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

Stable submanifolds in the product of projective spaces II

Shuli Chen And Alejandra Ramirez-Luna
Abstract.

We prove that there do not exist odd-dimensional stable compact minimal immersions in the product of two complex projective spaces. We also prove that the only stable compact minimal immersions in the product of a quaternionic projective space with any other Riemannian manifold are the products of quaternionic projective subspaces with compact stable minimal immersions of the second manifold in the Riemmanian product. These generalize similar results of the second-named author of immersions with low dimensions or codimensions to immersions with arbitrary dimensions. In addition, we prove that the only stable compact minimal immersions in the product of a octonionic projective plane with any other Riemannian manifold are the products of octonionic projective subspaces with compact stable minimal immersions of the second manifold in the Riemmanian product.

1. Introduction

A classical and fascinating problem in Riemannian geometry is the study of submanifolds that minimize area under perturbations in a given Riemannian manifold. This gives rise to the the study of stable minimal submanifolds, which are critical points of the area functional and whose second variation of the area functional is non-negative (equivalently, the mean curvature vector and Morse index both equal to zero [Sim68]). In particular, it is an interesting problem to obtain geometric information of the submanifold by just knowing that it is stable.

A lot of research has been done in that direction where the ambient manifold is well-known. For example, Fischer-Colbrie and Schoen [FCS80], do Carmo and Peng [dCP79], and Pogorelov [Pog81] independently proved that planes are the only stable complete minimal surfaces in the 33-dimensional Euclidean space, and recently Chodosh and Li proved that a complete, two-sided, stable minimal hypersurface in the 44-dimensional Euclidean space must be flat [CL21]. When the ambient manifold is compact, Simons proved that the are no compact stable minimal submanifolds in the Euclidean sphere [Sim68]. Later, Lawson and Simons characterized the complex submanifolds (in the sense that each tangent space is invariant under the complex structure) as the only compact stable minimal submanifolds in the complex projective space [LS73]. Finally, Ohnita completed the classification of compact stable minimal submanifolds in all compact rank one symmetric spaces by proving that the only stable submanifolds in the real and quaternionic projective space, and the Cayley plane, are the real and quaternionic projective subspaces, and the Cayley projective line, respectively [Ohn86].

In the case where the ambient manifold is a Riemannian product of well-known Riemannian manifolds we have the following: Torralbo and Urbano proved a characterization of stable submanifolds in the product of the Euclidean sphere with any other Riemannian manifold [TU14], and Chen and Wang proved a similar characterization of stable submanifolds in the product of any hypersurface of the Euclidean space with certain conditions and any Riemannian manifold [CW13].

Along these lines and following similar ideas of Torralbo, Urbano, Chen, and Wang, Ramirez-Luna in [RL22] proved a characterization theorem for stable submanifolds of specific dimensions in the product of a complex and quaternionic projective space with any other Riemannian manifold. In particular we recall the following results:

Theorem 1.1.

[RL22] The only compact stable minimal immersions of codimension d=2d=2 or dimension n=2n=2 in the product manifold M¯:=Pm12×Pm22\bar{M}:=\mathbb{C}P^{\frac{m_{1}}{2}}\times\mathbb{C}P^{\frac{m_{2}}{2}} are the complex ones, in the sense that each tangent space is invariant under the complex structure J1J_{1} or J2J_{2} of M¯\bar{M} (see Definition 3.7 in [RL22]).

Theorem 1.2.

[RL22] Let Φ=(ψ,ϕ):ΣPm14×M\Phi=(\psi,\phi):\Sigma\rightarrow\mathbb{H}P^{\frac{m_{1}}{4}}\times M be a compact stable minimal immersion of codimension dd and dimension nn, where MM is any Riemannian manifold of dimension m2m_{2}. Then,

  • If d=1d=1, Σ=Pm14×Σ^\Sigma=\mathbb{H}P^{\frac{m_{1}}{4}}\times\hat{\Sigma}, Φ=Id×ϕ^\Phi=\operatorname{Id}\times\hat{\phi} where ϕ^:Σ^M\hat{\phi}:\hat{\Sigma}\rightarrow M is a compact stable minimal immersion of codimension 11, and therefore Φ(Σ)=Pm14×ϕ^(Σ^)\Phi(\Sigma)=\mathbb{H}P^{\frac{m_{1}}{4}}\times\hat{\phi}(\hat{\Sigma}). In particular, for m2=1m_{2}=1, Σ=Pm14\Sigma=\mathbb{H}P^{\frac{m_{1}}{4}}, ϕ^\hat{\phi} is a constant function, and Φ(Σ)=Pm14×{q}\Phi(\Sigma)=\mathbb{H}P^{\frac{m_{1}}{4}}\times\{q\}, for qMq\in M.

  • If d=2d=2, Σ=Pm14×Σ^\Sigma=\mathbb{H}P^{\frac{m_{1}}{4}}\times\hat{\Sigma}, Φ=Id×ϕ^\Phi=\operatorname{Id}\times\hat{\phi} where ϕ^:Σ^M\hat{\phi}:\hat{\Sigma}\rightarrow M is a compact stable minimal immersion of codimension 22, and therefore Φ(Σ)=Pm14×ϕ^(Σ^)\Phi(\Sigma)=\mathbb{H}P^{\frac{m_{1}}{4}}\times\hat{\phi}(\hat{\Sigma}). In particular, for m2=1m_{2}=1, there are no compact stable minimal immersions of codimension 22 in Pm14×M\mathbb{H}P^{\frac{m_{1}}{4}}\times M. And for m2=2m_{2}=2, Σ=Pm14\Sigma=\mathbb{H}P^{\frac{m_{1}}{4}}, ϕ^\hat{\phi} is a constant function, and Φ(Σ)=Pm14×{q}\Phi(\Sigma)=\mathbb{H}P^{\frac{m_{1}}{4}}\times\{q\}, for qMq\in M.

  • If n=1n=1, ϕ:ΣM\phi:\Sigma\rightarrow M is a stable geodesic, ψ\psi is a constant function, and therefore Φ(Σ)={r}×ϕ(Σ)\Phi(\Sigma)=\{r\}\times\phi(\Sigma) with rr a point of Pm14\mathbb{H}P^{\frac{m_{1}}{4}}.

  • If n=2n=2, ϕ:ΣM\phi:\Sigma\rightarrow M is a stable minimal immersion of dimension 22, ψ\psi is a constant function, and therefore Φ(Σ)={r}×ϕ(Σ)\Phi(\Sigma)=\{r\}\times\phi(\Sigma) with rr a point of Pm14\mathbb{H}P^{\frac{m_{1}}{4}}.

Recall that the technique in these problems involving stable submanifolds, which goes back to Simons [Sim68], is to find appropriate normal sections, such that when we add the second variations along them, we obtain a non-positive sign. This together with the stability of the submanifold, allows us in some cases to obtain geometric information about the submanifold. Therefore, an essential part of the proofs of Theorems 1.1 and 1.2 is to prove that Equations (9) and (10) in Lemma 3.1 and (47) and (48) in Lemma 4.1 in [RL22] for the specific dimensions and codimensions have a non-positive sign (we call this the vector inequality). In [RL22], the vector inequality is proved in dimension one and two. In this paper we prove that the same vector inequality holds in general dimensions (see Proposition 2.2), and prove another inequality (see Proposition 3.10). This allows us to prove the nonexistence of odd-dimensional stable minimal submanifolds in the product of two complex projective spaces, and obtain a complete characterization of stable submanifolds in the product of a quaternionic projective space with any other Riemannian manifold, and stable minimal submanifolds in the product of a octonionic projective plane with any other Riemannian manifold.

From Theorem 1.1 (see also [TU14] and references in there), it is expected that stable minimal submanifolds in a Riemannian manifold with a complex structure behave well under the same complex structure. Towards this idea we prove that:

{restatable}

thmmainone There do not exist odd-dimensional stable compact minimal immersions in Pm1/2×Pm2/2\mathbb{C}P^{m_{1}/2}\times\mathbb{C}P^{m_{2}/2}. For even-dimensional stable compact minimal immersions in Pm1/2×Pm2/2\mathbb{C}P^{m_{1}/2}\times\mathbb{C}P^{m_{2}/2}, even though we obtain a non-positive sign in the second variation, this is not enough to conclude that the tangent space of the stable submanifold behaves well under a complex structure of Pm1/2×Pm2/2\mathbb{C}P^{m_{1}/2}\times\mathbb{C}P^{m_{2}/2}.

Theorem 1.2 shows that the stable minimal submanifolds of dimension or codimension 11 or 22 in Pm14×M\mathbb{H}P^{\frac{m_{1}}{4}}\times M are precisely the product of stable minimal submanifolds in Pm14\mathbb{H}P^{\frac{m_{1}}{4}} and MM. In this paper, we generalize this result to arbitrary dimension:

{restatable}

thmmaintwo Let Φ:ΣPm1/4×M\Phi:\Sigma\to\mathbb{H}P^{m_{1}/4}\times M be a stable compact minimal immersion, where MM is a Riemannian manifold of dimension m2m_{2}. Then Σ=Pm1/4×Σ2\Sigma=\mathbb{H}P^{m_{1}^{\prime}/4}\times\Sigma_{2}, Φ=ψ1×ψ2\Phi=\psi_{1}\times\psi_{2}, where ψ1:Pm1/4Pm1/4\psi_{1}:\mathbb{H}P^{m_{1}^{\prime}/4}\to\mathbb{H}P^{m_{1}/4}, ψ2:Σ2M\psi_{2}:\Sigma_{2}\to M are stable compact minimal immersions.

In this paper, we further consider the stable compact minimal immersions of arbitrary dimension in 𝕆P2×M\mathbb{O}P^{2}\times M. The linear algebra in this case is more involved due to non-associativity of octonions (see Proposition 3.10). Analogous to Theorem 1.2, we obtain the following result: {restatable}thmmainthree Let Φ:Σ𝕆P2×M\Phi:\Sigma\to\mathbb{O}P^{2}\times M be a stable compact minimal immersion, where MM is a Riemannian manifold of dimension m2m_{2}. Then Σ=Σ1×Σ2\Sigma=\Sigma_{1}\times\Sigma_{2}, Φ=ψ1×ψ2\Phi=\psi_{1}\times\psi_{2}, where Σ1\Sigma_{1} is either a point, 𝕆P1\mathbb{O}P^{1} or 𝕆P2\mathbb{O}P^{2}, and ψ1:Σ1𝕆P2\psi_{1}:\Sigma_{1}\to\mathbb{O}P^{2}, ψ2:Σ2M\psi_{2}:\Sigma_{2}\to M are stable compact minimal immersions.

This paper is structured as follows: In Section 2 we prove the vector inequality and necessary linear algebra facts for the following section. In Section 3, we give some preliminaries, and prove Theorems 1.2, 1.2, and 1.2.

Acknowledgments. The authors are grateful for the patience and useful discussions and suggestions of Otis Chodosh. S. C. is also grateful for useful discussions with Lie Qian. S. C. is partially sponsored by the Mary V. Sunseri Graduate Fellowship at Stanford University.

2. The Vector Inequality

First we prove some basic results in linear algebra that will be needed later.

Proposition 2.1.

Let XX be an m×nm\times n matrix. Then,

  • the row spaces of XTXXTX^{T}XX^{T} and XTX^{T} are the same.

  • the symmetric matrices XTXX^{T}X and XXTXX^{T} have the same nonzero eigenvalues with the same multiplicity.

Proof.
  • Let us denote by R(Y)R(Y) the row space of a matrix YY. It is clear that R(XTXXT)R(XT)R(X^{T}XX^{T})\subset R(X^{T}). For the reverse direction, it is enough to show ker(XTXXT)ker(XT)\operatorname{ker}(X^{T}XX^{T})\subset\operatorname{ker}(X^{T}). Let yker(XTXXT)y\in\operatorname{ker}(X^{T}XX^{T}). Then we have

    0=yTXXTXXTy=XXTy2,0=y^{T}XX^{T}XX^{T}y=\|XX^{T}y\|^{2},

    so XXTy=0XX^{T}y=0. Thus

    0=yTXXTy=XTy2,0=y^{T}XX^{T}y=\|X^{T}y\|^{2},

    showing XTy=0X^{T}y=0, so yker(XT)y\in\operatorname{ker}(X^{T}). Thus ker(XTXXT)ker(XT)\operatorname{ker}(X^{T}XX^{T})\subset\operatorname{ker}(X^{T}) as desired.

  • Let λ\lambda be a nonzero eigenvalue of XTXX^{T}X with multiplicity kk. Then there exist kk orthonormal eigenvectors v1,,vkv_{1},\dots,v_{k} with eigenvalue λ\lambda. Then, (XXT)(Xvi)=X(XTX)vi=λXvi(XX^{T})(Xv_{i})=X(X^{T}X)v_{i}=\lambda Xv_{i}, so XviXv_{i} are eigenvectors of XXTXX^{T} with eigenvalue λ\lambda. Moreover, we claim they are linearly independent. Let αi\alpha_{i} be such that iαiXvi=0\sum_{i}\alpha_{i}Xv_{i}=0. Then 0=XT(iαiXvi)=iαiXTXvi=iαiλvi=λiαivi0=X^{T}(\sum_{i}\alpha_{i}Xv_{i})=\sum_{i}\alpha_{i}X^{T}Xv_{i}=\sum_{i}\alpha_{i}\lambda v_{i}=\lambda\sum_{i}\alpha_{i}v_{i}. Since viv_{i}’s are linearly independent, we must have αi=0\alpha_{i}=0 for each ii. Thus XviXv_{i}’s are linearly independent.

