Since the density of A is invariant under the orthogonal group O(n), we can assume that B is spanned by the first l standard vectors, i.e.,

B=[𝟏𝟎]andA=[A1A2],

where 1 is the × identity matrix, A1 is an ×k matrix and A2 is an (n-)×k matrix. (Again, abusing notation, we denote by A and B also matrices whose columns span the corresponding spaces.) If we sample A with i.i.d. normal Gaussians, the corresponding probability distribution for the span of its columns is O(n) invariant, and consequently it coincides with the uniform distribution. In order to compute the principal angles between A and B using (3.1), we need to orthonormalize the columns of A. Defining

A^:=[A1A2](A1TA1+A2TA2)-12

we see that the span of the columns of A and A^ is the same, and the columns of A^ are orthonormal. The cosines of the principal angles between A and B are the singular values 1σ1σk0 of the matrix

A^TB=(A1TA1+A2TA2)-12A1T.

The σ1,,σk coincide with the square roots of the eigenvalues 1u1uk0 of the positive semidefinite matrix

(A1TA1+A2TA2)-12A1TA1(A1TA1+A2TA2)-12,

which are the same as the eigenvalues of N=(A1TA1+A2TA2)-1A1TA1 (eigenvalues are invariant under cyclic permutations). Consider the Cholesky decomposition

MTM=A1TA1+A2TA2.

Then the eigenvalues of N equal the eigenvalues of

U=(MT)-1A1TA1M-1.

We use now some facts about the multivariate Beta distribution (see [37, Section 3.3]). By its definition, the matrix U has a Betak(12l,12(n-l)) distribution, and [37, Theorem 3.3.4] states that the joint density of the eigenvalues of U is given by

(A.1)πk22Γk(n2)Γk(k2)Γk(l2)Γk(n-l2)j=1kujl-k-12(1-uj)n-l-k-12i<j(ui-uj).

Recall now that ui=(cosθi)2 for i=1,,k, which implies the change of variable

dui=-2cosθisinθidθi,

and thus (A.1) becomes the stated density in (3.2).

A.2 Proof of Lemma 4.3

We begin with a general reasoning. Assume that Ae(B). The unit normal vectors of e(B) at A, up to a sign, are uniquely determined by A. Lemma 4.1 provides an explicit description for them as follows. Let a1,a2,,ak and a1,b2,,bn-k be the orthonormal bases given by Lemma 3.1 for A and B, respectively; note that dim(AB)=1 since Ae(B). In particular, ai,bi=cosθi for ik, where θ1θk are the principal angles between A and B. Let f=(A+B) with f=1. According to Lemma 4.1, the unit vector ν:=fa2ak spans the normal space of e(B) at A. (We use here the representation of elements of G(k,n) and its tangent spaces by vectors in Λkn; cf. Section 2.1.)

Fix now BG(n-k,n) and recall that the functions θ1,θ2:G(k,n) give the smallest and second smallest principal angle, respectively, between AG(k,n) and B. Let 0<εδ<π2 and put

T:={AG(k,n)θ1(A)ε,θ2(A)δ}.

We first prove that 𝒯(e(B)δ,ε)T.

For Ae(B) we consider the curve NA(t):=(a1cost+fsint)a2ak for t. We note that NA(t) arises from A by a rotation with the angle t in the (oriented) plane spanned by a1 and f, fixing the vectors in the orthogonal complement spanned by a2,,ak. We have NA(0)=A and N˙A(0)=ν. It is well known (see, e.g., [13]) that NA(t) is the geodesic through A with speed vector ν, that is, NA(t)=expA(tν) and NA(0)=A. From the definition of the ε-tube, we therefore have

(A.2)𝒯(e(B)δ,ε)={NA(t)Ae(B)δ,|t|ε}.

Since both a1 and f are orthogonal to b2,,bn-k, the principal angles between NA(t) and B are |t|,θ2,,θk, where aj,bj=cosθj. By our assumption Ae(B)δ, we have δθ2θk. Hence we see that |t| is the smallest principle angle between NA(t) and B if |t|δ. We have thus verified that 𝒯(e(B)δ,ε)T.

