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

On the 3232-dimensional Rosenfeld projective plane

John Jones, Dmitriy Rumynin, Adam Thomas
Abstract.

Following on from [JRT23], we make a detailed study of the 3232-dimensional Rosenfeld projective plane which is the symmetric space EIII in Cartan’s list of compact symmetric spaces.

Introduction

In [JRT23] we made a systematic study of the classical topological invariants of homogeneous spaces with a particular emphasis on the twelve compact symmetric spaces for the exceptional Lie groups. Our main examples in that paper are the three Rosenfeld projective planes of dimensions 32,6432,64 and 128128. In this paper we make a much more detailed study of the 3232-dimensional Rosenfeld projective plane.

Throughout we will use the notation RR for the 3232-dimensional Rosenfeld projective plane. Explicitly,

R=E6Spin(10)×C4S1,R=\frac{E_{6}}{\mathrm{Spin}(10)\times_{C_{4}}S^{1}},

which is a symmetric space for the exceptional Lie group E6E_{6}. It is called R5R_{5} in [JRT23], EIII in Cartan’s list of compact symmetric spaces, and also known as P2(𝕆)P^{2}(\mathbb{O}\otimes\mathbb{C}). The subgroup C4C_{4} is the cyclic subgroup of Spin(10)×S1\mathrm{Spin}(10)\times S^{1} generated by (ϵ,i)Spin(10)×S1(\epsilon,i)\in\mathrm{Spin}(10)\times S^{1}. Here ϵ\epsilon is the element of the centre of Spin(10)\mathrm{Spin}(10) which acts as multiplication by ii on δ10+\delta_{10}^{+} and, therefore, multiplication by i-i on δ10\delta_{10}^{-}. As usual, δ10±\delta_{10}^{\pm} are the two 1616-dimensional complex spin representations of Spin(10)\mathrm{Spin}(10).

We also use the following notation

P=F4Spin(9),S=F4Spin(7),Q=F4Spin(7)×C2Spin(2).P=\frac{F_{4}}{\mathrm{Spin}(9)},\quad S=\frac{F_{4}}{\mathrm{Spin}(7)},\quad Q=\frac{F_{4}}{\mathrm{Spin}(7)\times_{C_{2}}\mathrm{Spin}(2)}.

These three homogeneous spaces for F4F_{4} have dimensions 16,31,3016,31,30 respectively, and all three are submanifolds of RR.

We study RR by using the action of F4E6F_{4}\subset E_{6} on RR. The idea for this approach comes from Atiyah and Berndt [AB03]. The action of F4F_{4} on RR has three orbit types. The principal (generic) orbit of codimension 11 is SS and there are 22 special orbits PP and QQ. This orbit structure is described in the Appendix to [AB03] and the paper [LM01]. It can also be derived from [Ada96, Chapter 14]. In standard terminology, the action of F4F_{4} on RR has cohomogeneity 11, and the orbit space R/F4R/F_{4} is a closed interval.

Using Mostert’s cohomogeneity one theorem [Mos57, Mos57a, GGZ18, Bre72] we get the following result.

Theorem 1.

Let NPN_{P}, NQN_{Q} be the normal bundles of PP and QQ in RR. Let DPD_{P}, DQD_{Q} be the disc bundles in NPN_{P}, NQN_{Q} and SPS_{P}, SQS_{Q} be the corresponding sphere bundles. There are diffeomorphisms

SPe1Se2SQ\begin{CD}S_{P}@<{e_{1}}<{}<S@>{e_{2}}>{}>S_{Q}\end{CD}

such that

R=DPSDQ.R=D_{P}\cup_{S}D_{Q}.

An equivalent way to describe RR is as the double mapping cylinder of the maps PSQP\leftarrow S\to Q given by the projections in the sphere bundles SPS_{P} and SQS_{Q}.

The 1616-dimensional real vector bundle NPN_{P} is the bundle over P=F4/Spin(9)P=F_{4}/\mathrm{Spin}(9) associated to the 1616-dimensional real spin representation of Spin(9)\mathrm{Spin}(9). This is also isomorphic to the tangent bundle of PP. The 22-dimensional real bundle NQN_{Q} over Q=F4/(Spin(7)×C2Spin(2))Q=F_{4}/(\mathrm{Spin}(7)\times_{C_{2}}\mathrm{Spin}(2)) is associated to the obvious representation Spin(7)×C2Spin(2)SO(2)\mathrm{Spin}(7)\times_{C_{2}}\mathrm{Spin}(2)\to\mathrm{SO}(2).

This geometric decomposition of RR allows us to give a new proof of a theorem originally proved by Toda and Watanabe [TW74].

Theorem 2.

The integral cohomology ring of RR is

H(R)=[t,w](r18,r24)H^{*}(R)=\frac{\mathbb{Z}[t,w]}{(r_{18},r_{24})}

where tH2(R)t\in H^{2}(R), wH8(R)w\in H^{8}(R) and the relations are

r18=t93w2t,r24=w39wt8+15w2t4.r_{18}=t^{9}-3w^{2}t,\quad r_{24}=w^{3}-9wt^{8}+15w^{2}t^{4}.

There is an equivalent way to present the cohomology of RR using Poincaré duality. Choosing a generator (fundamental class) [R][R] in H32(R)H_{32}(R)\cong\mathbb{Z} yields a non-degenerate bilinear pairing

μ:Hp(R)H32p(R),μ(a,b)=ab,[R].\mu:H^{p}(R)\otimes H^{32-p}(R)\to\mathbb{Z},\qquad\mu(a,b)=\langle ab,[R]\rangle.

This tells us that if we know all products that end up in the top degree, then we know all products. The following table, proved in Theorem 3.4, gives all products to the top degree.

t16t^{16} t12wt^{12}w t8w2t^{8}w^{2} t4w3t^{4}w^{3} w4w^{4}
3263\cdot 26 3153\cdot 15 2626 1515 99

Recall that RR is a 1616-dimensional smooth complex subvariety of 26\mathbb{C}\mathbb{P}^{26}, often called the fourth Severi variety [Zak85]. It is also a generalised flag variety, see [LV12, §1b]. We write Tc(R)T_{c}(R) for the 16-dimensional complex tangent bundle of RR. As an application, we explain how to calculate the Chern classes of Tc(R)T_{c}(R).