    Reverting the roles of XXTXX^{T} and XTXX^{T}X shows the λ\lambda-eigenspaces of XXTXX^{T} and XTXX^{T}X have the same dimension. Thus XTXX^{T}X and XXTXX^{T} have the same nonzero eigenvalues with multiplicity as desired.

For any two m×nm\times n matrices A,BMm×n()A,B\in M_{m\times n}(\mathbb{R}), one can define the Frobenius or trace inner product as

A,B:=Tr(ATB)=i=1mj=1naijbij,\langle A,B\rangle:=\operatorname{Tr}(A^{T}B)=\sum_{i=1}^{m}\sum_{j=1}^{n}a_{ij}b_{ij},

which induces the norm

A2:=Tr(ATA)=i=1mj=1naij2.\|A\|^{2}:=\operatorname{Tr}(A^{T}A)=\sum_{i=1}^{m}\sum_{j=1}^{n}a_{ij}^{2}.

This is the usual Eulidean inner product by identifying Mm×n()M_{m\times n}(\mathbb{R}) with mn\mathbb{R}^{mn}.

We thus have the familiar Cauchy-Schwarz inequality:

A,BAB,\langle A,B\rangle\leq\|A\|\|B\|,

and we obtain equality if and only if AA and BB are nonnegative multiples of each other.

The particular nice thing about the trace operator is that we have

Tr(ATB)=Tr(BAT)\operatorname{Tr}(A^{T}B)=\operatorname{Tr}(BA^{T})

for any A,BMm×n()A,B\in M_{m\times n}(\mathbb{R}).

Proposition 2.2.

Let AO(m)A\in O(m) be an orthogonal matrix. Let x1,xnx_{1},\dots x_{n} be vectors in m\mathbb{R}^{m}. Let XX be the associated m×nm\times n matrix

X=[x1xn].X=\begin{bmatrix}x_{1}&\dots&x_{n}\end{bmatrix}.

Then

i,j=1nxi,Axj2i,j=1nxi,xj20,\sum_{i,j=1}^{n}\langle x_{i},Ax_{j}\rangle^{2}-\sum_{i,j=1}^{n}\langle x_{i},x_{j}\rangle^{2}\leq 0,

and we have equality if and only if XXTXX^{T} commutes with AA. In particular, when we have equality, Span{x1,,xn}\operatorname{Span}\{x_{1},\dots,x_{n}\} is AA-invariant.

Proof.

Notice that XTX=[xi,xj]i,j=1nX^{T}X=[\langle x_{i},x_{j}\rangle]_{i,j=1}^{n}, and XTAX=[xi,Axj]i,j=1nX^{T}AX=[\langle x_{i},Ax_{j}\rangle]_{i,j=1}^{n}, so

XTX2=i,j=1nxi,xj2\displaystyle\|X^{T}X\|^{2}=\sum_{i,j=1}^{n}\langle x_{i},x_{j}\rangle^{2}, and XTAX2=i,j=1nxi,Axj2\displaystyle\|X^{T}AX\|^{2}=\sum_{i,j=1}^{n}\langle x_{i},Ax_{j}\rangle^{2}.

Moreover, we also have that

XXT2=Tr(XXTXXT)=Tr(XTXXTX)=XTX2\|XX^{T}\|^{2}=\operatorname{Tr}(XX^{T}XX^{T})=\operatorname{Tr}(X^{T}XX^{T}X)=\|X^{T}X\|^{2}.

We calculate

ATXXTA2\displaystyle\|A^{T}XX^{T}A\|^{2} =Tr(ATXXTAATXXTA)\displaystyle=\operatorname{Tr}(A^{T}XX^{T}AA^{T}XX^{T}A)
=Tr(ATXXTXXTA)\displaystyle=\operatorname{Tr}(A^{T}XX^{T}XX^{T}A)
=Tr(AATXXTXXT)\displaystyle=\operatorname{Tr}(AA^{T}XX^{T}XX^{T})
=Tr(XXTXXT)\displaystyle=\operatorname{Tr}(XX^{T}XX^{T})
=XXT2.\displaystyle=\|XX^{T}\|^{2}.

Thus

i,j=1nxi,Axj2\displaystyle\sum_{i,j=1}^{n}\langle x_{i},Ax_{j}\rangle^{2} =XTAX2\displaystyle=\|X^{T}AX\|^{2}
=Tr((XTAX)TXTAX)\displaystyle=\operatorname{Tr}((X^{T}AX)^{T}X^{T}AX)
=Tr((XTATXXTA)X)\displaystyle=\operatorname{Tr}((X^{T}A^{T}XX^{T}A)X)
=Tr(X(XTATXXTA))\displaystyle=\operatorname{Tr}(X(X^{T}A^{T}XX^{T}A))
=Tr((XXT)(ATXXTA))\displaystyle=\operatorname{Tr}((XX^{T})(A^{T}XX^{T}A))
=XXT,ATXXTA\displaystyle=\langle XX^{T},A^{T}XX^{T}A\rangle
XXTATXXTA\displaystyle\leq\|XX^{T}\|\|A^{T}XX^{T}A\|
=XXTXXT\displaystyle=\|XX^{T}\|\|XX^{T}\|
=XXT2\displaystyle=\|XX^{T}\|^{2}
=i,j=1nxi,xj2,\displaystyle=\sum_{i,j=1}^{n}\langle x_{i},x_{j}\rangle^{2},

as desired. We have equality if and only if we obtain equality in the Cauchy-Schwarz inequality, which happens if and only if XXTXX^{T} and ATXXTAA^{T}XX^{T}A are nonnegative multiples of each other. Since ATXXTA=XXT\|A^{T}XX^{T}A\|=\|XX^{T}\|, we have equality if and only if ATXXTA=XXTA^{T}XX^{T}A=XX^{T}, that is, when XXTA=AXXTXX^{T}A=AXX^{T}.

Lastly, suppose we have ATXXTA=XXTA^{T}XX^{T}A=XX^{T}. Multiplying XTX^{T} on the left to this equation yields XTXXT=XTATXXTA=XTATX(ATX)TX^{T}XX^{T}=X^{T}A^{T}XX^{T}A=X^{T}A^{T}X(A^{T}X)^{T}. This shows that the row space of XTXXTX^{T}XX^{T} is contained in the the row space of (ATX)T(A^{T}X)^{T}. By Proposition 2.1, this means the row space of XTX^{T} is contained in the row space of (ATX)T(A^{T}X)^{T}, so Span{x1,,xn}Span{AT(x1),,AT(xn)}\operatorname{Span}\{x_{1},\dots,x_{n}\}\subset\operatorname{Span}\{A^{T}(x_{1}),\dots,A^{T}(x_{n})\}. Since they have the same dimension, we have Span{x1,,xn}=Span{AT(x1),,AT(xn)}\operatorname{Span}\{x_{1},\dots,x_{n}\}=\operatorname{Span}\{A^{T}(x_{1}),\dots,A^{T}(x_{n})\}, so Span{x1,,xn}\operatorname{Span}\{x_{1},\dots,x_{n}\} is ATA^{T}-invariant. Since AA is orthogonal, AT=A1A^{T}=A^{-1}, AA-invariant spaces and ATA^{T}-invariant spaces coincide. ∎

We also have the following nice result.

Proposition 2.3.

Let x1,xnx_{1},\dots x_{n} be vectors such that dimSpan{x1,,xn}=m\dim\operatorname{Span}\{x_{1},\dots,x_{n}\}=m. Then

mi,j=1nxi,xj2i,j=1n|xi|2|xj|2,m\sum_{i,j=1}^{n}\langle x_{i},x_{j}\rangle^{2}\geq\sum_{i,j=1}^{n}|x_{i}|^{2}|x_{j}|^{2},

and we have equality if and only if all the xix_{i}’s have the same length and are mutually orthogonal.

Proof.

Without loss of generality we can assume x1,xnmx_{1},\dots x_{n}\in\mathbb{R}^{m}. Let XX be the associated m×nm\times n matrix

X=[x1xn].X=\begin{bmatrix}x_{1}&\dots&x_{n}\end{bmatrix}.

Then XXT2=XTX2=i,j=1nxi,xj2\|XX^{T}\|^{2}=\|X^{T}X\|^{2}=\sum_{i,j=1}^{n}\langle x_{i},x_{j}\rangle^{2}, and Tr(XXT)=Tr(XTX)=i=1n|xi|2\operatorname{Tr}(XX^{T})=\operatorname{Tr}(X^{T}X)=\sum_{i=1}^{n}|x_{i}|^{2}.

Using Cauchy-Schwarz inequality, we have

mi,j=1nxi,xj2\displaystyle m\sum_{i,j=1}^{n}\langle x_{i},x_{j}\rangle^{2} =Idm2XXT2\displaystyle=\|\operatorname{Id}_{m}\|^{2}\|XX^{T}\|^{2}
Idm,XXT2\displaystyle\geq\langle\operatorname{Id}_{m},XX^{T}\rangle^{2}
=Tr(XXT)2\displaystyle=\operatorname{Tr}(XX^{T})^{2}
=(i=1n|xi|2)2\displaystyle=(\sum_{i=1}^{n}|x_{i}|^{2})^{2}
=i,j=1n|xi|2|xj|2.\displaystyle=\sum_{i,j=1}^{n}|x_{i}|^{2}|x_{j}|^{2}.

To have equality, we need XXTXX^{T} to be a scalar multiple of the identity. By Proposition 2.1, XTXX^{T}X is a diagonalizable matrix with all eigenvalues equal. Thus it must also be a scalar multiple of the identity. This happens if and only if all the xix_{i}’s have the same length and are mutually orthogonal. ∎

3. Main results

3.1. Preliminaries

3.1.1. Second variation formula

Let Φ:(Σ,g)(M,g¯)\Phi:(\Sigma,g)\rightarrow(M,\bar{g}) be a minimal isometric immersion of an nn-dimensional compact Riemannian manifold (Σ,g)(\Sigma,g). Let C(ΦTM))C^{\infty}(\Phi^{*}TM)) denote the space of all CC^{\infty}-vector fields along Φ\Phi. For any VV in C(ΦTM)C^{\infty}(\Phi^{*}TM), let {Φt}\{\Phi_{t}\} be a CC^{\infty}-one-parameter family of immersions of Σ\Sigma into MM such that Φ0=Φ\Phi_{0}=\Phi and ddtΦt(x)|t=0=Vx\frac{d}{dt}\Phi_{t}(x)|_{t=0}=V_{x} for each xΣx\in\Sigma. Then we have the classical second variational formula

Theorem 3.1 (Second Variation formula).

Under the above assumptions,

d2dt2|Φt(Σ)||t=0=ΣJΣVN,VN𝑑Σ\displaystyle\frac{d^{2}}{dt^{2}}|\Phi_{t}(\Sigma)|\biggr{|}_{t=0}=-\int_{\Sigma}\langle J_{\Sigma}V^{N},V^{N}\rangle d\Sigma,

where VNV^{N} the normal component of VV and JΣJ_{\Sigma} is the elliptic Jacobi operator on the normal bundle N(Σ)N(\Sigma) defined by

JΣ(X):=ΔX+(i=1nRM(X,ei)ei)+i,j=1nB(ei,ej),XB(ei,ej)\displaystyle J_{\Sigma}(X):=\Delta^{\perp}X+(\sum_{i=1}^{n}R^{M}(X,e_{i})e_{i})^{\perp}+\sum_{i,j=1}^{n}\langle B(e_{i},e_{j}),X\rangle B(e_{i},e_{j}),

and the normal Laplacian on the normal bundle N(Σ)N(\Sigma) is given by

ΔX=i=1n(eieiX(eiei)TX)\displaystyle\Delta^{\perp}X=\sum_{i=1}^{n}(\nabla^{\perp}_{e_{i}}\nabla^{\perp}_{e_{i}}X-\nabla^{\perp}_{(\nabla_{e_{i}}e_{i})^{T}}X).

Here, {e1,,en}\{e_{1},\ldots,e_{n}\} is an orthonormal basis of TΣT\Sigma, \nabla is the connection of MM, \nabla^{\perp} is the normal connection of Σ\Sigma in MM, BB is the second fundamental form of Σ\Sigma in MM, and RMR^{M} is the curvature tensor of MM.

We say a minimal immersion Φ\Phi is stable if for all VC(ΦΓ(M))V\in C^{\infty}(\Phi^{*}\Gamma(M)),

ΣJΣVN,VN𝑑Σ0.-\int_{\Sigma}\langle J_{\Sigma}V^{N},V^{N}\rangle\,d\Sigma\geq 0.