For the other inclusion, let CT; assume that Ce(B)δ, otherwise we clearly have C𝒯(e(B)δ,ε). Let (c1,,ck) and (b1,,bn-k) be orthonormal bases of C and B, respectively, as provided by Lemma 3.1. So we have bj,cj=cosθj(C) for all j, where θ1(C)ε and δθ2(C)θk(C) by the assumption CT. In particular, b1 is orthogonal to c2,,ck and b1,c1 are linearly independent. We define AG(k,n) as the space spanned by b1,c2,,ck. By construction, θ1(A)=0 and θj(A)=θj(C) for j2. Hence, Ae(B)δ. Let NA(t)G(k,n) be defined as above as the space resulting from A by a rotation with the angle t in the oriented plane spanned by b1 and c1. By construction, we have C=NA(τ), where τ=θ1(C)ε. Therefore, we indeed have C𝒯(e(B)δ,ε) by (A.2).

A.3 Proof of Lemma 4.5

More generally, we consider a semialgebraic set YPn-1 of dimension dn-k-1. Consider the semialgebraic set

C(Y):={(A,y)G(k,n)×Yy(A)},

together with the projections on the two factors: π1:C(Y)G(k,n) and π2:B(Y)Y. As Z(Y)=π1(C(Y)), the set Z(Y) is semialgebraic. Note that C(Y) is compact if Y compact, and C(Y) is connected if Y is connected (since π1 is continuous). In order to determine the dimension of C(Y), we note that the fiber π2-1(y) over yY is isomorphic to G(k-1,n-1). As a consequence, we get

(A.3)dimC(Y)=dimY+(k-1)(n-k).

The fibers π1-1(A) over AZ(Y) consist of exactly one point, except for the A lying in the exceptional set

Z(2)(Y):={AZ(Y)(A)Y consists of at least two points}.

Note that (A)Y consists of one point only, for all AZ(Y)Z(2)(Y).

In order to show that dimZ(2)(Y)<dimZ(Y), we consider the semialgebraic set

C(2)(Y):={(A,y1,y2)G(k,n)×Y×Yy1,y2(A),y1y2}

with the corresponding projections π3:C(2)(Y)G(k,n) and π4:C(2)(Y)Y×Y. Note that Z(2)(Y)=π3(C(2)(Y)). For (y1,y2)Y×Y such that y1y2, we have

dimπ4-1(y1,y2)=k(n-k)-2(n-k),

since the fibers of π4 are isomorphic to G(k-2,n-2). Therefore,

(A.4)dimC(2)(Y)=2dimY+k(n-k)-2(n-k)
dimY+n-k-1+k(n-k)-2(n-k)
=dimY+(k-1)(n-k)-1=(A.3)dimC(Y)-1,

and we see that dim(C(Y)C(2)(Y))=dimC(Y). For the projection f:C(2)(Y)C(Y) defined by f(A,y1,y2):=(A,y1), we have π1-1(Z(2)(Y))=f(C(2)(Y)), hence

dimπ1-1(Z(2)(Y))dimf(C(2)(Y))dimC(2)(Y)<dimC(Y).

Moreover, the projection C(Y)C(2)(Y)Z(Y)Z(2)(Y), (A,y)y is bijective, hence

dim(Z(Y)Z(2)(Y))=dim(C(Y)C(2)(Y))=(A.4)dimC(Y).

Using dimZ(Y)dimC(Y), we see that

(A.5)dimZ(Y)=dimC(Y)=dimY+(k-1)(n-k)

and

(A.6)dimZ(2)(Y)dimC(2)(Y)<dimC(Y)=dimZ(Y).

In the special case where dimX=n-k-1, we conclude that Z(X) is a hypersurface.

We consider now the following set of “bad” AZ(X):

S(X):=Z(2)(X)Z(Sing(X))Sing(Z(X))π1(Sing(C(X))).

We claim that this semialgebraic set has dimension strictly less than Z(X). This follows for Z(2)(X) from (A.6), for Z(SingX) from (A.5) applied to Y=Sing(X), for π1(Sing(C(X)) from Proposition 2.2 and (A.5), and finally for Sing(Z(X)) by Proposition 2.2.

Thus generic points AZ(X) are not in S(X) and hence satisfy the following:

  1. The intersection (A)X consists of one point only (let us denote this point by p),

  2. The point p is a smooth point of X,

  3. The point A is a regular point of Z(X),

  4. The point (A,p) is a regular point of C(X).

It remains to prove that for every AZ(X)S(X) we have (4.3). To this end, let us take (A,p)C(X) with AS(X). We work in local coordinates (w,y)N×n-1 on a neighborhood U of (A,p) in G(k,n)×Pn-1, where N=k(n-k). For simplicity we center the coordinates on the origin, so that (0,0) are the coordinates of (A,p). In this coordinates, the set C(X)U can be described as

(A.7)C(X)U={(w,x)N×n-1F(w,x)=0,G(x)=0},

where F(w,x)=0 represents the reduced local equations describing the condition x(W) and G(x)=0 the reduced local equations giving the condition xX. Since (A,p) is a regular point of C(X), the tangent space of C(X) at (A,p) is described by

T(A,p)C(X)={(w˙,x˙)(D(0,0)F)(w˙,x˙)=0,(D0G)x˙=0}.