Theorem 3.

The Chern classes of Tc(R)T_{c}(R) are as follows.

c1\displaystyle c_{1} =12t,\displaystyle=12t,
c2\displaystyle c_{2} =69t2,\displaystyle=69t^{2},
c3\displaystyle c_{3} =252t3,\displaystyle=252t^{3},
c4\displaystyle c_{4} =657t46w,\displaystyle=657t^{4}-6w,
c5\displaystyle c_{5} =1296t536tw,\displaystyle=1296t^{5}-36tw,
c6\displaystyle c_{6} =1995t6102t2w,\displaystyle=1995t^{6}-102t^{2}w,
c7\displaystyle c_{7} =2448t7198t3w,\displaystyle=2448t^{7}-198t^{3}w,
c8\displaystyle c_{8} =2412t8288t4w+39w2,\displaystyle=2412t^{8}-288t^{4}w+39w^{2},
c9\displaystyle c_{9} =270t5w+5760tw2,\displaystyle=-270t^{5}w+5760tw^{2},
c10\displaystyle c_{10} =180t6w+3645t2w2,\displaystyle=-180t^{6}w+3645t^{2}w^{2},
c11\displaystyle c_{11} =432t7w+2430t3w2,\displaystyle=-432t^{7}w+2430t^{3}w^{2},
c12\displaystyle c_{12} =750t4w2136w3,\displaystyle=750t^{4}w^{2}-136w^{3},
c13\displaystyle c_{13} =360tw3,\displaystyle=360tw^{3},
c14\displaystyle c_{14} =84t2w3,\displaystyle=84t^{2}w^{3},
c15\displaystyle c_{15} =1512t3w3864t7w2\displaystyle=1512t^{3}w^{3}-864t^{7}w^{2}
c16\displaystyle c_{16} =3w4.\displaystyle=3w^{4}.

In Section 1 we explain the role that the triality automorphism of Spin(8)\mathrm{Spin}(8) plays in this study of RR. Section 2 is devoted to the calculation of H(Q)H^{*}(Q). In Section 3 we give a proof of Theorem 2. Finally, in Section 4 we turn to the calculation of the characteristic classes of vector bundles over RR, proving Theorem 3.

Acknowledgments

This research was funded in part by the EPSRC, EP/W000466/1 (Thomas). For the purpose of open access, the author has applied a Creative Commons Attribution (CC BY) licence to any Author Accepted Manuscript version arising from this submission.

1. Triality and homogeneous spaces of F4F_{4}

One of the special features of Spin(8)\mathrm{Spin}(8) is triality. Recall that Spin(8)\mathrm{Spin}(8) has three 88-dimensional real representations, the vector representation u8u_{8} and the two spin representations δ8±\delta^{\pm}_{8}. No two of these representations are isomorphic but given any two there is an outer automorphism of Spin(8)\mathrm{Spin}(8) which transforms one to the other. Indeed the group Out(Spin(8))\mathrm{Out}(\mathrm{Spin}(8)) of outer automorphisms of Spin(8)\mathrm{Spin}(8) can be identified with Σ3\Sigma_{3}, the group of permutations of the set {u8,δ8+,δ8}\{u_{8},\delta_{8}^{+},\delta_{8}^{-}\}.

The representation u8u_{8} gives a transitive action of Spin(8)\mathrm{Spin}(8) on S7S^{7} with stabiliser Spin(7)\mathrm{Spin}(7). By triality, the same is true for δ8+\delta_{8}^{+} and δ8\delta^{-}_{8}. So we get three (conjugacy classes) of embeddings

i,j+,j:Spin(7)Spin(8),i,j_{+},j_{-}:\mathrm{Spin}(7)\to\mathrm{Spin}(8),

and each of the homogeneous spaces

Spin(8)i(Spin(7)),Spin(8)j+(Spin(7)),Spin(8)j(Spin(7))\frac{\mathrm{Spin}(8)}{i(\mathrm{Spin}(7))},\quad\frac{\mathrm{Spin}(8)}{j_{+}(\mathrm{Spin}(7))},\quad\frac{\mathrm{Spin}(8)}{j_{-}(\mathrm{Spin}(7))}

are Spin(8)\mathrm{Spin}(8)-equivariariantly diffeomorphic to S7S^{7}.

Now we have the usual embeddings

Spin(8)Spin(9)F4.\mathrm{Spin}(8)\to\mathrm{Spin}(9)\to F_{4}.

The embeddings j+j_{+}, jj_{-} are conjugate in Spin(9)\mathrm{Spin}(9) and we get two conjugacy classes of embeddings

i,j:Spin(7)Spin(9).i,j:\mathrm{Spin}(7)\to\mathrm{Spin}(9).

The embeddings i,ji,j are conjugate in F4F_{4}, see [Ada81].

First we identify the homogeneous spaces

Spin(9)i(Spin(7)),Spin(9)j(Spin(7)).\frac{\mathrm{Spin}(9)}{i(\mathrm{Spin}(7))},\quad\frac{\mathrm{Spin}(9)}{j(\mathrm{Spin}(7))}.
Lemma 1.1.
  1. (i)

    The homogeneous space Spin(9)/i(Spin(7))\mathrm{Spin}(9)/i(\mathrm{Spin}(7)) is Spin(9)\mathrm{Spin}(9)-equivariantly diffeomorphic to the Stiefel manifold V2(9)V_{2}(\mathbb{R}^{9}) where Spin(9)\mathrm{Spin}(9) acts on V2(9)V_{2}(\mathbb{R}^{9}) via the vector representation.

  2. (ii)

    The homogeneous space Spin(9)/j(Spin(7))\mathrm{Spin}(9)/j(\mathrm{Spin}(7)) is Spin(9)\mathrm{Spin}(9)-equivariantly diffeomorphic to S15S^{15} where Spin(9)\mathrm{Spin}(9) acts on S15S^{15} as the sphere in the 1616-dimensional real spin representation of Spin(9)\mathrm{Spin}(9).

Proof.