3.1.2. Immersion into a Riemannian product

The general setting of our problem is as follows. Let Φ:ΣM¯=M1×M2\Phi:\Sigma\to\bar{M}=M_{1}\times M_{2} be an isometric immersion, where Σ\Sigma is an nn-dimensional compact Riemannian manifold, M1M_{1} is an m1m_{1}-dimensional compact Riemannian submanifold, and M2M_{2} is an m2m_{2}-dimensional Riemannian manifold. Let d=m1+m2nd=m_{1}+m_{2}-n denote the codimension of Σ\Sigma in M1×M2M_{1}\times M_{2}. We further let f1:M1N1f_{1}:M_{1}\to\mathbb{R}^{N_{1}} be an isometric embedding of M1M_{1} in an Euclidean space, and let F:M1×M2N1×M2F:M_{1}\times M_{2}\to\mathbb{R}^{N_{1}}\times M_{2} be defined by F=f1×IdM2F=f_{1}\times\operatorname{Id}_{M_{2}}. Let B1B_{1} denote the second fundamental form of the immersion f1f_{1}. We can then consider the chain of immersions ΣΦM1×M2𝐹N1×M2\Sigma\xrightarrow{\Phi}M_{1}\times M_{2}\xrightarrow{F}\mathbb{R}^{N_{1}}\times M_{2}.

Let pΣp\in\Sigma be a point. We choose a local orthonormal frame {e1,,en,η1,,ηd}\{e_{1},\dots,e_{n},\eta_{1},\dots,\eta_{d}\} near Φ(p)M1×M2\Phi(p)\in M_{1}\times M_{2} such that {e1,,en}\{e_{1},\dots,e_{n}\} is an orthonormal frame in dΦ(TΣ)d\Phi(T\Sigma) and {η1,,ηd}\{\eta_{1},\dots,\eta_{d}\} is an orthonormal frame in NΣT(M1×M2)N\Sigma\subset T(M_{1}\times M_{2}). For any tangent vector vv at Φ(p)=(p1,p2)M1×M2\Phi(p)=(p_{1},p_{2})\in M_{1}\times M_{2}, we decompose vv as v=(v1,v2)v=(v^{1},v^{2}), where viv^{i} is tangent to MiM_{i} at pip_{i}. We will thus write ei=(ei1,ei2)e_{i}=(e_{i}^{1},e_{i}^{2}) and ηβ=(ηβ1,ηβ2)\eta_{\beta}=(\eta_{\beta}^{1},\eta_{\beta}^{2}).

For a fixed vector UN1U\in\mathbb{R}^{N_{1}}, by identifying UU with (U,0)TF(Φ(p))(N1×M2)(U,0)\in T_{F(\Phi(p))}(\mathbb{R}^{N_{1}}\times M_{2}), we can decompose UU as

U=UT+UN,U=U^{T}+U^{N},

where UTdF(TΦ(p)(M1×M2))U^{T}\in dF(T_{\Phi(p)}(M_{1}\times M_{2})), and UNNΦ(p)(M1×M2)U^{N}\in N_{\Phi(p)}(M_{1}\times M_{2}). For UTdF(TΦ(p)(M1×M2))U^{T}\in dF(T_{\Phi(p)}(M_{1}\times M_{2})), by identifying it with its preimage in TΦ(p)(M1×M2)T_{\Phi(p)}(M_{1}\times M_{2}) and considering the immersion Φ:ΣM1×M2\Phi:\Sigma\to M_{1}\times M_{2}, we can further decompose it as

UT=TU+NU,U^{T}=T_{U}+N_{U},

where TUdΦ(TpΣ)T_{U}\in d\Phi(T_{p}\Sigma) and NUNpΣTΦ(p)(M1×M2)N_{U}\in N_{p}\Sigma\subset T_{\Phi(p)}(M_{1}\times M_{2}).

The case that will be interesting to us is when Σ\Sigma is a stable minimal submanifold of M1×M2M_{1}\times M_{2}. To study this case, we need to use some appropriate normal sections in the second variation formula. To this end, let {E1,,EN1}\{E_{1},\dots,E_{N_{1}}\} be an orthonormal basis of N1\mathbb{R}^{N_{1}}. Then NENAN_{E_{N_{A}}}, A=1,,N1A=1,\dots,N_{1} will be the normal sections we use. Computation shows that

Lemma 3.2.

[CW13, Equation (2.8)]

A=1N1NEA,JΣ(NEA)=i=1nβ=1d(B1(ηβ1,ηβ1),B1(ei1,ei1)2B1(ei1,ηβ1)2).\sum_{A=1}^{N_{1}}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle=\sum_{i=1}^{n}\sum_{\beta=1}^{d}\Big{(}\langle B_{1}(\eta_{\beta}^{1},\eta_{\beta}^{1}),B_{1}(e_{i}^{1},e_{i}^{1})\rangle-2\|B_{1}(e_{i}^{1},\eta_{\beta}^{1})\|^{2}\Big{)}.

In the next three subsections, we will let M1M_{1} be Pm1/2\mathbb{C}P^{m_{1}/2}, Pm1/4\mathbb{H}P^{m_{1}/4}, and 𝕆P2\mathbb{O}P^{2} respectively, and study the stable compact minimal immersions ΣM1×M2\Sigma\to M_{1}\times M_{2}. For these projective spaces, there exist isotropic embeddings into Euclidean space so that B1(X,X)\|B_{1}(X,X)\| is constant for any unit vector XTM1X\in TM_{1}. In all these cases, we will show A=1N1NEA,JΣ(NEA)0\sum_{A=1}^{N_{1}}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle\equiv 0, and therefore obtain geometric information about Σ\Sigma.

3.2. Stable minimal submanifolds in Pm12×Pm22\mathbb{C}P^{\frac{m_{1}}{2}}\times\mathbb{C}P^{\frac{m_{2}}{2}}.

We equip each Pmk/2\mathbb{C}P^{m_{k}/2}, k=1,2k=1,2 with the Fubini-Study metric of constant holomorphic sectional curvature λk2:=2mkmk+2\lambda_{k}^{2}:=\frac{2m_{k}}{m_{k}+2}.

Let Φ=(ϕ1,ϕ2):ΣM¯=Pm1/2×M2\Phi=(\phi_{1},\phi_{2})\colon\Sigma\to\bar{M}=\mathbb{C}P^{m_{1}/2}\times M_{2} be a stable compact minimal immersion of codimension dd and dimension nn.

Let f1:Pm12m~1f_{1}:\mathbb{C}P^{\frac{m_{1}}{2}}\rightarrow\mathbb{R}^{\tilde{m}_{1}} be the isotropic embedding induced by the generalized Veronese embedding (see Section 2 of [Sak77]). The nice property about this embedding is that, if we let BB denote its second fundamental form, then at each qPm12q\in\mathbb{C}P^{\frac{m_{1}}{2}}, we have

B(X,X)2=λ12 for any unit vector XTqPm12\|B(X,X)\|^{2}=\lambda_{1}^{2}\text{ for any unit vector $X\in T_{q}\mathbb{C}P^{\frac{m_{1}}{2}}$}

(see Proposition 3.6 of [Ohn86]).

For each vm~1v\in\mathbb{R}^{\tilde{m}_{1}}, we identify vv with (v,0)T(m~1×M2)(v,0)\in T(\mathbb{R}^{\tilde{m}_{1}}\times M_{2}). Same as in Section 3.1.2, we denote by NvN_{v} the projection of vv onto the normal space NpΣN_{p}\Sigma in TΦ(p)M¯T_{\Phi(p)}\bar{M}. We are now going to use the normal sections NvN_{v} in the second variation formula, where vv runs over an orthonormal basis of the Euclidean space m~1\mathbb{R}^{\tilde{m}_{1}}.

Let pΣp\in\Sigma. Let {e1,,en}\{e_{1},\dots,e_{n}\} be an orthonormal basis of dΦ(TpΣ)d\Phi(T_{p}\Sigma). Let ej1e_{j}^{1} be the projection of eje_{j} onto Tϕ1(p)Pm1/2T_{\phi_{1}(p)}\mathbb{C}P^{m_{1}/2} and ej2e_{j}^{2} be the projection of eje_{j} onto Tϕ2(p)M2T_{\phi_{2}(p)}M_{2}. Let J1J_{1} be the complex structure of Tϕ1(p)Pm1/2T_{\phi_{1}(p)}\mathbb{C}P^{m_{1}/2}.

Lemma 3.3.

[RL22, Lemma 3.1] Same assumptions as above. Let {E1,,Em~1}\{E_{1},\ldots,E_{\tilde{m}_{1}}\} be the usual canonical basis of m~1\mathbb{R}^{\tilde{m}_{1}}. Then

(1) A=1m~1NEA,JΣ(NEA)=λ12(i,j=1nJ1(ej1),ei12ej1,ei12).\displaystyle\sum_{A=1}^{\tilde{m}_{1}}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle=\lambda_{1}^{2}\Big{(}\sum_{i,j=1}^{n}\langle J_{1}(e^{1}_{j}),e_{i}^{1}\rangle^{2}-\langle e^{1}_{j},e^{1}_{i}\rangle^{2}\Big{)}.

The stability of Φ\Phi requires

ΣNEA,JΣ(NEA)dΣ0for each A=1,,m~1,\int_{\Sigma}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle\,d\Sigma\geq 0\quad\text{for each }A=1,\dots,\tilde{m}_{1},

so

ΣA=1m~1NEA,JΣ(NEA)dΣ0.\int_{\Sigma}\sum_{A=1}^{\tilde{m}_{1}}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle\,d\Sigma\geq 0.

On the other hand, combining Lemma 3.3 and Proposition 2.2, we have

A=1m~1NEA,JΣ(NEA)=λ12i,j=1nJ1(ej1),ei12ej1,ei120.\displaystyle\sum_{A=1}^{\tilde{m}_{1}}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle=\lambda_{1}^{2}\sum_{i,j=1}^{n}\langle J_{1}(e_{j}^{1}),e_{i}^{1}\rangle^{2}-\langle e_{j}^{1},e_{i}^{1}\rangle^{2}\leq 0.

The above two equations together show that the integrand must be exactly zero, namely,

(2) i,j=1nJ1(ej1),ei12ej1,ei12=0.\sum_{i,j=1}^{n}\langle J_{1}(e_{j}^{1}),e_{i}^{1}\rangle^{2}-\langle e_{j}^{1},e_{i}^{1}\rangle^{2}=0.

Now we specialize to the case where M2=Pm2/2M_{2}=\mathbb{C}P^{m_{2}/2}, and consider the stable compact minimal immersion Φ=(ϕ1,ϕ2):ΣPm1/2×Pm2/2\Phi=(\phi_{1},\phi_{2})\colon\Sigma\to\mathbb{C}P^{m_{1}/2}\times\mathbb{C}P^{m_{2}/2} of dimension nn. For pΣp\in\Sigma, let J2J_{2} be the complex structure of Tϕ2(p)Pm2/2T_{\phi_{2}(p)}\mathbb{C}P^{m_{2}/2}. Then letting Pm2/2\mathbb{C}P^{m_{2}/2} play the role of Pm1/2\mathbb{C}P^{m_{1}/2} in the above arguments, we have

(3) i,j=1nJ2(ej2),ei22ej2,ei22=0.\sum_{i,j=1}^{n}\langle J_{2}(e_{j}^{2}),e_{i}^{2}\rangle^{2}-\langle e_{j}^{2},e_{i}^{2}\rangle^{2}=0.

as well.

For k=1,2k=1,2, choose a local orthonormal basis {vik}i=1mk\{v_{i}^{k}\}_{i=1}^{m_{k}} for Tϕk(p)Pmk/2T_{\phi_{k}(p)}\mathbb{C}P^{m_{k}/2} and write ejke_{j}^{k} under this basis as ejk=bi,jkvie_{j}^{k}=b^{k}_{i,j}v_{i}. Let XkX_{k} denote the mk×nm_{k}\times n matrix Xk=[bi,jk]i,jX_{k}=\begin{bmatrix}b^{k}_{i,j}\end{bmatrix}_{i,j}. Then X1TX1=[ei1,ej1i,j=1n]X_{1}^{T}X_{1}=\begin{bmatrix}\langle e_{i}^{1},e_{j}^{1}\rangle_{i,j=1}^{n}\end{bmatrix} and X2TX2=[ei2,ej2i,j=1n]X_{2}^{T}X_{2}=\begin{bmatrix}\langle e_{i}^{2},e_{j}^{2}\rangle_{i,j=1}^{n}\end{bmatrix}, and since {(e11,e12),,(en1,en2)}\{(e_{1}^{1},e_{1}^{2}),\dots,(e_{n}^{1},e_{n}^{2})\} forms an orthonormal basis of dΦ(TpΣ)d\Phi(T_{p}\Sigma) we have

X1TX1+X2TX2=In.X_{1}^{T}X_{1}+X_{2}^{T}X_{2}=I_{n}.