Note that p is a smooth point of X, hence (D0G)x˙=0 is the equation for the tangent space to X at p. On the other hand, since Z(X)=π1(C(X)), we have

(A.8)TAZ(X)=D(0,0)π1(T(A,p)C(X)).

Let now Bn be the linear space corresponding to TpX and let us write the equations for C(B) in the same coordinates as above (by construction we have (A,p)C(B)):

C(B)U={(w,x)F(w,x)=0,(D0G)x=0}.

Note that the same equation F(w,x)=0 as in (A.7) appears here (recall that this is the equation describing x(W)), but now G=0 is replaced with its linearization at zero. In particular,

T(A,p)C(B)={(w˙,x˙)(D(0,0)F)(w˙,x˙)=0,(D0G)x˙=0},

which coincides with (A.8). Since Ω(B)=π1(C(B)), this finally implies

TAΩ(B)=D(0,0)π1(T(A,p)C(B))=D(0,0)π1(T(A,p)C(X))=TAZ(X),

which finishes the proof.

A.4 Proof of Proposition 5.12

We extend the function f to k×m{0} by setting

f(X):=f(XX),

denoting it by the same symbol. Similarly, we extend g by setting

g(σ):=f(diagk,m(σ)).

We assume now that Xk×m has i.i.d. standard Gaussian entries. Then we can write

1|Skm-1|Skm-1f(X)𝑑Skm-1=𝔼f(X)
=σ1>>skg(σ)pSVD(σ1,,σk)𝑑σ1𝑑σk,

where pSVD denotes the joint density of the ordered singular values of X. This density can be derived as follows. The joint density of the ordered eigenvalues λ1>>λk>0 of the Wishart distributed matrix XXT is known to be [37, Corollary 3.2.19]

p(λ1,,λk)=ck,me-12i=1kλii=1kλim-k-12i<j(λi-λj),

where

ck,m:=2kπk222km2Γk(k2)Γk(m2).

From this, using λi(XXT)=σi(X)2 and the change of variable dλi=2σidσi, we obtain

pSVD(σ1,,σk)=ck,me-12i=1kσi2i=1kσim-ki<j(σi2-σj2).

As a consequence, we obtain

𝔼f(X)=ck,mσ1>>σk>0g(σ)e-12i=1kσi2i=1kσim-ki<j(σi2-σj2)dσ1dσk
=ck,m0rkm-1e-12r2𝑑rθS+k-1g(θ)i=1kθim-ki<j(θi2-θj2)dSk-1(θ),

where in the second line we have switched to polar coordinates σ=rθ with θSk-1 and r0. Note that the power of the r-variable arises as

k(m-k)+2(k2)+k-1=km-1.

Using

0rkm-1e-12r2𝑑r=Γ(km2)2km2-1,

we obtain

𝔼f(X)=ck,mΓ(km2)2km2-1S+k-1g(θ)i=1kθim-k1i<jk(θi2-θj2)dSk-1.

It is immediate to verify that

|Skm-1|ck,mΓ(km2)2km2-1=|O(k)||S(k,m)|2k,

which completes the proof.

A.5 Generalized Poincaré’s formula in homogeneous spaces

The purpose of this subsection is to prove the kinematic formula in homogeneous spaces for multiple intersections and to derive Theorem 3.19. The proofs are similar to [26], to which we refer the reader for more details.

A.5.1 Definitions and statement of the theorem

In the following, G denotes a compact Lie group with a left and right invariant Riemannian metric.[2] See [6] for background on Lie groups. We denote by eG the identity element and by Lg:GG,xgx the left translation by gG. The derivatives of Lg will be denoted by g*:TeGTgG. By assumption, this map is isometric.

In (3.5) we defined a quantity for capturing the relative position of linear subspaces of a Euclidean vector space. We can extend this notion to linear subspaces ViTgiG in tangent spaces of G at any points g1,,gmG, assuming idimVidimG. This is done by left-translating the gi to the identity e: so we define

σ(V1,,Vm):=σ((g1)*-1V1,,(gm)*-1Vm).