The embedding ii is conjugate to the usual embedding of Spin(7)\mathrm{Spin}(7) in Spin(9)\mathrm{Spin}(9), and so the homogeneous space Spin(9)/i(Spin(7))\mathrm{Spin}(9)/i(\mathrm{Spin}(7)) is the Stiefel manifold V2(9)V_{2}(\mathbb{R}^{9}) of 22-frames in 9\mathbb{R}^{9}. The spin representation of Spin(9)\mathrm{Spin}(9) is a 1616-dimensional real representation and Spin(9)\mathrm{Spin}(9) acts transitively on S15S^{15}, the sphere in the spin representation. This action is transitive and the stabiliser of a point is (conjugate to) j(Spin(7))j(\mathrm{Spin}(7)), see [Bry20]. ∎

Next we identify the homogeneous spaces

F4i(Spin(7)),F4j(Spin(7)).\frac{F_{4}}{i(\mathrm{Spin}(7))},\quad\frac{F_{4}}{j(\mathrm{Spin}(7))}.

The first is the total space of the fibre bundle

Spin(9)i(Spin(7))F4i(Spin(7))F4Spin(9).\frac{\mathrm{Spin}(9)}{i(\mathrm{Spin}(7))}\to\frac{F_{4}}{i(\mathrm{Spin}(7))}\to\frac{F_{4}}{\mathrm{Spin}(9)}.

The fibre of this bundle is the Stiefel manifold V2(9)V_{2}(\mathbb{R}^{9}). Let U9U_{9} be the real 99-dimensional real vector bundle over F4/Spin(9)F_{4}/\mathrm{Spin}(9) associated to the 99-dimensional vector representation of Spin(9)\mathrm{Spin}(9). It follows from the previous lemma that F4/i(Spin(7))F_{4}/i(\mathrm{Spin}(7)) is F4F_{4}-equivariantly diffeomorphic to the fibrewise Stiefel manifold V2(U9)V_{2}(U_{9}), that is the fibre bundle over PP with fibre over xPx\in P equal to V2(U9,x)V_{2}(U_{9,x}), where U9,xU_{9,x} is the fibre of U9U_{9} over xx.

The second is the total space of the fibre bundle

Spin(9)j(Spin(7)).F4j(Spin(7))F4Spin(9).\frac{\mathrm{Spin}(9)}{j(\mathrm{Spin}(7))}.\to\frac{F_{4}}{j(\mathrm{Spin}(7))}\to\frac{F_{4}}{\mathrm{Spin}(9)}.

Let Δ9\Delta_{9} be the 1616-dimensional real vector bundle over P=F4/Spin(9)P=F_{4}/\mathrm{Spin}(9) associated to the (real) spin representation of Spin(9)\mathrm{Spin}(9). This time the previous lemma tells us that F4/j(Spin(7))F_{4}/j(\mathrm{Spin}(7)) is F4F_{4}-equivariantly diffeomorphic to S(Δ9)S(\Delta_{9}), the sphere bundle of Δ9\Delta_{9}.

As mentioned above, the embeddings i,j:Spin(7)F4i,j:\mathrm{Spin}(7)\to F_{4} are conjugate. This proves the following theorem.

Theorem 1.2.

There are F4F_{4}-equivariant diffeomorphisms

S(Δ9)e1F4Spin(7)e2V2(U9)S(\Delta_{9})\xleftarrow{\hphantom{he}e_{1}\hphantom{he}}\frac{F_{4}}{\mathrm{Spin}(7)}\xrightarrow{\hphantom{he}e_{2}\hphantom{he}}V_{2}(U_{9})

This theorem is quite surprising. We get two fibre bundles

S15S(Δ9)P,V2(9)V2(U9)PS^{15}\to S(\Delta_{9})\to P,\quad V_{2}(\mathbb{R}^{9})\to V_{2}(U_{9})\to P

with diffeomorphic total spaces and bases. Clearly, since the fibres of these bundles are not diffeomorphic, there is no fibre preserving diffeomorphism. The E2E_{2} pages of the Serre spectral sequences of the two fibre bundles look very different but they converge to the same answer.

2. The integral cohomology groups of QQ.

2.1. The integral cohomology of PP and SS

The homogeneous space P=F4/Spin(9)P=F_{4}/\mathrm{Spin}(9) is the Cayley projective plane. Its integral cohomology and Pontryagin classes are well known, see [BH58, §19]. As a ring

H(P)=[a](a3) with aH8(P).H^{*}(P)=\frac{\mathbb{Z}[a]}{(a^{3})}\quad\text{ with }a\in H^{8}(P).

Furthermore aa can be chosen so that the total Pontryagin class p(TP)p(TP) and the Euler class e(TP)e(TP) are given by

p(TP)=1+6a+39a2 and e(TP)=3a2.p(TP)=1+6a+39a^{2}\quad\text{ and }\quad e(TP)=3a^{2}.

The homogeneous space SS is the sphere bundle S(TP)S(TP). A simple argument with the Gysin sequence of this sphere bundle shows that

Hi(S)={if i=0,8,23,31,/3if i=16,0 otherwise.H^{i}(S)=\begin{cases}\mathbb{Z}\quad&\text{if $i=0,8,23,31$,}\\ \mathbb{Z}/3&\text{if $i=16$,}\\ 0&\text{ otherwise}.\end{cases}

2.2. The integral cohomology groups of QQ

An argument along the lines of [JRT23, Section 7.4], shows that

H(Q;)=[a2,a8](ρ16,ρ24)H^{*}(Q;\mathbb{Q})=\frac{\mathbb{Q}[a_{2},a_{8}]}{(\rho_{16},\rho_{24})}

and its Poincaré polynomial is

(1+x2+x4++x14)(1+x8+x16).(1+x^{2}+x^{4}+\dots+x^{14})(1+x^{8}+x^{16}).

We also know from [Bot56, Theorem A] that H(Q)H^{*}(Q) is torsion free.

Let q:QPq:Q\to P be the fibre bundle with fibres diffeomorphic to the Grassmannian of oriented 22-planes in 9\mathbb{R}^{9}

Gr2(9)=Spin(9)Spin(7)×C2Spin(2),\mathrm{Gr}_{2}(\mathbb{R}^{9})=\frac{\mathrm{Spin}(9)}{\mathrm{Spin}(7)\times_{C_{2}}\mathrm{Spin}(2)},

also known as the complex quadric.