Notice that X1TX1X_{1}^{T}X_{1} and X2TX2X_{2}^{T}X_{2} are positive semi-definite since they are Gram matrices. Moreover, if μ\mu is an eigenvalue of X1TX1X_{1}^{T}X_{1} with eigenvector vv, then we have X1TX1v+X2TX2v=vX_{1}^{T}X_{1}v+X_{2}^{T}X_{2}v=v, so X2TX2v=vX1TX1v=vμv=(1μ)vX_{2}^{T}X_{2}v=v-X_{1}^{T}X_{1}v=v-\mu v=(1-\mu)v, showing (1μ)(1-\mu) is an eigenvalue of X2TX2X_{2}^{T}X_{2} with eigenvector vv.
Thus if 1μ1μn01\geq\mu_{1}\geq\dots\geq\mu_{n}\geq 0 are eigenvalues of X1TX1X_{1}^{T}X_{1} in non-increasing order with corresponding orthonormal eigenbasis v1,,vnv_{1},\dots,v_{n}, then 01μ11μn10\leq 1-\mu_{1}\leq\dots\leq 1-\mu_{n}\leq 1 are eigenvalues of X2TX2X_{2}^{T}X_{2} in non-decreasing order with corresponding orthonormal eigenbasis v1,,vnv_{1},\dots,v_{n}. In particular, the multiplicity of 11 as an eigenvalue of X1TX1X_{1}^{T}X_{1} equals the multiplicity of 0 as an eigenvalue of X2TX2X_{2}^{T}X_{2}.

By Proposition 2.2, Equation (2) implies that X1X1TX_{1}X_{1}^{T} commutes with J1J_{1}, and Equation (3) implies that X2X2TX_{2}X_{2}^{T} commutes with J2J_{2}. Thus the eigenspaces of X1X1TX_{1}X_{1}^{T} are J1J_{1}-invariant, therefore even dimensional. Thus all the eigenvalues of X1X1TX_{1}X_{1}^{T} have even multiplicity. By Proposition 2.1, this shows all the nonzero eigenvalues of X1TX1X_{1}^{T}X_{1} have even multiplicity. Similarly, all the nonzero eigenvalues of X2TX2X_{2}^{T}X_{2} have even multiplicity.
In particular, 11 as an eigenvalue of X2TX2X_{2}^{T}X_{2} has even multiplicity. Using the fact that the multiplicity of 11 as an eigenvalue of X1TX1X_{1}^{T}X_{1} equals the multiplicity of 0 as an eigenvalue of X2TX2X_{2}^{T}X_{2}, we get that the multiplicity of 0 as an eigenvalue of X1TX1X_{1}^{T}X_{1} is also even. Thus all the eigenvalues of X1TX1X_{1}^{T}X_{1} have even multiplicity. However, sum of the multiplicities of the eigenvalues of X1TX1X_{1}^{T}X_{1} equals its dimension, which is nn. This shows nn must be an even number.

We have thus proved Theorem 1.2.

Remark 3.4.

Notice that by Proposition 2.2, from Equations (2) and (3) we can deduce that Span{e1k,,enk}\operatorname{Span}\{e^{k}_{1},\dots,e^{k}_{n}\} is JkJ_{k}-invariant, for k=1,2k=1,2. However, this is not enough to conclude that the tangent space of the stable submanifold Σ\Sigma behaves well under a complex structure of Pm1/2×Pm2/2\mathbb{C}P^{m_{1}/2}\times\mathbb{C}P^{m_{2}/2}.

3.3. Stable minimal submanifolds in Pm14×M\mathbb{H}P^{\frac{m_{1}}{4}}\times M.

Equip Pm14\mathbb{H}P^{\frac{m_{1}}{4}} with its standard metric as a Riemannian symmetric space, with the maximum λ2\lambda^{2} of the sectional curvatures given by λ2:=2m1m1+4\lambda^{2}:=\frac{2m_{1}}{m_{1}+4}. Let Φ=(ϕ1,ϕ2):ΣM¯:=Pm14×M\Phi=(\phi_{1},\phi_{2}):\Sigma\rightarrow\bar{M}:=\mathbb{H}P^{\frac{m_{1}}{4}}\times M be a stable compact minimal immersion of codimension dd and dimension nn, where MM is any Riemannian manifold of dimension m2m_{2}.

Let f1:Pm14mf_{1}:\mathbb{H}P^{\frac{m_{1}}{4}}\rightarrow\mathbb{R}^{m} be the isotropic embedding induced by the generalized Veronese embedding (see Section 2 of [Sak77]). The nice property about this embedding is that, if we let BB denote its the second fundamental form, then at each qPm14q\in\mathbb{H}P^{\frac{m_{1}}{4}}, we have

B(X,X)2=λ2 for any unit vector XTqPm14\|B(X,X)\|^{2}=\lambda^{2}\text{ for any unit vector $X\in T_{q}\mathbb{H}P^{\frac{m_{1}}{4}}$}

(see Proposition 3.6 of [Ohn86]).

For each vmv\in\mathbb{R}^{m}, we identify vv with (v,0)T(m×M)(v,0)\in T(\mathbb{R}^{m}\times M). Same as in Section 3.1.2, we denote by NvN_{v} the projection of vv onto the normal space NpΣN_{p}\Sigma in TΦ(p)M¯T_{\Phi(p)}\bar{M}. As in the complex case, we are going to use the normal sections NvN_{v} in the second variation formula, where vv runs over an orthonormal basis of m\mathbb{R}^{m}.

Let pΣp\in\Sigma. Let {e1,,en}\{e_{1},\ldots,e_{n}\} be an orthonormal basis of dΦ(TpΣ)d\Phi(T_{p}\Sigma), and let {η1,,ηd}\{\eta_{1},\ldots,\eta_{d}\} be an orthonormal basis of NpΣN_{p}\Sigma. Let ej1e_{j}^{1} be the projection of eje_{j} onto Tϕ1(p)Pm1/4T_{\phi_{1}(p)}\mathbb{H}P^{m_{1}/4} and ej2e_{j}^{2} be the projection of eje_{j} onto Tϕ2(p)MT_{\phi_{2}(p)}M. Similarly, let ηj1\eta_{j}^{1} be the projection of ηj\eta_{j} onto Tϕ1(p)Pm1/4T_{\phi_{1}(p)}\mathbb{H}P^{m_{1}/4} and ηj2\eta_{j}^{2} be the projection of ηj\eta_{j} onto Tϕ2(p)MT_{\phi_{2}(p)}M. Let {J1,J2,J3}\{J_{1},J_{2},J_{3}\} be the quaternionic structure on Tϕ1(p)Pm1/4T_{\phi_{1}(p)}\mathbb{H}P^{m_{1}/4}.

Lemma 3.5.

[RL22, Lemma 4.1] Same assumptions as above. Let {E1,,Em}\{E_{1},\ldots,E_{m}\} be the usual canonical basis of m\mathbb{R}^{m}. Then, for s{1,2,3}s\in\{1,2,3\}

(4) A=1mNEA,JΣ(NEA)=λ2(ks3j=1nβ=1dJk(ηβ1),ej12+i,j=1n(Js(ei1),ej12ei1,ej12)).\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle=\lambda^{2}\Big{(}-\sum_{k\neq s}^{3}\sum_{j=1}^{n}\sum_{\beta=1}^{d}\langle J_{k}(\eta^{1}_{\beta}),e^{1}_{j}\rangle^{2}+\sum_{i,j=1}^{n}(\langle J_{s}(e^{1}_{i}),e_{j}^{1}\rangle^{2}-\langle e^{1}_{i},e_{j}^{1}\rangle^{2})\Big{)}.

The stability of Φ\Phi requires

ΣNEA,JΣ(NEA)dΣ0for each A=1,,m,\int_{\Sigma}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle\,d\Sigma\geq 0\quad\text{for each }A=1,\dots,m,

so

(5) ΣA=1mNEA,JΣ(NEA)dΣ0.\int_{\Sigma}\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle\,d\Sigma\geq 0.

On the other hand, combining Lemma 3.5 and Proposition 2.2, we have for s=1,2,3s=1,2,3

(6) A=1mNEA,JΣ(NEA)\displaystyle\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle
=\displaystyle= λ2(ks3j=1nβ=1dJk(ηβ1),ej12+i,j=1n(Js(ei1),ej12ei1,ej12))\displaystyle\,\lambda^{2}\Big{(}-\sum_{k\neq s}^{3}\sum_{j=1}^{n}\sum_{\beta=1}^{d}\langle J_{k}(\eta^{1}_{\beta}),e^{1}_{j}\rangle^{2}+\sum_{i,j=1}^{n}(\langle J_{s}(e^{1}_{i}),e_{j}^{1}\rangle^{2}-\langle e^{1}_{i},e_{j}^{1}\rangle^{2})\Big{)}
\displaystyle\leq λ2i,j=1n(Js(ei1),ej12ei1,ej12)\displaystyle\,\lambda^{2}\sum_{i,j=1}^{n}(\langle J_{s}(e^{1}_{i}),e_{j}^{1}\rangle^{2}-\langle e^{1}_{i},e_{j}^{1}\rangle^{2})
\displaystyle\leq  0.\displaystyle\,0.

Equations (5) and (6) together show that both inequalities in (6) are equalities. Hence for s=1,2,3s=1,2,3,

(7) i,j=1nJs(ej1),ei12ej1,ei12=0,\sum_{i,j=1}^{n}\langle J_{s}(e_{j}^{1}),e_{i}^{1}\rangle^{2}-\langle e_{j}^{1},e_{i}^{1}\rangle^{2}=0,

and consequently

(8) ks3j=1nβ=1dJk(ηβ1),ej12=0.-\sum_{k\neq s}^{3}\sum_{j=1}^{n}\sum_{\beta=1}^{d}\langle J_{k}(\eta^{1}_{\beta}),e^{1}_{j}\rangle^{2}=0.

From Equation (8) we have that for t{1,2,3}t\in\{1,2,3\}, β{1,,d}\beta\in\{1,\dots,d\}, and j{1,,n}j\in\{1,\dots,n\}

Jt(ηβ1),ej1=0.\langle J_{t}(\eta_{\beta}^{1}),e_{j}^{1}\rangle=0.

Thus for t,βt,\beta fixed, Jt(ηβ1)Span{el1}lJ_{t}(\eta_{\beta}^{1})\perp\operatorname{Span}\{e_{l}^{1}\}_{l}. By Proposition 2.2, Equation (7) also implies that Span{el1}l\operatorname{Span}\{e_{l}^{1}\}_{l} is JtJ_{t}-invariant, so Jt(ej1)Span{el1}lJ_{t}(e_{j}^{1})\in\operatorname{Span}\{e_{l}^{1}\}_{l} for each jj. Thus

(9) ηβ1,ej1=Jt(ηβ1),Jt(ej1)=0.\langle\eta_{\beta}^{1},e_{j}^{1}\rangle=\langle J_{t}(\eta_{\beta}^{1}),J_{t}(e_{j}^{1})\rangle=0.

But since we have 0=ηβ,ej=ηβ1,ej1+ηβ2,ej20=\langle\eta_{\beta},e_{j}\rangle=\langle\eta_{\beta}^{1},e_{j}^{1}\rangle+\langle\eta_{\beta}^{2},e_{j}^{2}\rangle, Equation (9) shows that we necessarily have

(10) ηβ2,ej2=0\langle\eta_{\beta}^{2},e_{j}^{2}\rangle=0

as well.

For the product manifold Pm1/4×M\mathbb{H}P^{m_{1}/4}\times M, let

P:T(Pm1/4×M)TPm1/40T(Pm1/4×M),P:T(\mathbb{H}P^{m_{1}/4}\times M)\to T\mathbb{H}P^{m_{1}/4}\oplus 0\subseteq T(\mathbb{H}P^{m_{1}/4}\times M),
Q:T(Pm1/4×M)0TMT(Pm1/4×M)Q:T(\mathbb{H}P^{m_{1}/4}\times M)\to 0\oplus TM\subseteq T(\mathbb{H}P^{m_{1}/4}\times M)

denote the projection maps. Let F=PQF=P-Q, then F2=IF^{2}=I. For pΣp\in\Sigma, we claim that F(dΦ(TpΣ))dΦ(TpΣ)F(d\Phi(T_{p}{\Sigma}))\subseteq d\Phi(T_{p}{\Sigma}). In fact, since F(ei)=F(ei1,ei2)=(ei1,ei2)F(e_{i})=F(e_{i}^{1},e_{i}^{2})=(e_{i}^{1},-e_{i}^{2}), for each ηβ\eta_{\beta}, we have

F(ei),ηβ=ei1,ηβ1ei2,ηβ2=0,\langle F(e_{i}),\eta_{\beta}\rangle=\langle e_{i}^{1},\eta_{\beta}^{1}\rangle-\langle e_{i}^{2},\eta_{\beta}^{2}\rangle=0,

where we have used Equations (9) and (10) in the second equality. Therefore, F(ei)(dΦ(TpΣ))F(e_{i})\perp(d\Phi(T_{p}\Sigma))^{\perp}. This shows F(ei)dΦ(TpΣ)F(e_{i})\subset d\Phi(T_{p}\Sigma), so F(dΦ(TpΣ))dΦ(TpΣ)F(d\Phi(T_{p}{\Sigma}))\subseteq d\Phi(T_{p}{\Sigma}). Thus Σ\Sigma is an invariant submanifold of Pm1/4×M\mathbb{H}P^{m_{1}/4}\times M.