Let now KG be a closed Lie subgroup and denote by p:GG/K the quotient map. We endow K with the Riemannian structure induced by its inclusion in G, and G/K with the Riemannian structure defined by declaring p to be a Riemannian submersion. For example, when G=O(n) with the invariant metric defined in Section 2.8 and K=O(k)×O(n-k), then G/K with the quotient metric is isometric to the Grassmannian G(k,n) with the metric defined in Section 2.1.

Note that G acts naturally by isometries on G/K; if gG and yG/K, we denote by gy the result of the action. Further, we denote by y0=p(e) the projection of the identity element. The multiplication with an element kK fixes the point y0; as a consequence, the differential of k induces a map denoted k*:Ty0G/KTy0G/K, so that we have an induced action of K on Ty0G/K.

Given a submanifold X of a Riemannian manifold M, we denote by NX its normal bundle in M (i.e., for all xX the vector space NxX is the orthogonal complement to TxX in TxM). Also, the restriction of the Riemannian metric of M to X allows to define a volume density on X; if f:X is an integrable function, we denote its integral with respect to this density by Xf(x)𝑑x.

Definition A.1.

For given submanifolds Y1,,YmG/K, we define the function

σK:Y1××Ym

as follows. For (y1,,ym)Y1××Ym let ξiG be such that ξiyi=y0 for all i. We define

σK(y1,,ym):=𝔼(k1,,km)Kmσ(k1*ξ1*Ny1Y1,,km*ξm*NymYm),

where the expectation is taken over a uniform (k1,,km)K××K.

The reader should compare this definition with [26, Definition 3.3], which is just a special case. The main result of this section is the following generalization of Poincaré’s kinematic formula for homogeneous spaces, as stated in [26, Theorem 3.8] for the intersection of two manifolds. We provide a proof, since the more general result is crucial for our work and we were unable to find it in the literature.

Theorem A.2.

Let Y1,,Ym be submanifolds of G/K such that

i=1mcodimG/KYidimG/K.

Then, for almost all (g1,,gm)Gm, the manifolds g1Y1,,gmYm intersect transversely, and

𝔼(g1,,gm)Gm|g1Y1gmYm|
  =1|G/K|m-1Y1××YmσK(y1,,ym)𝑑y1𝑑ym,

where the expectation is taken over a uniform (g1,,gm)G××G.

We note that when G acts transitively on the tangent spaces to each Yi (formally defined as in Definition 3.4), then the function σK:Y1××Ym introduced in Definition A.1 is constant. As a consequence we obtain:

Corollary A.3.

Under the assumptions of Theorem A.2, if moreover G acts transitively on the tangent spaces to Yi for i=1,,m, then we have

𝔼(g1,,gm)Gm|g1Y1gmYm|=σK(y1,,ym)|G/K|i=1n|Yi||G/K|,

where (y1,,ym) is any point of Y1××Ym.

Let us look now at the special case G=O(n) and K=O(k)×O(n-k). If Y1,,Ym are coisotropic hypersurfaces of G(k,n), then G acts transitively on their tangent spaces by Proposition 4.6. Moreover, when m=k(n-k), it is easy to check that the constant value of σK equals the real average scaling factor α(k,n-k) defined in Definition 3.18. Hence in this case, the statement of Corollary A.3 coincides with the statement of Theorem 3.19.

A.5.2 The kinematic formula in G

As before, G denotes a compact Lie group. We derive first Theorem A.2 in the special case K={e}, which is the following result (we can without loss of generality assume gm=e).

Lemma A.4.

Let X1,,Xm be submanifolds of G such that

i=1mcodimGXidimG.

Then

Gm-1|g1X1gm-1Xm-1Xm|𝑑g1𝑑gm-1
  =X1××Xmσ(Nx1X1,,NxmXm)𝑑x1𝑑xm.

In the special case of intersecting two submanifolds, this is an immediate consequence of the following “basic integral formula” from [26, Section 2.7] (take h=1).

Proposition A.5.

Let M1,M2 be submanifolds of G such that

codimGM1+codimGM2dimG.

For almost all gG, the manifolds M1 and gM2 intersect transversely, and if h is an integrable function on M1×M2, then

GgM1M2h(φg(y))𝑑y𝑑g=M1×M2h(x1,x2)σ(Nx1M1,Nx2M2)𝑑x1𝑑x2,

where φg:gM1M2M1×M2 is the function given by φg(y):=(g-1y,y).