There are elements eH2(Gr2(9))e\in H^{2}(\mathrm{Gr}_{2}(\mathbb{R}^{9})) and bH8(Gr2(9))b\in H^{8}(\mathrm{Gr}_{2}(\mathbb{R}^{9})) such that

H(Gr2(9))=[e,b](e42b,b2).H^{*}(\mathrm{Gr}_{2}(\mathbb{R}^{9}))=\frac{\mathbb{Z}[e,b]}{(e^{4}-2b,b^{2})}.

See [Bot58, Section 9] for a discussion of this result.

The ring homomorphism q:H(P)H(Q)q^{*}:H^{*}(P)\to H^{*}(Q) makes H(Q)H^{*}(Q) into a module over H(P)H^{*}(P). The following result gives the structure of H(Q)H^{*}(Q) as a module over H(P)H^{*}(P).

Lemma 2.1.

Let i:Gr2(9)Qi:\mathrm{Gr}_{2}(\mathbb{R}^{9})\to Q be the inclusion of a fibre of q:QPq:Q\to P.

  1. (i)

    i:H(Q)H(Gr2(9))i^{*}:H^{*}(Q)\to H^{*}(\mathrm{Gr}_{2}(\mathbb{R}^{9})) is surjective, and q:H(P)H(Q)q^{*}:H^{*}(P)\to H^{*}(Q) is injective.

  2. (ii)

    Choose sH2(Q)s\in H^{2}(Q), vH8(Q)v\in H^{8}(Q) such that i(s)=ei^{*}(s)=e and i(v)=bi^{*}(v)=b. Then H(Q)H^{*}(Q) is a free module over H(P)H^{*}(P) with basis

    1,s,s2,s3,v,sv,s2v,s3v.1,\quad s,\quad s^{2},\quad s^{3},\quad v,\quad sv,\quad s^{2}v,\quad s^{3}v.
  3. (iii)

    The ring H(Q)H^{*}(Q) is generated by s,v,as,v,a.

Proof.

Since both H(P)H^{*}(P) and H(Gr2(9))H^{*}(\mathrm{Gr}_{2}(\mathbb{R}^{9})) are zero in odd degrees the Serre spectral sequence of the fibre bundle q:QPq:Q\to P collapses at the E2E_{2} page. The result follows from the Leray-Hirsch theorem. ∎

Since q:H(P)=[a]/(a3)H(Q)q^{*}:H^{*}(P)=\mathbb{Z}[a]/(a^{3})\to H^{*}(Q) is injective, from now on we regard [a]/(a3)\mathbb{Z}[a]/(a^{3}) as a subring of H(Q)H^{*}(Q). We have a basis for H(Q)H^{*}(Q) and we need to calculate products between ss, vv, and aa. To do this we use characteristic classes.

2.3. Products in H(Q)H^{*}(Q)

The real representations Spin(7)×C2Spin(2)SO(7)\mathrm{Spin}(7)\times_{C_{2}}\mathrm{Spin}(2)\to\mathrm{SO}(7) and Spin(7)×C2Spin(2)SO(2)\mathrm{Spin}(7)\times_{C_{2}}\mathrm{Spin}(2)\to\mathrm{SO}(2) define real homogeneous vector bundles U7U_{7} and U2U_{2} of dimensions 77 and 22 over QQ. The real representation Spin(9)SO(9)\mathrm{Spin}(9)\to\mathrm{SO}(9) defines a real vector bundle U9U_{9} over PP. Evidently

U7U2=q(U9)U_{7}\oplus U_{2}=q^{*}(U_{9})

and applying characteristic classes to this identity of vector bundles will give us relations between products of s,v,as,v,a.

Lemma 2.2.

Let i:Gr2(9)Qi:\mathrm{Gr}_{2}(\mathbb{R}^{9})\rightarrow Q be the inclusion of a fibre of q:QPq:Q\to P. There exists a unique element vH8(Q)v\in H^{8}(Q) such that

i(v)=b,2v=s4+a.i^{*}(v)=b,\quad 2v=s^{4}+a.
Proof.

Let ww be the total Stiefel-Whitney class. Then applying ww to the above relation between bundles gives

w(U7)w(U2)=q(w(U9)).w(U_{7})w(U_{2})=q^{*}(w(U_{9})).

Simple arguments show that in H(P;/2)H^{*}(P;\mathbb{Z}/2), and H(Q;/2)H^{*}(Q;\mathbb{Z}/2) respectively,

w(U9)=1+a,w(U2)=1+smod2.w(U_{9})=1+a,\quad w(U_{2})=1+s\mod 2.

It follows that

w(U7)(1+s)=1+amod2,w(U_{7})(1+s)=1+a\mod 2,

and in particular

w8(U7)=s4+amod2.w_{8}(U_{7})=s^{4}+a\mod 2.

Since U7U_{7} is 77-dimensional, w8(U7)=0w_{8}(U_{7})=0 and we conclude that

s4+a=0mod2.s^{4}+a=0\mod 2.

This shows that s4+aH8(Q)s^{4}+a\in H^{8}(Q) is divisible by 22 and since H(Q)H^{*}(Q) is torsion free it is uniquely divisible by 22. ∎

Corollary 2.3.

Let ss and vv be as in the above lemma. The ring H(Q)H^{*}(Q) is generated by sH2(Q)s\in H^{2}(Q) and vH8(Q)v\in H^{8}(Q).

Proof.

Lemma 2.1 shows that H(Q)H^{*}(Q) is generated by s,a,vs,a,v and Lemma 2.2 shows that 2v=s4+a2v=s^{4}+a. ∎

This allows us to fix the generators of H(Q)H^{*}(Q), explicitly:

  1. (i)

    sH2(Q)s\in H^{2}(Q) is the Euler class of U2U_{2},

  2. (ii)

    vH8(Q)v\in H^{8}(Q) is the unique class such i(v)=bi^{*}(v)=b and 2v=s4+a2v=s^{4}+a.

We now need to find two relations between these generators, one in degree 1616 and the other in degree 2424.

Lemma 2.4.

In H16(Q)H^{16}(Q) we have the relation

s8=3v2.s^{8}=3v^{2}.
Proof.