By Theorem 1 in [XN00], Σ\Sigma is isometric to a product manifold Σ1×Σ2\Sigma_{1}\times\Sigma_{2}, where M1M_{1} is an immersion into Pm1/4\mathbb{H}P^{m_{1}/4} and Σ2\Sigma_{2} is an immersion into MM. Since Φ:ΣPm1/4×M\Phi:\Sigma\to\mathbb{H}P^{m_{1}/4}\times M is a stable compact minimal immersion, we further have that Σ1\Sigma_{1} and Σ2\Sigma_{2} are stable compact minimal immersions into Pm1/4\mathbb{H}P^{m_{1}/4} and MM, respectively. However, Ohnita [Ohn86, Theorem D] showed that the only stable minimal immersed submanifolds of a quaternionic protective subspace are precisely the quaternionic protective subspaces. Using this result, we have thus established Theorem 1.2.

Remark 3.6.

Notice that the diagonal map Φ:Pm/4Pm/4×Pm/4,x(x,x)\Phi:\mathbb{H}P^{m/4}\to\mathbb{H}P^{m/4}\times\mathbb{H}P^{m/4},x\mapsto(x,x) is not stable.

3.4. Stable minimal submanifolds in 𝕆P2×M\mathbb{O}P^{2}\times M.

Let 𝕆P2\mathbb{O}P^{2} denote the octonionic (Cayley) projective plane, endowed with the standard metric as a Riemannian symmetric space. Classically, 𝕆P2\mathbb{O}P^{2} can be seen as a 16-dimensional quotient manifold F4/Spin(9)F_{4}/\operatorname{Spin}(9), where the isometry group of 𝕆P2\mathbb{O}P^{2} is isomophic to the exceptional Lie group F4F_{4} and the isotropy group at any point q𝕆P2q\in\mathbb{O}P^{2} is isomorphic to Spin(9)\operatorname{Spin}(9).

Alternatively, we can also define 𝕆P2\mathbb{O}P^{2} in terms of equivalence classes (cf. [HSV09]). Consider the relation \sim on 𝕆3\mathbb{O}^{3}, where [a,b,c][d,e,f][a,b,c]\sim[d,e,f] if and only if there exists u𝕆{0}u\in\mathbb{O}\setminus\{0\} such that a=du,b=eu,c=fua=du,b=eu,c=fu. This relation is symmetric and reflexive. However, it is not necessarily transitive due to non-associativity of octonions. As a remedy, we instead consider the following subsets of 𝕆3\mathbb{O}^{3}:

U1:={1}×𝕆×𝕆,U2:=𝕆×{1}×𝕆,U3:=𝕆×𝕆×{1},U_{1}:=\{1\}\times\mathbb{O}\times\mathbb{O},\quad U_{2}:=\mathbb{O}\times\{1\}\times\mathbb{O},\quad U_{3}:=\mathbb{O}\times\mathbb{O}\times\{1\},

and their union 𝕆3:=U1U2U3\mathbb{O}^{3}_{\bullet}:=U_{1}\cup U_{2}\cup U_{3}. The relation \sim on 𝕆3\mathbb{O}^{3}_{\bullet} is then an equivalence relation, and we can define the octonionic projective plane 𝕆P2\mathbb{O}P^{2} as the set of equivalence classes of 𝕆3\mathbb{O}^{3}_{\bullet} by \sim. That is,

𝕆P2:=𝕆3/.\mathbb{O}P^{2}:=\mathbb{O}^{3}_{\bullet}/\sim.

3.4.1. Basic octonion algebra

First we review some basic octonion algebra. We refer the reader to [BG72], [HSV09], and [Kot20] for details. Let I0=1,I1,,I7I_{0}=1,I_{1},\dots,I_{7} denote the usual octonion basis. For a=a0+s=17asIs𝕆a=a_{0}+\sum_{s=1}^{7}a_{s}I_{s}\in\mathbb{O}, aia_{i}\in\mathbb{R}, let

a:=a0s=17asIsa^{*}:=a_{0}-\sum_{s=1}^{7}a_{s}I_{s}

denote its conjugate. Then the inner product on 𝕆\mathbb{O} is given by

a,b=12(ab+ba)=12(ba+ab).\langle a,b\rangle=\frac{1}{2}(a^{*}b+b^{*}a)=\frac{1}{2}(ba^{*}+ab^{*}).

Moreover, for a,b,c,d,x,y𝕆a,b,c,d,x,y\in\mathbb{O}, we have the identities

(11) ax,y\displaystyle\langle ax,y\rangle =x,ay,\displaystyle=\langle x,a^{*}y\rangle,
(12) xa,y\displaystyle\langle xa,y\rangle =x,ya,\displaystyle=\langle x,ya^{*}\rangle,
(13) ab,cd+ad,cb\displaystyle\langle ab,cd\rangle+\langle ad,cb\rangle =2a,cb,d.\displaystyle=2\langle a,c\rangle\langle b,d\rangle.

The 16-dimensional real vector space 𝕆𝕆\mathbb{O}\oplus\mathbb{O} decomposes into octonionic lines of the form

m:={(u,mu)u𝕆}or:={(0,u)u𝕆},\ell_{m}:=\{(u,mu)\mid u\in\mathbb{O}\}\quad\text{or}\quad\ell_{\infty}:=\{(0,u)\mid u\in\mathbb{O}\},

that intersect each other only at the origin (0,0)𝕆𝕆(0,0)\in\mathbb{O}\oplus\mathbb{O}. Each line is an 8-dimensional real vector space. Here m𝕆{}𝕆P1S8m\in\mathbb{O}\cup\{\infty\}\cong\mathbb{O}P^{1}\cong S^{8} parametrizes the set of octonionic lines. Be cautious here that the octonionic line through (0,0)(0,0) and (a,b)(a,b) is not the set {(au,bu)u𝕆}\{(au,bu)\mid u\in\mathbb{O}\} due to the non-associativity of octonion multiplication; the correct line is given by ba1\ell_{ba^{-1}}.

Proposition 3.7 (see e.g.,[Kot20, Section 2.2]).

Here we collect the properties of Spin(9)\operatorname{Spin}(9) that we will use later.

  • The group Spin(9)\operatorname{Spin}(9) acts transitively and effectively on the unit sphere S15𝕆𝕆S^{15}\subset\mathbb{O}\oplus\mathbb{O}, and we can view Spin(9)SO(16)\operatorname{Spin}(9)\subset SO(16) as a matrix group acting on 𝕆𝕆\mathbb{O}\oplus\mathbb{O}.

  • Spin(9)\operatorname{Spin}(9) maps octonionic lines to octonionic lines.

  • The setwise stabilizer of an octonionic line is isomorphic to Spin(8)\operatorname{Spin}(8), and the stabilizer of a point is isomorphic to Spin(7)\operatorname{Spin}(7).

Define the map L:𝕆𝕆0𝕆P1L:\mathbb{O}\oplus\mathbb{O}\setminus 0\to\mathbb{O}P^{1} by mapping x𝕆𝕆x\in\mathbb{O}\oplus\mathbb{O} to the unique octonionic line L(x)L(x) that contains xx. The restriction of the map LL to S15𝕆𝕆S^{15}\subset\mathbb{O}\oplus\mathbb{O} defines the octonionic Hopf fibration

S15S8,or as homogeneous fibrationSpin(9)/Spin(7)Spin(9)/Spin(8)S^{15}\to S^{8},\quad\text{or as homogeneous fibration}\quad\operatorname{Spin}(9)/\operatorname{Spin}(7)\to\operatorname{Spin}(9)/\operatorname{Spin}(8)

with S7Spin(8)/Spin(7)S^{7}\cong\operatorname{Spin}(8)/\operatorname{Spin}(7) as the fiber. Spin(9)\operatorname{Spin}(9) is the symmetry group of this fibration. (cf. [OPPV13])

Lemma 3.8.

Let m1,m2\ell_{m_{1}},\ell_{m_{2}}, m1,m2𝕆{}m_{1},m_{2}\in\mathbb{O}\cup\{\infty\} be two octonionic lines defined as above. Let x1,,x8x_{1},\dots,x_{8} be an orthonormal basis of m1\ell_{m_{1}} and y1,,y8y_{1},\dots,y_{8} be an orthonormal basis of m2\ell_{m_{2}}. Let XX, YY be the associated 16×816\times 8 matrices

X=[x1x8],Y=[y1y8].X=\begin{bmatrix}x_{1}&\dots&x_{8}\end{bmatrix},\quad Y=\begin{bmatrix}y_{1}&\dots&y_{8}\end{bmatrix}.

Then XTY=[xi,yj]i,j=18=cQX^{T}Y=[\langle x_{i},y_{j}\rangle]_{i,j=1}^{8}=cQ, where QQ is an 8×88\times 8 orthogonal matrix, and 0c10\leq c\leq 1 is some constant. We have c=0c=0 if and only if m1m2\ell_{m_{1}}\perp\ell_{m_{2}}, and c=1c=1 if and only if m1=m2m_{1}=m_{2}.

Proof.

By Proposition 3.7, by an action of Spin(9)\operatorname{Spin}(9) we can assume that m1=0m_{1}=0 and consequently m1=𝕆0\ell_{m_{1}}=\mathbb{O}\oplus 0. For m2=m_{2}=\infty, since 0,\ell_{0},\ell_{\infty} are orthogonal complements, we have XTY=0=cQX^{T}Y=0=cQ, for c=0c=0 and Q=Id8Q=\operatorname{Id}_{8}. We can thus assume m2𝕆m_{2}\in\mathbb{O}.

Let x~s=(Is1,0)\tilde{x}_{s}=(I_{s-1},0), s=1,,8s=1,\dots,8, where I0=1,I1,,I7I_{0}=1,I_{1},\dots,I_{7} are the usual octonion basis. Then x~1,,x~8\tilde{x}_{1},\dots,\tilde{x}_{8} form an orthogonal basis for the 8-dimensional vector space 0\ell_{0}. Let

X~=[x~1x~8].\tilde{X}=\begin{bmatrix}\tilde{x}_{1}&\dots&\tilde{x}_{8}\end{bmatrix}.

Therefore, there exists an 8×88\times 8 orthogonal matrix Q1Q_{1} such that X=X~Q1X=\tilde{X}Q_{1}.

Let y~s=11+|m2|2(Is1,m2Is1)\tilde{y}_{s}=\frac{1}{\sqrt{1+|m_{2}|^{2}}}(I_{s-1},m_{2}I_{s-1}), s=1,,8s=1,\dots,8. Then y~1,,y~8\tilde{y}_{1},\dots,\tilde{y}_{8} form an orthogonal basis for the 8-dimensional vector space m2\ell_{m_{2}}. Let

Y~=[y~1y~8].\tilde{Y}=\begin{bmatrix}\tilde{y}_{1}&\dots&\tilde{y}_{8}\end{bmatrix}.

Therefore, there exists an 8×88\times 8 orthogonal matrix Q2Q_{2} such that Y=Y~Q2Y=\tilde{Y}Q_{2}.

It is clear to see that

X~TY~=[x~i,y~j]i,j=1n=11+|m2|2Id8.\tilde{X}^{T}\tilde{Y}=[\langle\tilde{x}_{i},\tilde{y}_{j}\rangle]_{i,j=1}^{n}=\frac{1}{\sqrt{1+|m_{2}|^{2}}}\operatorname{Id}_{8}.

Thus

XTY=Q1TX~TY~Q2=Q1T(11+|m2|2Id8)Q2=11+|m2|2(Q1TQ2),X^{T}Y=Q_{1}^{T}\tilde{X}^{T}\tilde{Y}Q_{2}=Q_{1}^{T}(\frac{1}{\sqrt{1+|m_{2}|^{2}}}\operatorname{Id}_{8})Q_{2}=\frac{1}{\sqrt{1+|m_{2}|^{2}}}(Q_{1}^{T}Q_{2}),

where Q:=Q1TQ2Q:=Q_{1}^{T}Q_{2} is an 8×88\times 8 orthogonal matrix, and c=11+|m2|2(0,1]c=\frac{1}{\sqrt{1+|m_{2}|^{2}}}\in(0,1]. We have c=0c=0 precisely when m2=0m_{2}=0 and consequently m1=m2\ell_{m_{1}}=\ell_{m_{2}}.

3.4.2. Geometry of 𝕆P2\mathbb{O}P^{2}

Here we review the basic Riemannian geometry of 𝕆P2\mathbb{O}P^{2}. We refer the reader to [BG72], [Ohn86], and [HSV09] for details. In each of the chart U1,U2,U3U_{1},U_{2},U_{3} defined at the beginning of this subsection, let (u,v)(u,v) with u,v𝕆u,v\in\mathbb{O} denote the coordinate functions on the chart. Then they give rise to octonion valued 1-forms dudu and dvdv. The standard metric of 𝕆P2\mathbb{O}P^{2} is then given by (see Equation (3.1) in [HSV09])

ds2=4λ2|du|2(1+|v|2)+|dv|2(1+|u|2)2Re[(uv)(dvdu)](1+|u|2+|v|2)2,ds^{2}=\frac{4}{\lambda^{2}}\frac{|du|^{2}(1+|v|^{2})+|dv|^{2}(1+|u|^{2})-2\operatorname{Re}[(uv^{*})(dvdu^{*})]}{(1+|u|^{2}+|v|^{2})^{2}},

where λ2\lambda^{2} is a scaling factor that denotes the maximum of the sectional curvatures of 𝕆P2\mathbb{O}P^{2}. Notice that this expression resembles the expression of the Fubini–Study metric of a complex projective plane.