In order to reduce the general case to that of intersecting two submanifolds, we first establish a linear algebra identity.

Lemma A.6.

For subspaces V1,,V,W,Z of a Euclidean vector space, we have

σ(V1,,V,W,Z)=σ(V1,,V,W+Z)σ(W,Z).

Proof.

We may assume that WZ=0, since otherwise both sides of the identity are zero. Let us denote by (vi,1,,vi,di), (w1,,wa) and (z1,,zb) orthonormal bases of Vi, W and Z, respectively. Moreover, we denote by (w1,,wa,z~1,,z~b) an orthonormal basis for Z+W obtained by completing (w1,,wa). By definition we have

(A.9)σ(V1,,V,W+Z)=(i,jvi,j)w1waz~1z~b.

On the other hand, since W+Z=span{w1,,wa,z1,,zb}, we have

w1waz~1z~b=±w1waz1zbw1waz1zb
=±w1waz1zbσ(W,Z).

By substituting the last line into (A.9) and recalling the definition of σ(V1,,V,W,Z) from equation (3.5), the assertion follows. ∎

Corollary A.7.

Let M1,M2 be submanifolds of G and let V1,,V be linear subspaces of tangent spaces of G (possibly at different points) such that

codimGM1+codimGM2+idimVidimG.

Then we have

GgM1M2σ(V1,,V,Ny(gM1M2))𝑑y𝑑g
  =M1×M2σ(V1,,V,Nx1M1,Nx2M2)𝑑x1𝑑x2.

Proof.

We apply Proposition A.5 with the function h:M1×M2 defined by

h(x1,x2):=σ(V1,,V,(x1)*-1Nx1M1+(x2)*-1Nx2M2).

When gM1 and M2 intersect transversely, we have Ny(gM1M2)=Ny(gM1)+Ny(M2), which implies h(φg(y))=σ(V1,,V,Ny(gM1gM2)) for ygM1M2. Hence we obtain with Proposition A.5,

GgM1M2σ(V1,,V,Ny(gM1M2))𝑑y𝑑g
  =G(gM1M2h(φg(y))𝑑y)𝑑g
  =M1×M2h(x1,x2)σ(Nx1M1,Nx2M2)𝑑x1𝑑x2
  =M1×M2σ(V1,,V,Nx1M1,Nx2M2)𝑑x1𝑑x2,

where the last equality is due to Lemma A.6. ∎

Proof of Lemma A.4.

Recall that we already established Lemma A.4 in the case m=2 as a consequence of Proposition A.5. Let now m3 and abbreviate Y:=g2X2Z, where Z:=g3X3gm-1Xm-1Xm. Then we have

Gm-2(g1G|g1X1(g2X2gm-1Xm-1Xm)|𝑑g1)𝑑g2𝑑gm-1
  =Gm-2(X1Yσ(Nx1X1,NyY)𝑑x1𝑑y)𝑑g2𝑑gm-1
  =X1Gm-3(g2GYσ(Nx1X1,Ny(g2X2Z)dg2dy)dg3dgm-1dx1
  =X1Gm-3(x2X2zZσ(Nx1X1,Nx2X2,NzZ)𝑑g2)𝑑g3𝑑gm-1,

where we first applied Lemma A.4 in the case m=2, then interchanged the order of integration, and after that used Corollary A.7. Proceeding analogously, we see that the above integral indeed equals

X1××Xmσ(Nx1X1,,NxmXm)𝑑x1𝑑xm,

which completes the proof. ∎

Proof of Theorem A.2.

We consider Xi:=p-1(Yi), which is a submanifold, since the projection p:GG/K is a submersion, cf. [9, Theorem A.15]. Moreover, g1X1,,gmXm intersect transversally if Y1,,Ym do so. We can rewrite the integral in the statement as

E:=𝔼(g1,,gm)Gm|g1Y1gmYm|
=𝔼(g1,,gm-1)Gm-1|g1Y1gm-1Ym-1Ym|
=1|K|1|G|m-1Gm-1|g1X1gm-1Xm-1Xm|𝑑g1𝑑gm-1,

For justifying the last equality, note that the coarea formula [9, Theorem 17.8] yields

|p-1(Y)|=|K||Y|

for any submanifold Y of G/K, since p is a Riemannian submersion. Applying Lemma A.4 to the integral in the last line, we obtain

E=1|K|1|G|m-1X1××Xmσ(Nx1X1,,NxmXm)𝑑x1𝑑xm.