Let pp be the total Pontryagin class. Since H(Q)H^{*}(Q) is torsion free, the relation U7+U2=q(U9)U_{7}+U_{2}=q^{*}(U_{9}) between bundles gives the relation

p(U2)p(U7)=q(p(U9)).p(U_{2})p(U_{7})=q^{*}(p(U_{9})).

Now p(U2)=1s2p(U_{2})=1-s^{2} and from [BH58, Theorem 19.4] we know that p(U9)=16a3a2p(U_{9})=1-6a-3a^{2}. It follows that

p4(U7)=s86as43a2.p_{4}(U_{7})=s^{8}-6as^{4}-3a^{2}.

Since U7U_{7} is 77-dimensional, pi(U7)=0p_{i}(U_{7})=0 for i4i\geq 4 and so

s86as43a2=0.s^{8}-6as^{4}-3a^{2}=0.

Substitute a=2vs4a=2v-s^{4} and this simplifies to

4s8=12v2.4s^{8}=12v^{2}.

Since H(Q)H^{*}(Q) is torsion free this proves the lemma. ∎

The relation in degree 2424 is the obvious one coming from PP:

a3=(2vs4)3=0.a^{3}=(2v-s^{4})^{3}=0.
Theorem 2.5.

The integral cohomology ring of QQ is given by

H(Q)=[s,v](ρ16,ρ24),H^{*}(Q)=\frac{\mathbb{Z}[s,v]}{(\rho_{16},\rho_{24})},

where sH2(Q)s\in H^{2}(Q), vH8(Q)v\in H^{8}(Q) are as above, and the relations are

ρ16=s83v2,ρ24=(2vs4)3.\rho_{16}=s^{8}-3v^{2},\quad\rho_{24}=(2v-s^{4})^{3}.
Proof.

We have constructed a surjective ring homomorphism

[s,v](ρ16,ρ24)H(Q).\frac{\mathbb{Z}[s,v]}{(\rho_{16},\rho_{24})}\to H^{*}(Q).

A simple counting argument shows the rank of the homogeneous degree kk part of [s,v]/(ρ16,ρ24)\mathbb{Z}[s,v]/(\rho_{16},\rho_{24}) is free abelian with the same rank as Hk(Q)H^{k}(Q). Therefore this ring homomorphism is an isomorphism. ∎

Finally we choose a generator [Q]H30(Q)=[Q]\in H^{30}(Q)=\mathbb{Z} and use this to write down the products in H(Q)H^{*}(Q) to the top dimension.

Theorem 2.6.

The products to H30(Q)H^{30}(Q) are given by the following table.

s15s^{15} s11vs^{11}v s7v2s^{7}v^{2} s3v3s^{3}v^{3}
3263\cdot 26 3153\cdot 15 2626 1515
Proof.

The abelian group H30(Q)H^{30}(Q) has the following presentation. There are four generators

s15,s11v,s7v2,s3v3s^{15},\quad s^{11}v,\quad s^{7}v^{2},\quad s^{3}v^{3}

and three relations

s7ρ16=0,s3vρ16=0,s3ρ24=0.s^{7}\rho_{16}=0,\quad s^{3}v\rho_{16}=0,\quad s^{3}\rho_{24}=0.

We let

a=s7v2,b=s3v3.a=s^{7}v^{2},\quad b=s^{3}v^{3}.

Then the relations show that

s15=3a,s11v=3b,15a=26b.s^{15}=3a,\quad s^{11}v=3b,\quad 15a=26b.

So H30(Q)H^{30}(Q) is the abelian group generated by a,ba,b with the single relation 15a=26b15a=26b. Since 1515 and 2626 are coprime this abelian group is isomorphic to \mathbb{Z}, and the isomorphism is the unique homomorphism such that a26a\mapsto 26 and b15b\mapsto 15. This gives the values in the table. ∎

3. The integral cohomology ring of RR.

We know that the cohomology of both RR and QQ is torsion free and zero in odd degrees. From [JRT23, Section 7.4], we see that the Poincaré polynomial of RR is

(1+x2+x4++x16)(1+x8+x16),(1+x^{2}+x^{4}+\dots+x^{16})(1+x^{8}+x^{16}),

and recall from Section 2.2 that the Poincaré polynomial of QQ is

(1+x2+x4++x14)(1+x8+x16).(1+x^{2}+x^{4}+\dots+x^{14})(1+x^{8}+x^{16}).

The following table, which highlights the very small difference between RR and QQ, lists the non-zero Betti numbers of RR and QQ.

2k 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32
b2k(R5)b_{2k}(R_{5}) 1 1 1 1 2 2 2 2 3 2 2 2 2 1 1 1 1
b2k(Q)b_{2k}(Q) 1 1 1 1 2 2 2 2 2 2 2 2 1 1 1 1

Let i:PRi:P\to R be the embedding of PP in RR. Then we have the usual induced homomorphisms i,ii_{*},i^{*} in homology and cohomology, respectively. However both PP and RR are orientable manifolds and ii is a codimension 1616 embedding, so we also have the umkehr homomorphisms

i!:Hr(P)Hr+16(R).i_{!}:H^{r}(P)\to H^{r+16}(R).

Recall that i!i_{!} is defined by the following commutative diagram

Hr(P)i!Hr+16(R)H16r(Q)iH16r(R)\begin{CD}H^{r}(P)@>{i_{!}}>{}>H^{r+16}(R)\\ @V{}V{}V@V{}V{}V\\ H_{16-r}(Q)@>{}>{i_{*}}>H_{16-r}(R)\end{CD}

where the vertical arrows are the Poincaré duality isomorphisms. Now let j:QRj:Q\to R be the codimension 22 embedding of QQ. In this case we have j,jj_{*},j^{*} and

j!:Hr(Q)Hr+2(R).j_{!}:H^{r}(Q)\to H^{r+2}(R).

We know that the cohomology groups of P,Q,RP,Q,R are zero in odd degrees.

Lemma 3.1.