Let RR denote the Riemannian curvature tensor of 𝕆P2\mathbb{O}P^{2}. Let qq be a point in 𝕆P2\mathbb{O}P^{2}. We can identify Tq(𝕆P2)T_{q}(\mathbb{O}P^{2}) with 𝕆𝕆\mathbb{O}\oplus\mathbb{O} in a natural manner. Using the structure of the octonionic algebra, the curvature tensor RR is given by

(14) R((a,b),(c,d))(e,f),(g,h)=λ24(\displaystyle\langle R\Big{(}(a,b),(c,d)\Big{)}(e,f),(g,h)\rangle=\frac{\lambda^{2}}{4}\bigg{(} 4a,ec,g+4c,ea,g\displaystyle-4\langle a,e\rangle\langle c,g\rangle+4\langle c,e\rangle\langle a,g\rangle
4b,fd,h+4d,fb,h\displaystyle-4\langle b,f\rangle\langle d,h\rangle+4\langle d,f\rangle\langle b,h\rangle
+ed,gbeb,gd+cf,ahaf,ch\displaystyle+\langle ed^{*},gb^{*}\rangle-\langle eb^{*},gd^{*}\rangle+\langle cf^{*},ah^{*}\rangle-\langle af^{*},ch^{*}\rangle
+adcb,gfeh),\displaystyle+\langle ad^{*}-cb^{*},gf^{*}-eh^{*}\rangle\bigg{)},

where a,b,c,d,e,f,g,h𝕆a,b,c,d,e,f,g,h\in\mathbb{O}. For this formula, see [HSV09, Theorem 5.3 (1)], or see [Ohn86, Section 1.2 and Table 2] and apply the basic identities (11) and (12). Notice that we use a slightly different identification for Tp(𝕆P2)T_{p}(\mathbb{O}P^{2}) from the one in [Ohn86] (𝕆𝕆\mathbb{O}\oplus\mathbb{O} versus 𝕆𝕆¯\mathbb{O}\oplus\overline{\mathbb{O}}) and there is a typo in Equation (1.3) of [Ohn86], where vv at the end of the second line should be vv^{*}. Here we use this identification so as to be consistent with out definition of octonionic lines earlier.

In particular, we have

(15) R((a,b),(c,d))(a,b),(c,d)=λ24(\displaystyle\langle R\Big{(}(a,b),(c,d)\Big{)}(a,b),(c,d)\rangle=\frac{\lambda^{2}}{4}\bigg{(} 4a,ac,c+4a,c24b,bd,d+4b,d2\displaystyle-4\langle a,a\rangle\langle c,c\rangle+4\langle a,c\rangle^{2}-4\langle b,b\rangle\langle d,d\rangle+4\langle b,d\rangle^{2}
+2ad,cb2ab,cdadcb,adcb).\displaystyle+2\langle ad^{*},cb^{*}\rangle-2\langle ab^{*},cd^{*}\rangle-\langle ad^{*}-cb^{*},ad^{*}-cb^{*}\rangle\bigg{)}.

Let f1:𝕆P2mf_{1}:\mathbb{O}P^{2}\rightarrow\mathbb{R}^{m} be the isotropic embedding induced by the generalized Veronese embedding (see Section 2 of [Sak77]). Let BB denote the second fundamental form of this embedding. Then at each q𝕆P2q\in\mathbb{\mathbb{O}}P^{2}, we have

B(X,X)2=λ2 for any unit vector X in Tq𝕆P2.\|B(X,X)\|^{2}=\lambda^{2}\text{ for any unit vector $X$ in $T_{q}\mathbb{O}P^{2}$}.

Applying the Gauss equation yields

(16) 3B(X,Y),B(Z,W)=\displaystyle 3\langle B(X,Y),B(Z,W)\rangle= R(X,Z)W,Y+R(X,W)Z,Y+λ2X,YZ,W\displaystyle\langle R(X,Z)W,Y\rangle+\langle R(X,W)Z,Y\rangle+\lambda^{2}\langle X,Y\rangle\langle Z,W\rangle
+λ2X,WY,Z+λ2X,ZW,Y,\displaystyle+\lambda^{2}\langle X,W\rangle\langle Y,Z\rangle+\lambda^{2}\langle X,Z\rangle\langle W,Y\rangle,

where X,Y,Z,WTq𝕆P2X,Y,Z,W\in T_{q}\mathbb{\mathbb{O}}P^{2} (see Proposition 3.6 in [Ohn86]).

3.4.3. Proof of Theorem 1.2

Let Φ=(ϕ1,ϕ2):ΣM¯:=𝕆P2×M\Phi=(\phi_{1},\phi_{2}):\Sigma\rightarrow\bar{M}:=\mathbb{O}P^{2}\times M be a stable compact minimal immersion of codimension dd and dimension nn, where MM is any Riemannian manifold of dimension m2m_{2}.

Let pΣp\in\Sigma, {e1,,en}\{e_{1},\ldots,e_{n}\} be an orthonormal basis of dΦ(TpΣ)d\Phi(T_{p}\Sigma), and {η1,,ηd}\{\eta_{1},\ldots,\eta_{d}\} be an orthonormal basis of NpΣN_{p}\Sigma. Let ej1e_{j}^{1} be the projection of eje_{j} onto Tϕ1(p)𝕆P2T_{\phi_{1}(p)}\mathbb{O}P^{2} and ej2e_{j}^{2} be the projection of eje_{j} onto Tϕ2(p)MT_{\phi_{2}(p)}M. Similarly, let ηj1\eta_{j}^{1} be the projection of ηj\eta_{j} onto Tϕ1(p)𝕆P2T_{\phi_{1}(p)}\mathbb{O}P^{2} and ηj2\eta_{j}^{2} be the projection of ηj\eta_{j} onto Tϕ2(p)MT_{\phi_{2}(p)}M. We identify Tϕ1(p)𝕆P2T_{\phi_{1}(p)}\mathbb{O}P^{2} with 𝕆𝕆\mathbb{O}\oplus\mathbb{O}.

For each vmv\in\mathbb{R}^{m}, we identify vv with (v,0)T(m×M)(v,0)\in T(\mathbb{R}^{m}\times M). Same as in Section 3.1.2, we denote by NvN_{v} the projection of vv onto the normal space NpΣN_{p}\Sigma in TΦ(p)M¯T_{\Phi(p)}\bar{M}. As in the complex and quaternionic cases, we are going to use the normal sections NvN_{v} in the second variation formula, where vv runs over an orthonormal basis of m\mathbb{R}^{m}.

Lemma 3.9.

Same assumptions as above. Let {E1,,Em}\{E_{1},\ldots,E_{m}\} be the usual canonical basis of m\mathbb{R}^{m}. Then,

A=1mNEA,JΣ(NEA)\displaystyle\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle
(17) =\displaystyle= λ2(j=1ni=1n|ej1|2|ProjL(ej1)ei1|28j=1ni=1nej1,ei126j=1nk=1dej1,ηk12)\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}|e_{j}^{1}|^{2}|\operatorname{Proj}_{L(e_{j}^{1})}e_{i}^{1}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)}
(18) =\displaystyle= λ2(k=1dl=1d|ηk1|2|ProjL(ηk1)ηl1|28k=1dl=1dηk1,ηl126j=1nk=1dej1,ηk12),\displaystyle{\lambda^{2}}\bigg{(}\sum_{k=1}^{d}\sum_{l=1}^{d}|\eta_{k}^{1}|^{2}|\operatorname{Proj}_{L(\eta_{k}^{1})}\eta_{l}^{1}|^{2}-8\sum_{k=1}^{d}\sum_{l=1}^{d}\langle\eta^{1}_{k},\eta^{1}_{l}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)},

where L(ej1)L(e_{j}^{1}) is the octonionic line containing ej1e_{j}^{1} as defined above when ej10e_{j}^{1}\neq 0. When ej1=0e_{j}^{1}=0, we just let L(ej1)L(e_{j}^{1}) be any octonionic line. Same for L(ηk1)L(\eta_{k}^{1}).

Proof.

We proceed in the same way as in the proof of Lemma 3.1 in [RL22]. By Lemma 3.2, we have the following

(19) A=1mNEA,JΣ(NEA)=j=1nk=1d2|B(ej1,ηk1)|2B(ηk1,ηk1),B(ej1,ej1).\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle=\sum_{j=1}^{n}\sum_{k=1}^{d}2|B(e_{j}^{1},\eta_{k}^{1})|^{2}-\langle B(\eta_{k}^{1},\eta_{k}^{1}),B(e_{j}^{1},e_{j}^{1})\rangle.

Using Equation (16), we have

3|B(ej1,ηk1)|2=R(ej1,ηk1)ej1,ηk1+2λ2ej1,ηk12+λ2|ej1|2|ηk1|2\displaystyle 3|B(e^{1}_{j},\eta^{1}_{k})|^{2}=\langle R(e^{1}_{j},\eta^{1}_{k})e^{1}_{j},\eta^{1}_{k}\rangle+2\lambda^{2}\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}+\lambda^{2}|e^{1}_{j}|^{2}|\eta^{1}_{k}|^{2}

and

3B(ηk1,ηk1),B(ej1,ej1)=2R(ej1,ηk1)ej1,ηk1+λ2|ej1|2|ηk1|2+2λ2ej1,ηk12\displaystyle 3\langle B(\eta_{k}^{1},\eta_{k}^{1}),B(e_{j}^{1},e_{j}^{1})\rangle=-2\langle R(e^{1}_{j},\eta^{1}_{k})e^{1}_{j},\eta^{1}_{k}\rangle+\lambda^{2}|e^{1}_{j}|^{2}|\eta^{1}_{k}|^{2}+2\lambda^{2}\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}.

Now applying the last two equalities in Equation (19), we obtain

A=1mNEA,JΣ(NEA)\displaystyle\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle =j=1nk=1d43R(ej1,ηk1)ηk1,ej1+2λ23ej1,ηk12+λ23|ej1|2|ηk1|2\displaystyle=\sum_{j=1}^{n}\sum_{k=1}^{d}-\frac{4}{3}\langle R(e^{1}_{j},\eta^{1}_{k})\eta^{1}_{k},e^{1}_{j}\rangle+\frac{2\lambda^{2}}{3}\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}+\frac{\lambda^{2}}{3}|e^{1}_{j}|^{2}|\eta^{1}_{k}|^{2}
(20) =j=1nk=1d43R(ej1,ηk1)ej1,ηk1+2λ23ej1,ηk12+λ23|ej1|2|ηk1|2.\displaystyle=\sum_{j=1}^{n}\sum_{k=1}^{d}\frac{4}{3}\langle R(e^{1}_{j},\eta^{1}_{k})e^{1}_{j},\eta^{1}_{k}\rangle+\frac{2\lambda^{2}}{3}\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}+\frac{\lambda^{2}}{3}|e^{1}_{j}|^{2}|\eta^{1}_{k}|^{2}.

We look at the summand. Notice that each term in the summand is invariant under isometries. Since Spin(9)\operatorname{Spin}(9) is isomorphic to the isotropy group at pp and Spin(9)\operatorname{Spin}(9) acts transitively and effectively on S15S^{15}, there exists some gjSpin(9)g_{j}\in\operatorname{Spin}(9) such that gj(ej1)=(|ej1|,0)𝕆𝕆g_{j}(e_{j}^{1})=(|e_{j}^{1}|,0)\in\mathbb{O}\oplus\mathbb{O}. We then have gj(ej1)=(a,0)g_{j}(e_{j}^{1})=(a,0) for some aa\in\mathbb{R}, a>0a>0 and gj(ηk1)=(c,d)g_{j}(\eta_{k}^{1})=(c,d) for some c,d𝕆c,d\in\mathbb{O}.

Also note that the setwise stabilizer of the 8-dimensional subspace 𝕆0\mathbb{O}\oplus 0 is isomorphic to Spin(8)\operatorname{Spin}(8). For the vectors (Is,0)(I_{s},0), s=0,,7s=0,\dots,7, let hsh_{s} be an element of Spin(8)Spin(9)\operatorname{Spin}(8)\subset\operatorname{Spin}(9) such that hs((I0,0))=(Is,0)h_{s}((I_{0},0))=(I_{s},0) and we take h0=Idh_{0}=\operatorname{Id}.