The projection P:X1××XmY1××Ym defined by

p(x1,,xm):=(p(x1),,p(xm))

is a Riemannian submersion with fibers isometric to Km. Using the coarea formula

X1××Xmσ(Nx1X1,,NxmXm)𝑑x1𝑑xm
  =yY1×YmxP-1(y)σ(Nx1X1,,NxmXm)𝑑x𝑑y
  =|K|myY1×YmσK(y1,,ym)𝑑y,

where the last equality is due to Definition A.1. (Note that here is where we use the right invariance of the metric under the action of the elements in K, as one can verify by a careful inspection of the last steps. The reader can see the proof of [26, Theorem 3.8], which is almost identical and where all the details of the calculation are shown). This completes the proof. ∎

Acknowledgements

We are very grateful to Frank Sottile who originally suggested this line of research. We thank Paul Breiding, Kathlén Kohn, Chris Peterson, and Bernd Sturmfels for discussions. We also thank the anonymous referees for their comments.

References

[1] P.-A. Absil, A. Edelman and P. Koev, On the largest principal angle between random subspaces, Linear Algebra Appl. 414 (2006), no. 1, 288–294. 10.1016/j.laa.2005.10.004Search in Google Scholar

[2] D. Amelunxen, Geometric analysis of the condition of the convex feasibility problem, PhD thesis, University of Paderborn, Paderborn 2011. Search in Google Scholar

[3] S. Basu, A. Lerario, E. Lundberg and C. Peterson, Random fields and the enumerative geometry of lines on real and complex hypersurfaces, preprint (2016), https://arxiv.org/abs/1610.01205. 10.1007/s00208-019-01837-0Search in Google Scholar

[4] J. Bochnak, M. Coste and M.-F. Roy, Real algebraic geometry, Ergeb. Math. Grenzgeb. (3) 36, Springer, Berlin 1998. 10.1007/978-3-662-03718-8Search in Google Scholar

[5] P. Breiding, The expected number of eigenvalues of a real Gaussian tensor, SIAM J. Appl. Algebra Geom. 1 (2017), no. 1, 254–271. 10.1137/16M1089769Search in Google Scholar

[6] T. Bröcker and T. Tom Dieck, Representations of compact Lie groups, Grad. Texts in Math. 98, Springer, New York 1995. Search in Google Scholar

[7] P. Bürgisser, Average Euler characteristic of random real algebraic varieties, C. R. Math. Acad. Sci. Paris 345 (2007), no. 9, 507–512. 10.1016/j.crma.2007.10.013Search in Google Scholar

[8] P. Bürgisser, Condition of intersecting a projective variety with a varying linear subspace, SIAM J. Appl. Algebra Geom. 1 (2017), no. 1, 111–125. 10.1137/16M1077222Search in Google Scholar

[9] P. Bürgisser and F. Cucker, Condition. The geometry of numerical algorithms, Grundlehren Math. Wiss. 349, Springer, Heidelberg 2013. 10.1007/978-3-642-38896-5Search in Google Scholar

[10] A. Dimca, Topics on real and complex singularities. An introduction, Adv. Lect. Math., Friedrich Vieweg & Sohn, Braunschweig 1987. 10.1007/978-3-663-13903-4Search in Google Scholar

[11] J. Draisma and E. Horobeţ, The average number of critical rank-one approximations to a tensor, Linear Multilinear Algebra 64 (2016), no. 12, 2498–2518. 10.1080/03081087.2016.1164660Search in Google Scholar

[12] J. Draisma, E. Horobeţ, G. Ottaviani, B. Sturmfels and R. R. Thomas, The Euclidean distance degree of an algebraic variety, Found. Comput. Math. 16 (2016), no. 1, 99–149. 10.1007/s10208-014-9240-xSearch in Google Scholar

[13] A. Edelman, T. A. Arias and S. T. Smith, The geometry of algorithms with orthogonality constraints, SIAM J. Matrix Anal. Appl. 20 (1999), no. 2, 303–353. 10.1137/S0895479895290954Search in Google Scholar

[14] D. Eisenbud and J. Harris, 3264 and all that—a second course in algebraic geometry, Cambridge University Press, Cambridge 2016. 10.1017/CBO9781139062046Search in Google Scholar

[15] S. Finashin and V. Kharlamov, Abundance of 3-planes on real projective hypersurfaces, Arnold Math. J. 1 (2015), no. 2, 171–199. 10.1007/s40598-015-0015-5Search in Google Scholar