There are short exact sequences

0H2k2(Q)j!H2k(R)iH2k(P)0,\begin{CD}0@>{}>{}>H^{2k-2}(Q)@>{}>{j_{!}}>H^{2k}(R)@>{}>{i^{*}}>H^{2k}(P)@>{}>{}>0,\end{CD}
0H2k16(P)i!H2k(R)jH2k(Q)0.\begin{CD}0@>{}>{}>H^{2k-16}(P)@>{}>{i_{!}}>H^{2k}(R)@>{}>{j^{*}}>H^{2k}(Q)@>{}>{}>0.\end{CD}
Proof.

By Theorem 1, R/PR/P is the Thom space Th(NQ)\mathrm{Th}(N_{Q}) of the normal bundle of j:QRj:Q\to R. The dimension of this bundle is 22 so the Thom isomorphism

Hs2(Q)Hs(Th(NQ))H^{s-2}(Q)\to H^{s}(\mathrm{Th}(N_{Q}))

shows R/P=Th(NQ)R/P=\mathrm{Th}(N_{Q}) has no cohomology in odd degrees. Therefore the connecting homomorphism in the long exact sequence of the pair (R,P)(R,P) is always zero. Finally the composite

Hs2(Q)Hs(Th(NQ))=Hs(R/P)Hs(R)H^{s-2}(Q)\to H^{s}(\mathrm{Th}(N_{Q}))=H^{s}(R/P)\to H^{s}(R)

is an alternative definition of j!j_{!}. This gives the first exact sequence. The second follows by identifying R/QR/Q with Th(NP)\mathrm{Th}(N_{P}) and repeating the argument in this context. ∎

Corollary 3.2.
  1. (i)

    The map j:Hs(R)Hs(Q)j^{*}:H^{s}(R)\to H^{s}(Q) is an isomorphism for s15s\leq 15.

  2. (ii)

    The map j!:Hs(Q)Hs+2(R)j_{!}:H^{s}(Q)\to H^{s+2}(R) is an isomorphism for s17s\geq 17.

  3. (iii)

    There is a short exact sequence

    0H14(Q)j!H16(R)iH16(P)0\begin{CD}0@>{}>{}>H^{14}(Q)@>{}>{j_{!}}>H^{16}(R)@>{}>{i^{*}}>H^{16}(P)@>{}>{}>0\end{CD}

First, we define tH2(R)t\in H^{2}(R) and wH8(R)w\in H^{8}(R) by

j(t)=s,j(w)=v.j^{*}(t)=s,\quad j^{*}(w)=v.
Lemma 3.3.

The ring H(R)H^{*}(R) is generated by tt and ww.

Proof.

Suppose xHk(R)x\in H^{k}(R) and k15k\leq 15. Then since s,vs,v generate H(Q)H^{*}(Q) the above corollary shows that x=j(p(s,v))x=j^{*}(p(s,v)) for some polynomial pp in two variables. By definition, j(s)=tj^{*}(s)=t and j(v)=wj^{*}(v)=w, so x=p(t,w)x=p(t,w). Therefore xx is a polynomial in t,wt,w.

Now suppose xHk(R)x\in H^{k}(R) and k17k\geq 17. This time, the corollary shows that x=j!(q(s,v))x=j_{!}(q(s,v)) for some polynomial qq in two variables. Since j(t)=sj^{*}(t)=s and j(w)=vj^{*}(w)=v it follows that q(s,v)=j(q(t,w))q(s,v)=j^{*}(q(t,w)). Now the general properties of j!j_{!} show that

j(j!(1))=s,j!(j(b)c)=bj!(c),j!(j(b))=bt.j^{*}(j_{!}(1))=s,\quad j_{!}(j^{*}(b)c)=bj_{!}(c),\quad j_{!}(j^{*}(b))=bt.

The first formula shows that j!(1)=tj_{!}(1)=t so the third formula follows from the first two in the special case c=1c=1. It follows that

x=j!(q(s,v))=j!(j(q(t,w)))=q(t,w)t.x=j_{!}(q(s,v))=j_{!}(j^{*}(q(t,w)))=q(t,w)t.

This shows that if k17k\geq 17 any xHk(R)x\in H^{k}(R) can be written as a polynomial in tt and ww.

We are left to prove that the same conclusion is true if xH16(R)x\in H^{16}(R). We claim that i(w2)=a2i^{*}(w^{2})=a^{2}. Assuming this claim is true, consider the exact sequence

0H14(Q)H16(R)H16(P)00\to H^{14}(Q)\to H^{16}(R)\to H^{16}(P)\to 0

in Corollary 3.2. Then H14(Q)H^{14}(Q) is \mathbb{Z}\oplus\mathbb{Z} with basis s7,s3vs^{7},s^{3}v. It follows that j!(s7),j!(s3v),w2j_{!}(s^{7}),j_{!}(s^{3}v),w^{2} is a basis for H16(R)H^{16}(R). Repeating the argument of the previous paragraph shows that j!(s7)=t8j_{!}(s^{7})=t^{8} and j!(s3v)=t4wj_{!}(s^{3}v)=t^{4}w. So every element in H16(R)H^{16}(R) is also given by a polynomial in tt and ww.

It remains to prove that i(w2)=a2i^{*}(w^{2})=a^{2}. Corollary 3.2 shows that the homomorphism i:H8(R)H8(P)i^{*}:H^{8}(R)\to H^{8}(P) is surjective. We also know that t8,wt^{8},w is a basis for H8(R)H^{8}(R). For degree reasons i(t)=0i^{*}(t)=0 and so i(t4)=0i^{*}(t^{4})=0. Thus, i(w)=±ai^{*}(w)=\pm a. ∎

Next we show how to compute the products to the top degree in RR.

Theorem 3.4.

The products to H32(R)H^{32}(R) are given by the following table.

t16t^{16} t12wt^{12}w t8w2t^{8}w^{2} t4w3t^{4}w^{3} w4w^{4}
3263\cdot 26 3153\cdot 15 2626 1515 99
Proof.

Note that

j!:H30(Q)H32(R)j_{!}:H^{30}(Q)\to H^{32}(R)

is an isomorphism. The first four entries in the table follow by applying j!j_{!} to the entries in the table in Theorem 2.6 and using the formula j!(j(x))=xtj_{!}(j^{*}(x))=xt.