Now using Equation (15) for the curvature tensor, we have

43R(ej1,ηk1)ej1,ηk1+2λ23ej1,ηk12+λ23|ej1|2|ηk1|2\displaystyle\frac{4}{3}\langle R(e^{1}_{j},\eta^{1}_{k})e^{1}_{j},\eta^{1}_{k}\rangle+\frac{2\lambda^{2}}{3}\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}+\frac{\lambda^{2}}{3}|e^{1}_{j}|^{2}|\eta^{1}_{k}|^{2}
=\displaystyle= 43R(gj(ej1),gj(ηk1))gj(ej1),gj(ηk1)+2λ23gj(ej1),gj(ηk1)2+λ23|gj(ej1)|2|gj(ηk1)|2\displaystyle\frac{4}{3}\langle R\Big{(}g_{j}(e^{1}_{j}),g_{j}(\eta^{1}_{k})\Big{)}g_{j}(e^{1}_{j}),g_{j}(\eta^{1}_{k})\rangle+\frac{2\lambda^{2}}{3}\langle g_{j}(e^{1}_{j}),g_{j}(\eta^{1}_{k})\rangle^{2}+\frac{\lambda^{2}}{3}|g_{j}(e^{1}_{j})|^{2}|g_{j}(\eta^{1}_{k})|^{2}
=\displaystyle= 43R((a,0),(c,d))(a,0),(c,d)+2λ23a,c2+λ23|a|2(|c|2+|d|2)\displaystyle\frac{4}{3}\langle R\Big{(}(a,0),(c,d)\Big{)}(a,0),(c,d)\rangle+\frac{2\lambda^{2}}{3}\langle a,c\rangle^{2}+\frac{\lambda^{2}}{3}|a|^{2}\Big{(}|c|^{2}+|d|^{2}\Big{)}
=\displaystyle= λ23(4|a|2|c|2+4a,c2|ad|2)+2λ23a,c2+λ23(|a|2|c|2+|a|2|d|2)\displaystyle\frac{\lambda^{2}}{3}\bigg{(}-4|a|^{2}|c|^{2}+4\langle a,c\rangle^{2}-|ad|^{2}\bigg{)}+\frac{2\lambda^{2}}{3}\langle a,c\rangle^{2}+\frac{\lambda^{2}}{3}\Big{(}|a|^{2}|c|^{2}+|a|^{2}|d|^{2}\Big{)}
=\displaystyle= λ2(|a|2|c|2+2a,c2)\displaystyle{\lambda^{2}}\bigg{(}-|a|^{2}|c|^{2}+2\langle a,c\rangle^{2}\bigg{)}
=\displaystyle= λ2(|a|2s=07Is,c2+2a,c2)\displaystyle{\lambda^{2}}\bigg{(}-|a|^{2}\sum_{s=0}^{7}\langle I_{s},c\rangle^{2}+2\langle a,c\rangle^{2}\bigg{)}
=\displaystyle= λ2(s=07(aIs,c+0,d)2+2a,c2)\displaystyle{\lambda^{2}}\bigg{(}-\sum_{s=0}^{7}\big{(}\langle aI_{s},c\rangle+\langle 0,d\rangle\big{)}^{2}+2\langle a,c\rangle^{2}\bigg{)}
=\displaystyle= λ2(s=07(hsgj(ej1),gj(ηk1)2+2gj(ej1),gj(ηk1)2)\displaystyle{\lambda^{2}}\bigg{(}-\sum_{s=0}^{7}\big{(}\langle h_{s}g_{j}(e_{j}^{1}),g_{j}(\eta_{k}^{1})\rangle^{2}+2\langle g_{j}(e^{1}_{j}),g_{j}(\eta^{1}_{k})\rangle^{2}\bigg{)}
=\displaystyle= λ2(s=07(gj1hsgj(ej1),ηk12+2ej1,ηk12).\displaystyle{\lambda^{2}}\bigg{(}-\sum_{s=0}^{7}\big{(}\langle g_{j}^{-1}h_{s}g_{j}(e_{j}^{1}),\eta_{k}^{1}\rangle^{2}+2\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}\bigg{)}.

Therefore Equation (20) becomes

(21) A=1mNEA,JΣ(NEA)=λ2(j=1nk=1ds=07(gj1hsgj(ej1),ηk12+2j=1nk=1dej1,ηk12).\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle={\lambda^{2}}\bigg{(}-\sum_{j=1}^{n}\sum_{k=1}^{d}\sum_{s=0}^{7}\big{(}\langle g_{j}^{-1}h_{s}g_{j}(e_{j}^{1}),\eta_{k}^{1}\rangle^{2}+2\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e^{1}_{j},\eta^{1}_{k}\rangle^{2}\bigg{)}.

Since e1,,en,η1,.ηde_{1},\dots,e_{n},\eta_{1},\dots.\eta_{d} form a basis of TΦ(p)(𝕆P2×MT_{\Phi(p)}(\mathbb{O}P^{2}\times M), we have

(ej1,0)=i=1n(ej1,0),eiei+k=1d(ej1,0),ηkηk,(e_{j}^{1},0)=\sum_{i=1}^{n}\langle(e_{j}^{1},0),e_{i}\rangle e_{i}+\sum_{k=1}^{d}\langle(e_{j}^{1},0),\eta_{k}\rangle\eta_{k},

so looking at the projection onto Tϕ1(p)𝕆P2T_{\phi_{1}(p)}\mathbb{O}P^{2} yields

ej1=i=1nej1,ei1ei1+k=1dej1,ηk1ηk1,e_{j}^{1}=\sum_{i=1}^{n}\langle e_{j}^{1},e_{i}^{1}\rangle e_{i}^{1}+\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle\eta_{k}^{1},

and taking inner product of ej1e_{j}^{1} with itself gives

(22) |ej1|2=i=1nej1,ei12+k=1dej1,ηk12.|e^{1}_{j}|^{2}=\sum_{i=1}^{n}\langle e_{j}^{1},e_{i}^{1}\rangle^{2}+\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}.

Similarly, using gj1hsgj(ej1)g_{j}^{-1}h_{s}g_{j}(e_{j}^{1}) instead of ej1e_{j}^{1} gives

(23) |ej1|2=|gj1hsgj(ej1)|2=i=1ngj1hsgj(ej1),ei12+k=1dgj1hsgj(ej1),ηk12.|e^{1}_{j}|^{2}=|g_{j}^{-1}h_{s}g_{j}(e_{j}^{1})|^{2}=\sum_{i=1}^{n}\langle g_{j}^{-1}h_{s}g_{j}(e_{j}^{1}),e_{i}^{1}\rangle^{2}+\sum_{k=1}^{d}\langle g_{j}^{-1}h_{s}g_{j}(e_{j}^{1}),\eta_{k}^{1}\rangle^{2}.

Substituting Equations (22) and (23) back into Equation (21), we have

A=1mNEA,JΣ(NEA)\displaystyle\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle
=\displaystyle= λ2(j=1ns=07(i=1ngj1hsgj(ej1),ei12|ej1|2)+2j=1n(|ej1|2i=1nej1,ei12))\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{s=0}^{7}\Big{(}\sum_{i=1}^{n}\langle g_{j}^{-1}h_{s}g_{j}(e_{j}^{1}),e_{i}^{1}\rangle^{2}-|e_{j}^{1}|^{2}\Big{)}+2\sum_{j=1}^{n}\Big{(}|e_{j}^{1}|^{2}-\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}\Big{)}\bigg{)}
=\displaystyle= λ2(j=1ni=1ns=07hsgj(ej1),gj(ei1)22j=1ni=1nej1,ei126j=1n|ej1|2)\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}\sum_{s=0}^{7}\langle h_{s}g_{j}(e_{j}^{1}),g_{j}(e_{i}^{1})\rangle^{2}-2\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}|e_{j}^{1}|^{2}\bigg{)}
=\displaystyle= λ2(j=1ni=1ns=07hs((|ej1|,0)),gj(ei1)22j=1ni=1nej1,ei126j=1n(i=1nej1,ei12+k=1dej1,ηk12))\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}\sum_{s=0}^{7}\langle h_{s}((|e_{j}^{1}|,0)),g_{j}(e_{i}^{1})\rangle^{2}-2\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\Big{(}\sum_{i=1}^{n}\langle e_{j}^{1},e_{i}^{1}\rangle^{2}+\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\Big{)}\bigg{)}
=\displaystyle= λ2(j=1ni=1ns=07|ej1|2(Is,0),gj(ei1)28j=1ni=1nej1,ei126j=1nk=1dej1,ηk12)\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}\sum_{s=0}^{7}|e_{j}^{1}|^{2}\langle(I_{s},0),g_{j}(e_{i}^{1})\rangle^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)}
=\displaystyle= λ2(j=1ni=1n|ej1|2|Proj𝕆0gj(ei1)|28j=1ni=1nej1,ei126j=1nk=1dej1,ηk12)\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}|e_{j}^{1}|^{2}|\operatorname{Proj}_{\mathbb{O}\oplus 0}g_{j}(e_{i}^{1})|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)}
=\displaystyle= λ2(j=1ni=1n|ej1|2|Projgj1(𝕆0)ei1|28j=1ni=1nej1,ei126j=1nk=1dej1,ηk12)\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}|e_{j}^{1}|^{2}|\operatorname{Proj}_{g_{j}^{-1}(\mathbb{O}\oplus 0)}e_{i}^{1}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)}
=\displaystyle= λ2(j=1ni=1n|ej1|2|ProjL(ej1)ei1|28j=1ni=1nej1,ei126j=1nk=1dej1,ηk12),\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}|e_{j}^{1}|^{2}|\operatorname{Proj}_{L(e_{j}^{1})}e_{i}^{1}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)},

which is Equation (17).

Equation (18) follows from symmetry between eje_{j} and ηk\eta_{k} in Equation (20) and Equation (15) for the curvature tensor.

Proposition 3.10.

Let x1,xnx_{1},\dots x_{n} be nonzero vectors in 𝕆𝕆\mathbb{O}\oplus\mathbb{O}. Then

j=1ni=1n|xj|2|ProjL(xj)xi|28j=1ni=1nxi,xj20.\sum_{j=1}^{n}\sum_{i=1}^{n}|x_{j}|^{2}|\operatorname{Proj}_{L(x_{j})}x_{i}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle x_{i},x_{j}\rangle^{2}\leq 0.
Proof.

Group the xix_{i}’s according to the octonionic lines they belong to. Let m1,,mk\ell_{m_{1}},\dots,\ell_{m_{k}} be the resulting distinct octonionic lines. For each ms\ell_{m_{s}}, let x1s,,xnssx_{1}^{s},\dots,x_{n_{s}}^{s} be the vectors among x1,,xnx_{1},\dots,x_{n} that are on the line ms\ell_{m_{s}}. Let XsX_{s} be the 16×ns16\times n_{s} matrix

Xs=[x1sxnss].X_{s}=[x_{1}^{s}\dots x_{n_{s}}^{s}].

Since XsTXs=[xis,xjs]i,j=1nsX_{s}^{T}X_{s}=[\langle x_{i}^{s},x_{j}^{s}\rangle]_{i,j=1}^{n_{s}} is a Gram matrix, it is symmetric positive semi-definite. There thus exists some orthogonal matrix QsQ_{s} of size nsn_{s} such that QsTXsTXsQsQ_{s}^{T}X_{s}^{T}X_{s}Q_{s} is diagonal. Further, since the columns of XsX_{s} are from the same octonionic line, XsX_{s} has at most rank 88. Thus we can assume that QsTXsTXsQs=diag(λs,1,λs,8,0,0)Q_{s}^{T}X_{s}^{T}X_{s}Q_{s}=\operatorname{diag}(\lambda_{s,1}\dots,\lambda_{s,8},0\dots,0), where λs,1,λs,80\lambda_{s,1}\dots,\lambda_{s,8}\geq 0 are the eigenvalues of XsTXsX_{s}^{T}X_{s}. The matrix identity

[(XsQs)i,(XsQs)j]i,j=1ns=QsTXsTXsQs=diag(λs,1,λs,8,0,0)[\langle(X_{s}Q_{s})_{i},(X_{s}Q_{s})_{j}\rangle]_{i,j=1}^{n_{s}}=Q_{s}^{T}X_{s}^{T}X_{s}Q_{s}=\operatorname{diag}(\lambda_{s,1}\dots,\lambda_{s,8},0\dots,0)

tells us that the 16×ns16\times n_{s} matrix XsQsX_{s}Q_{s} has orthogonal columns, and all but the first eight columns are zero vectors. Since XsQsX_{s}Q_{s} and XsX_{s} have the same column span, there exist orthonormal vectors y1s,,y8smsy_{1}^{s},\dots,y_{8}^{s}\in\ell_{m_{s}} such that if we define an 16×816\times 8 matrix YsY_{s} and an 8×ns8\times n_{s} matrix DsD_{s} as

Ys=[y1sy8s]andDs=[diag(λs,1,λs,8) 00],Y_{s}=[y_{1}^{s}\dots y_{8}^{s}]\quad\text{and}\quad D_{s}=[\operatorname{diag}(\sqrt{\lambda_{s,1}}\dots,\sqrt{\lambda_{s,8}})\ 0\dots 0],

then XsQs=YsDsX_{s}Q_{s}=Y_{s}D_{s}. This gives Xs=YsDsQsTX_{s}=Y_{s}D_{s}Q_{s}^{T}.