[16] W. Fulks and J. O. Sather, Asymptotics. II. Laplace’s method for multiple integrals, Pacific J. Math. 11 (1961), 185–192. 10.2140/pjm.1961.11.185Search in Google Scholar

[17] W. Fulton, Intersection theory, 2nd ed., Ergeb. Math. Grenzgeb. (3) 2, Springer, Berlin 1998. 10.1007/978-1-4612-1700-8Search in Google Scholar

[18] Y. V. Fyodorov, A. Lerario and E. Lundberg, On the number of connected components of random algebraic hypersurfaces, J. Geom. Phys. 95 (2015), 1–20. 10.1016/j.geomphys.2015.04.006Search in Google Scholar

[19] D. Gayet and J.-Y. Welschinger, Lower estimates for the expected Betti numbers of random real hypersurfaces, J. Lond. Math. Soc. (2) 90 (2014), no. 1, 105–120. 10.1112/jlms/jdu018Search in Google Scholar

[20] D. Gayet and J.-Y. Welschinger, Expected topology of random real algebraic submanifolds, J. Inst. Math. Jussieu 14 (2015), no. 4, 673–702. 10.1017/S1474748014000115Search in Google Scholar

[21] D. Gayet and J.-Y. Welschinger, Betti numbers of random real hypersurfaces and determinants of random symmetric matrices, J. Eur. Math. Soc. (JEMS) 18 (2016), no. 4, 733–772. 10.4171/JEMS/601Search in Google Scholar

[22] I. M. Gelfand, M. M. Kapranov and A. V. Zelevinsky, Discriminants, resultants and multidimensional determinants, Mod. Birkhäuser Class., Birkhäuser, Boston 2008. Search in Google Scholar

[23] A. Gray and L. A. Vanhecke, The volumes of tubes in a Riemannian manifold, Rend. Sem. Mat. Univ. Politec. Torino 39 (1981), no. 3, 1–50. Search in Google Scholar

[24] P. Griffiths and J. Harris, Principles of algebraic geometry, Wiley Class. Libr., John Wiley & Sons, New York 1994. 10.1002/9781118032527Search in Google Scholar

[25] J. Harris, Algebraic geometry, Grad. Texts in Math. 133, Springer, New York 1995. Search in Google Scholar

[26] R. Howard, The kinematic formula in Riemannian homogeneous spaces, Mem. Amer. Math. Soc. 106 (1993), no. 509, 1–69. 10.1090/memo/0509Search in Google Scholar

[27] S. L. Kleiman, The transversality of a general translate, Compos. Math. 28 (1974), 287–297. Search in Google Scholar

[28] S. L. Kleiman and D. Laksov, Schubert calculus, Amer. Math. Monthly 79 (1972), 1061–1082. 10.1080/00029890.1972.11993188Search in Google Scholar

[29] K. Kohn, Coisotropic hypersurfaces in the Grassmannian, preprint (2016), https://arxiv.org/abs/1607.05932. 10.1016/j.jsc.2019.12.002Search in Google Scholar

[30] E. Kostlan, On the distribution of roots of random polynomials, From topology to computation: Proceedings of the Smalefest (Berkeley 1990), Springer, New York (1993), 419–431. 10.1007/978-1-4612-2740-3_38Search in Google Scholar

[31] S. E. Kozlov, Geometry of real Grassmannian manifolds. I, II, III, Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) 246 (1997), 84–107, 108–129, 197–198. Search in Google Scholar

[32] A. Lerario, Random matrices and the average topology of the intersection of two quadrics, Proc. Amer. Math. Soc. 143 (2015), no. 8, 3239–3251. 10.1090/proc/12324Search in Google Scholar

[33] A. Lerario and E. Lundberg, Gap probabilities and Betti numbers of a random intersection of quadrics, Discrete Comput. Geom. 55 (2016), no. 2, 462–496. 10.1007/s00454-015-9741-7Search in Google Scholar

[34] S. Lojasiewicz, Ensembles semi-analytiques, Institut des Hautes Études Scientifiques, Bures-sur-Yvette 1965. Search in Google Scholar

[35] L. Manivel, Symmetric functions, Schubert polynomials and degeneracy loci, SMF/AMS Texts Monogr. 6, American Mathematical Society, Providence 2001. Search in Google Scholar

[36] J. W. Milnor and J. D. Stasheff, Characteristic classes, Ann. of Math. Stud. 76, Princeton University Press, Princeton 1974. 10.1515/9781400881826Search in Google Scholar