Now we need to compute xx, the entry in the table corresponding to w4w^{4}. The abelian group H16(R)H^{16}(R) is free of rank 33 with basis

t8,t4w,w2.t^{8},\quad t^{4}w,\quad w^{2}.

The intersection matrix with respect to this basis is

(3263152631526152615x).\begin{pmatrix}3\cdot 26&3\cdot 15&26\\ 3\cdot 15&26&15\\ 26&15&x\end{pmatrix}.

This matrix must have determinant ±1\pm 1. The determinant is 3x263x-26 so it follows that xx must be 99. ∎

This completely determines the ring H(R)H^{*}(R) but it would seem quite perverse not to extract the degree 1818 and degree 2424 relations it implies.

Lemma 3.5.
  1. (i)

    In H18(R)H^{18}(R) we have the relation

    t93w2t=0.t^{9}-3w^{2}t=0.
  2. (ii)

    In H24(R)H^{24}(R) we have the relation

    w3+15w2t49wt8=0.w^{3}+15w^{2}t^{4}-9wt^{8}=0.
Proof.

To prove (i) we argue as follows. In H16(Q)H^{16}(Q)

j(t83w2)=s83v2=0j^{*}(t^{8}-3w^{2})=s^{8}-3v^{2}=0

and therefore

0=j!((j(t83w2))=(t83w2)t.0=j_{!}((j^{*}(t^{8}-3w^{2}))=(t^{8}-3w^{2})t.

To prove (ii) we simply check that

t4(w3+15w2t49wt8)=w(w3+15w2t49wt8)=0t^{4}(w^{3}+15w^{2}t^{4}-9wt^{8})=w(w^{3}+15w^{2}t^{4}-9wt^{8})=0

and it follows that w3+15w2t49wt8=0w^{3}+15w^{2}t^{4}-9wt^{8}=0 by Poincaré duality. ∎

We now complete the proof of Theorem 2 in the same way we completed the proof of Theorem 2.5. We have constructed a surjective ring homomorphism

[t,w](r16,r24)H(R).\frac{\mathbb{Z}[t,w]}{(r_{16},r_{24})}\to H^{*}(R).

We know that Hk(R)H^{k}(R) is a free abelian group and we know its rank. A counting argument shows that the rank of the homogeneous degree kk part of [t,w]/(r16,r24)\mathbb{Z}[t,w]/(r_{16},r_{24}) is free abelian and has the same rank as Hk(R)H^{k}(R).

4. The characteristic classes of the natural bundles over RR

4.1. Bundles over RR

We start with the relevant representation theory. As usual we write ρ10\rho_{10} for the 1010-dimensional vector representation of Spin(10)\mathrm{Spin}(10), and δ10±\delta_{10}^{\pm} for the two 1616-dimensional spin representations. As in the introduction, we use (ϵ,i)(\epsilon,i) for the relevant central subgroup C4Spin(10)×S1C_{4}\subset\mathrm{Spin}(10)\times S^{1}. Recall that ϵ\epsilon acts as multiplication by ±i\pm i on δ10±\delta_{10}^{\pm} and as multiplication by 1-1 on ρ10\rho_{10}. A necessary and sufficient condition for a representation of Spin(10)×S1\mathrm{Spin}(10)\times S^{1} to descend to Spin(10)×C4S1\mathrm{Spin}(10)\times_{C_{4}}S^{1} is that this central subgroup C4C_{4} acts trivially.

Let ξ\xi be the usual representation of S1S^{1} on \mathbb{C}. Then

ρ10ξ2,δ10+ξ3,δ10ξ3,ξ4\rho_{10}\otimes\xi^{2},\quad\delta_{10}^{+}\otimes\xi^{3},\quad\delta_{10}^{-}\otimes\xi^{-3},\quad\xi^{4}

descend to Spin(10)×C4S1\mathrm{Spin}(10)\times_{C_{4}}S^{1} and so they define complex vector bundles

V,D,D¯,LV,\quad D,\quad\bar{D},\quad L

of dimensions 10,16,16,110,16,16,1 over R5R_{5}, respectively. From [Min77, Section 2] we have the following relation between bundles over RR

L1+V+DL1=R×27.L^{-1}+V+D\otimes L^{-1}=R\times\mathbb{C}^{27}.

We know the Chern classes of LL so once we know the Chern classes of DD we know the Chern classes of D¯\bar{D} and VV. We choose to compute the Chern classes of DD since it is straightforward to check from [Ada96, Chapter 6] that Tc(R)=DT_{c}(R)=D. We choose to give these Chern classes in terms of the presentation of Theorem 2 rather than the presentation in [JRT23, Theorem 8.2.1] since this leads to the most compact formulas. To do this, we must compare the two presentations.

4.2. Comparing the two presentations of the cohomology of RR

We work in the ring

H(R;)=[t,w](r18,r24)H^{*}(R;\mathbb{Q})=\frac{\mathbb{Q}[t,w]}{(r_{18},r_{24})}

as in Theorem 2.

In [JRT23, Theorem 8.2.1] we give a presentation of H(R;)H^{*}(R;\mathbb{Q}) with two generators a2H2(R;)a_{2}\in H^{2}(R;\mathbb{Q}), a8H8(R;)a_{8}\in H^{8}(R;\mathbb{Q}) and two relations. It is straightforward to check that t=4a2t=4a_{2} and when we substitute a2=t/4a_{2}=t/4 the two relations become

0\displaystyle 0 =ta82+274t5a83964t9,\displaystyle=ta_{8}^{2}+\frac{27}{4}t^{5}a_{8}-\frac{39}{64}t^{9}, (1)
0\displaystyle 0 =a83+3698t4a82299764t8a8+1539512t12.\displaystyle=a_{8}^{3}+\frac{369}{8}t^{4}a_{8}^{2}-\frac{2997}{64}t^{8}a_{8}+\frac{1539}{512}t^{12}. (2)

We must calculate a8a_{8} as a polynomial in t,wt,w.

Theorem 4.1.

The element

a8=6w278t4a_{8}=6w-\frac{27}{8}t^{4}

is the unique indecomposable element of the ring H(R;)H^{*}(R;\mathbb{Q}) satisfying relations (1) and (2).

Proof.