[37] R. J. Muirhead, Aspects of multivariate statistical theory, Wiley Ser. Probab. Math. Stat., John Wiley & Sons, New York 1982. 10.1002/9780470316559Search in Google Scholar

[38] D. Mumford, Algebraic geometry. I. Complex Projective Varieties, Grundlehren Math. Wiss. 221, Springer, Berlin 1976. Search in Google Scholar

[39] I. Rivin, Surface area and other measures of ellipsoids, Adv. in Appl. Math. 39 (2007), no. 4, 409–427. 10.1016/j.aam.2006.08.009Search in Google Scholar

[40] L. A. Santaló, Integral geometry and geometric probability, Encyclopedia Math. Appl., Addison–Wesley, Reading 1976. Search in Google Scholar

[41] R. Schneider, Convex bodies: The Brunn–Minkowski theory, expanded ed., Encyclopedia Math. Appl. 151, Cambridge University Press, Cambridge 2014. Search in Google Scholar

[42] H. Schubert, Anzahl-Bestimmungen für lineare Räume, Acta Math. 8 (1886), no. 1, 97–118. 10.1007/BF02417085Search in Google Scholar

[43] I. R. Shafarevich, Basic algebraic geometry. 2, 3rd ed., Springer, Heidelberg 2013. 10.1007/978-3-642-38010-5_1Search in Google Scholar

[44] M. Shub and S. Smale, Complexity of Bezout’s theorem. II. Volumes and probabilities, Computational algebraic geometry (Nice 1992), Progr. Math. 109, Birkhäuser, Boston (1993), 267–285. 10.1007/978-1-4612-2752-6_19Search in Google Scholar

[45] F. Sottile, Pieri’s formula via explicit rational equivalence, Canad. J. Math. 49 (1997), no. 6, 1281–1298. 10.4153/CJM-1997-063-7Search in Google Scholar

[46] F. Sottile, The special Schubert calculus is real, Electron. Res. Announc. Amer. Math. Soc. 5 (1999), 35–39. 10.1090/S1079-6762-99-00058-XSearch in Google Scholar

[47] F. Sottile, Enumerative real algebraic geometry, Algorithmic and quantitative real algebraic geometry (Piscataway 2001), DIMACS Ser. Discrete Math. Theoret. Comput. Sci. 60, American Mathematical Society, Providence (2003), 139–179. 10.1090/dimacs/060/11Search in Google Scholar

[48] F. Sottile, Real solutions to equations from geometry, Univ. Lecture Ser. 57, American Mathematical Society, Providence 2011. 10.1090/ulect/057Search in Google Scholar

[49] M. Spivak, A comprehensive introduction to differential geometry. Vol. III, 2nd ed., Publish or Perish, Wilmington 1979. Search in Google Scholar

[50] J. Stasica, Smooth points of a semialgebraic set, Ann. Polon. Math. 82 (2003), no. 2, 149–153. 10.4064/ap82-2-5Search in Google Scholar

[51] B. Sturmfels, The Hurwitz form of a projective variety, J. Symbolic Comput. 79 (2017), 186–196. 10.1016/j.jsc.2016.08.012Search in Google Scholar

[52] T. Tao, Topics in random matrix theory, Grad. Stud. Math. 132, American Mathematical Society, Providence 2012. 10.1090/gsm/132Search in Google Scholar

[53] R. C. Thompson, Singular values, diagonal elements, and convexity, SIAM J. Appl. Math. 32 (1977), no. 1, 39–63. 10.1137/0132003Search in Google Scholar

[54] R. Vakil, Schubert induction, Ann. of Math. (2) 164 (2006), no. 2, 489–512. 10.4007/annals.2006.164.489Search in Google Scholar

[55] R. A. Vitale, Expected absolute random determinants and zonoids, Ann. Appl. Probab. 1 (1991), no. 2, 293–300. 10.1214/aoap/1177005938Search in Google Scholar

[56] R. Wong, Asymptotic approximations of integrals, Class. Appl. Math. 34, Society for Industrial and Applied Mathematics (SIAM), Philadelphia 2001. 10.1137/1.9780898719260Search in Google Scholar

[57] Y.-C. Wong, Conjugate loci in Grassmann manifolds, Bull. Amer. Math. Soc. 74 (1968), 240–245. 10.1090/S0002-9904-1968-11903-2Search in Google Scholar

Received: 2017-04-23
Revised: 2018-01-13
Published Online: 2018-05-03
Published in Print: 2020-03-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/crelle-2018-0009/html
Scroll to top button