Let xH8(R;)x\in H^{8}(R;\mathbb{Q}) be such that

x=λt4+μw,tx2+αt5x=βt9x=\lambda t^{4}+\mu w,\quad tx^{2}+\alpha t^{5}x=\beta t^{9}

where α,β,λ,μ\alpha,\beta,\lambda,\mu\in\mathbb{Q} and μ0\mu\neq 0. The condition μ0\mu\neq 0 ensures that xx is indecomposable.

We work with the basis u=t9u=t^{9}, v=t5wv=t^{5}w for H18(R;)H^{18}(R;\mathbb{Q}). Calculating tx2+αt5x=βt9tx^{2}+\alpha t^{5}x=\beta t^{9} in this basis (using r18r_{18}), and comparing the coefficients of uu and vv gives two equations:

2λμ+αμ\displaystyle 2\lambda\mu+\alpha\mu =0,\displaystyle=0,
λ2+μ23+αλ\displaystyle\lambda^{2}+\frac{\mu^{2}}{3}+\alpha\lambda =β.\displaystyle=\beta.

Since μ0\mu\neq 0, we find that

λ=α2 and μ2=3(β+α24).\lambda=-\frac{\alpha}{2}\quad\text{ and }\quad\mu^{2}=3\left(\beta+\frac{\alpha^{2}}{4}\right).

Fixing α=27/4\alpha=27/4 and β=39/64\beta=39/64 leads to μ=±6\mu=\pm 6 and λ=27/8\lambda=-27/8. Therefore

a8=±6w278t4.a_{8}=\pm 6w-\frac{27}{8}t^{4}.

A final calculation shows that

a8=6w278t4a_{8}=6w-\frac{27}{8}t^{4}

is the only solution to relation (2). ∎

4.3. Calculating characteristic classes

In [JRT23, Section 8.2] we construct a homomorphism

ϕ:H(BSpin(10)×S1;)H(R;)\phi:H^{*}(B\mathrm{Spin}(10)\times S^{1};\mathbb{Q})\to H^{*}(R;\mathbb{Q})

such that

ϕ(ck(δ10+ξ3))=ck(Tc(R)).\phi(c_{k}(\delta_{10}^{+}\otimes\xi^{3}))=c_{k}(T_{c}(R)).

In Section 8.1 of [JRT23] we explain how to calculate ϕ\phi in terms of the standard choice of generators for the ring H(BSpin(10)×S1;)H^{*}(B\mathrm{Spin}(10)\times S^{1};\mathbb{Q}) and the generators a2a_{2} and a8a_{8} of H(R;)H^{*}(R;\mathbb{Q}) referred to in the previous section. We complete the proof of Theorem 3 in three steps.

  1. (i)

    First compute ck(δ10+ξ3)c_{k}(\delta_{10}^{+}\otimes\xi^{3}) using the splitting principle, see [JRT23, Section 8.6].

  2. (ii)

    Next compute ϕ(ck(δ10+ξ3))\phi(c_{k}(\delta_{10}^{+}\otimes\xi^{3})) in terms of the generators a2,a8a_{2},a_{8}.

  3. (iii)

    Finally use Section 4.2 to express ck(Tc(R))c_{k}(T_{c}(R)) in terms of the generators t,wt,w.

References

  • [AB03] Michael Atiyah and Jürgen Berndt “Projective planes, Severi varieties and spheres” In Surveys in differential geometry, Vol. VIII (Boston, MA, 2002) 8, Surv. Differ. Geom. Int. Press, Somerville, MA, 2003, pp. 1–27
  • [Ada81] J. F. Adams “Spin(8), Triality, F4F_{4} and all that” In Superspace & Supergravity Cambridge University Press, 1981, pp. 435–445
  • [Ada96] J. F. Adams “Lectures on exceptional Lie groups”, Chicago Lectures in Mathematics University of Chicago Press, Chicago, IL, 1996
  • [BH58] A. Borel and F. Hirzebruch “Characteristic classes and homogeneous spaces. I” In Amer. J. Math. 80, 1958, pp. 458–538
  • [Bot56] Raoul Bott “An application of the Morse theory to the topology of Lie-groups” In Bull. Soc. Math. France 84, 1956, pp. 251–281
  • [Bot58] Raoul Bott “The space of loops on a Lie group.” In Michigan Mathematical Journal 5.1 University of Michigan, Department of Mathematics, 1958, pp. 35 –61
  • [Bre72] G.E. Bredon “Introduction to Compact Transformation Groups”, Pure and Applied Mathematics Elsevier Science, 1972
  • [Bry20] Robert L. Bryant “Notes on spinors in low dimension”, 2020 arXiv:2011.05568 [math.GR]
  • [GGZ18] Fernando Galaz-García and Masoumeh Zarei “Cohomogeneity one topological manifolds revisited” In Math. Z. 288.3-4, 2018, pp. 829–853
  • [JRT23] John Jones, Dmitriy Rumynin and Adam Thomas “The classical topological invariants of homogeneous spaces”, 2023 arXiv:2310.1465 [math.AT]
  • [LM01] J. M. Landsberg and L. Manivel “The Projective Geometry of Freudenthal’s Magic Square” In Journal of Algebra 239.2, 2001, pp. 477–512
  • [LV12] R. Lazarsfeld and A. Ven “Topics in the Geometry of Projective Space”, Oberwolfach Seminars Birkhäuser Basel, 2012
  • [Min77] Haruo Minami KK-groups of EIIIEIII and FIIFII In Osaka Math. J. 14.1, 1977, pp. 173–177
  • [Mos57] Paul S. Mostert “On a compact Lie group acting on a manifold” In Ann. of Math. (2) 65, 1957, pp. 447–455
  • [Mos57a] Paul S. Mostert “Errata, “On a compact Lie group acting on a manifold”” In Ann. of Math. (2) 66, 1957, pp. 589
  • [TW74] Hirosi Toda and Takashi Watanabe “The integral cohomology ring of 𝐅4/𝐓{\bf F}_{4}/{\bf T} and 𝐄6/𝐓{\bf E}_{6}/{\bf T} In J. Math. Kyoto Univ. 14, 1974, pp. 257–286
  • [Zak85] F. L. Zak “Severi varieties” In Mat. Sb. (N.S.) 126(168).1, 1985, pp. 115–132, 144