The strategy for the proof is to rewrite

j=1ni=1n|xj|2|ProjL(xj)xi|28j=1ni=1nxi,xj2\displaystyle\sum_{j=1}^{n}\sum_{i=1}^{n}|x_{j}|^{2}|\operatorname{Proj}_{L(x_{j})}x_{i}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle x_{i},x_{j}\rangle^{2}

in terms of the eigenvalues λs,j\lambda_{s,j} so that we get a nicer expression, for which we can do maximization. In doing so, we need the ms\ell_{m_{s}}’s to be octonionic lines instead of aribitrary 8-dimensional spaces, and Lemma 3.8 plays a crucial role.

We first compute

j=1ni=1nxi,xj2\displaystyle\sum_{j=1}^{n}\sum_{i=1}^{n}\langle x_{i},x_{j}\rangle^{2} =r=1ks=1ki=1nrj=1nsxir,xjs2\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{n_{r}}\sum_{j=1}^{n_{s}}\langle x^{r}_{i},x^{s}_{j}\rangle^{2}
=r=1ks=1kTr((XrTXs)TXrTXs)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\operatorname{Tr}((X_{r}^{T}X_{s})^{T}X_{r}^{T}X_{s})
=r=1ks=1kTr(XsTXrXrTXs)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\operatorname{Tr}(X_{s}^{T}X_{r}X_{r}^{T}X_{s})
=r=1ks=1kTr(XrXrTXsXsT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\operatorname{Tr}(X_{r}X_{r}^{T}X_{s}X_{s}^{T})
=r=1ks=1kTr(YrDrQrT(YrDrQrT)TYsDsQsT(YsDsQsT)T)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\operatorname{Tr}(Y_{r}D_{r}Q_{r}^{T}(Y_{r}D_{r}Q_{r}^{T})^{T}Y_{s}D_{s}Q_{s}^{T}(Y_{s}D_{s}Q_{s}^{T})^{T})
=r=1ks=1kTr(YsTYrDrDrTYrTYsDsDsT).\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\operatorname{Tr}\Big{(}Y_{s}^{T}Y_{r}D_{r}D_{r}^{T}Y_{r}^{T}Y_{s}D_{s}D_{s}^{T}).

Lemma 3.8 shows that YrTYs=cr,sAr,sY_{r}^{T}Y_{s}=c_{r,s}A_{r,s} for some orthogonal matrix Ar,sA_{r,s} of size 88, and some constant 0cr,s10\leq c_{r,s}\leq 1. Also, cr,s=cs,rc_{r,s}=c_{s,r} and Ar,s=As,rTA_{r,s}=A_{s,r}^{T}. In particular, when r=sr=s, we have cr,r=1c_{r,r}=1 and Ar,r=Id8A_{r,r}=\operatorname{Id}_{8}. Thus

j=1ni=1nxi,xj2\displaystyle\sum_{j=1}^{n}\sum_{i=1}^{n}\langle x_{i},x_{j}\rangle^{2} =r=1ks=1kTr(YsTYrDrDrTYrTYsDsDsT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\operatorname{Tr}(Y_{s}^{T}Y_{r}D_{r}D_{r}^{T}Y_{r}^{T}Y_{s}D_{s}D_{s}^{T})
=r=1ks=1kcr,s2Tr(Ar,sTDrDrTAr,sDsDsT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}c_{r,s}^{2}\operatorname{Tr}(A_{r,s}^{T}D_{r}D_{r}^{T}A_{r,s}D_{s}D_{s}^{T})
(24) =r=1ks=1ki=18j=18cr,s2λr,iλs,j(Ar,s)ij2.\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}(A_{r,s})_{ij}^{2}.

We then look at the quantity

j=1ni=1n|xj|2|ProjL(xj)xi|2\displaystyle\sum_{j=1}^{n}\sum_{i=1}^{n}|x_{j}|^{2}|\operatorname{Proj}_{L(x_{j})}x_{i}|^{2}.

Notice that since XsTXs=[xis,xjs]i,j=1nsX_{s}^{T}X_{s}=[\langle x_{i}^{s},x_{j}^{s}\rangle]_{i,j=1}^{n_{s}}, we have

Tr(XsTXs)=j=1ns|xjs|2\displaystyle\operatorname{Tr}(X_{s}^{T}X_{s})=\sum_{j=1}^{n_{s}}|x_{j}^{s}|^{2}.

On the other hand, since XsTXsX_{s}^{T}X_{s} is a symmetric matrix, its trace equals to the sum of its eigenvalues. Thus j=1ns|xjs|2=Tr(XsTXs)=j=18λs,j\sum_{j=1}^{n_{s}}|x_{j}^{s}|^{2}=\operatorname{Tr}(X_{s}^{T}X_{s})=\sum_{j=1}^{8}\lambda_{s,j}. Using this and the identities Xr=YrDrQrTX_{r}=Y_{r}D_{r}Q_{r}^{T} and YrTYs=cr,sAr,sY_{r}^{T}Y_{s}=c_{r,s}A_{r,s}, we get

j=1ni=1n|xj|2|ProjL(xj)xi|2\displaystyle\sum_{j=1}^{n}\sum_{i=1}^{n}|x_{j}|^{2}|\operatorname{Proj}_{L(x_{j})}x_{i}|^{2} =r=1ks=1ki=1nrj=18λs,j|Projmsxir|2\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{n_{r}}\sum_{j=1}^{8}\lambda_{s,j}|\operatorname{Proj}_{\ell_{m_{s}}}x_{i}^{r}|^{2}
=r=1ks=1kj=18λs,ji=1nrp=18xir,yps2\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\sum_{i=1}^{n_{r}}\sum_{p=1}^{8}\langle x_{i}^{r},y_{p}^{s}\rangle^{2}
=r=1ks=1kj=18λs,jTr((XrTYs)TXrTYs)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}((X_{r}^{T}Y_{s})^{T}X_{r}^{T}Y_{s})
=r=1ks=1kj=18λs,jTr(YsTXrXrTYs)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}(Y_{s}^{T}X_{r}X_{r}^{T}Y_{s})
=r=1ks=1kj=18λs,jTr(YsYsTXrXrT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}(Y_{s}Y_{s}^{T}X_{r}X_{r}^{T})
=r=1ks=1kj=18λs,jTr(YsYsT(YrDrQrT)(YrDrQrT)T)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}(Y_{s}Y_{s}^{T}(Y_{r}D_{r}Q_{r}^{T})(Y_{r}D_{r}Q_{r}^{T})^{T})
=r=1ks=1kj=18λs,jTr(YsYsTYrDrQrTQrDrTYrT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}(Y_{s}Y_{s}^{T}Y_{r}D_{r}Q_{r}^{T}Q_{r}D_{r}^{T}Y_{r}^{T})
=r=1ks=1kj=18λs,jTr(YsYsTYrDrDrTYrT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}(Y_{s}Y_{s}^{T}Y_{r}D_{r}D_{r}^{T}Y_{r}^{T})
=r=1ks=1kj=18λs,jTr(YsTYrDrDrTYrTYs)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}\operatorname{Tr}(Y_{s}^{T}Y_{r}D_{r}D_{r}^{T}Y_{r}^{T}Y_{s})
=r=1ks=1kj=18λs,jcr,s2Tr(Ar,sTDrDrTAr,s)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}c_{r,s}^{2}\operatorname{Tr}(A_{r,s}^{T}D_{r}D_{r}^{T}A_{r,s})
=r=1ks=1kj=18λs,jcr,s2Tr(Ar,sAr,sTDrDrT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}c_{r,s}^{2}\operatorname{Tr}(A_{r,s}A_{r,s}^{T}D_{r}D_{r}^{T})
=r=1ks=1kj=18λs,jcr,s2Tr(DrDrT)\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}c_{r,s}^{2}\operatorname{Tr}(D_{r}D_{r}^{T})
=r=1ks=1kj=18λs,jcr,s2i=18λr,i\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}\lambda_{s,j}c_{r,s}^{2}\sum_{i=1}^{8}\lambda_{r,i}
(25) =r=1ks=1ki=18j=18cr,s2λr,iλs,j.\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}.

Using (24) and (25), we have that

j=1ni=1n|xj|2|ProjL(xj)xi|28j=1ni=1nxi,xj2\displaystyle\sum_{j=1}^{n}\sum_{i=1}^{n}|x_{j}|^{2}|\operatorname{Proj}_{L(x_{j})}x_{i}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle x_{i},x_{j}\rangle^{2}
=\displaystyle= r=1ks=1ki=18j=18cr,s2λr,iλs,j8r=1ks=1ki=18j=18cr,s2λr,iλs,j(Ar,s)ij2\displaystyle\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}-8\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}(A_{r,s})_{ij}^{2}
=:\displaystyle=: f(,λr,i,),\displaystyle f(\dots,\lambda_{r,i},\dots),

where ff is a function of each λr,i\lambda_{r,i}.

Since

fλr,i=2s=1kj=18cr,s2λs,j16s=1kj=18cr,s2λs,j(Ar,s)ij2,\frac{\partial f}{\partial\lambda_{r,i}}=2\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{s,j}-16\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{s,j}(A_{r,s})_{ij}^{2},

summing over ii yields

(26) i=18fλr,i\displaystyle\sum_{i=1}^{8}\frac{\partial f}{\partial\lambda_{r,i}} =16s=1kj=18cr,s2λs,j16s=1kj=18cr,s2λs,ji=18(Ar,s)ij2\displaystyle=16\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{s,j}-16\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{s,j}\sum_{i=1}^{8}(A_{r,s})_{ij}^{2}
=16s=1kj=18cr,s2λs,j16s=1kj=18cr,s2λs,j\displaystyle=16\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{s,j}-16\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{s,j}
=0,\displaystyle=0,

where in the second to last step we used the fact that Ar,sA_{r,s} is an orthogonal matrix.

We maximize ff subject to the constraint

r=1ki=18λr,i2=C\displaystyle\sum_{r=1}^{k}\sum_{i=1}^{8}\lambda_{r,i}^{2}=C

for some constant C0C\geq 0. Using Lagrange multipliers, at the maximum of ff subject to the constraint, there exists some α\alpha\in\mathbb{R} such that for all r,ir,i,

(27) fλr,i=αλr,i.\displaystyle\frac{\partial f}{\partial\lambda_{r,i}}=\alpha\lambda_{r,i}.

Using (26) and summing over ii yields

0=i=18fλr,i\displaystyle 0=\sum_{i=1}^{8}\frac{\partial f}{\partial\lambda_{r,i}} =αi=18λr,i.\displaystyle=\alpha\sum_{i=1}^{8}\lambda_{r,i}.

Thus either α=0\alpha=0, or i=18λr,i=0\sum_{i=1}^{8}\lambda_{r,i}=0 for all rr.

If α=0\alpha=0, then by Equation (27), we have fλr,i=0\frac{\partial f}{\partial\lambda_{r,i}}=0 for all r,ir,i. Thus

0\displaystyle 0 =r=1ki=18λr,ifλr,i\displaystyle=\sum_{r=1}^{k}\sum_{i=1}^{8}\lambda_{r,i}\frac{\partial f}{\partial\lambda_{r,i}}
=2r=1ki=18s=1kj=18cr,s2λr,iλs,j16r=1ki=18s=1kj=18cr,s2λr,iλs,j(Ar,s)ij2\displaystyle=2\sum_{r=1}^{k}\sum_{i=1}^{8}\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}-16\sum_{r=1}^{k}\sum_{i=1}^{8}\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}(A_{r,s})_{ij}^{2}
=2f.\displaystyle=2f.

If instead i=18λr,i=0\sum_{i=1}^{8}\lambda_{r,i}=0 for all rr, then

f\displaystyle f =r=1ks=1ki=18j=18cr,s2λr,iλs,j8r=1ks=1ki=18j=18cr,s2λr,iλs,j(Ar,s)ij2\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}-8\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}(A_{r,s})_{ij}^{2}
=r=1ks=1kj=18cr,s2(i=18λr,i)(j=18λs,j)8r=1ks=1ki=18j=18cr,s2λr,iλs,j(Ar,s)ij2\displaystyle=\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{j=1}^{8}c_{r,s}^{2}\Big{(}\sum_{i=1}^{8}\lambda_{r,i}\Big{)}\Big{(}\sum_{j=1}^{8}\lambda_{s,j}\Big{)}-8\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}(A_{r,s})_{ij}^{2}
=8r=1ks=1ki=18j=18cr,s2λr,iλs,j(Ar,s)ij2\displaystyle=-8\sum_{r=1}^{k}\sum_{s=1}^{k}\sum_{i=1}^{8}\sum_{j=1}^{8}c_{r,s}^{2}\lambda_{r,i}\lambda_{s,j}(A_{r,s})_{ij}^{2}
=8j=1ni=1nxi,xj2\displaystyle=-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle x_{i},x_{j}\rangle^{2}
0.\displaystyle\leq 0.

In either case, at the maximum of ff subject to r=1ki=18λr,i2=C\sum_{r=1}^{k}\sum_{i=1}^{8}\lambda_{r,i}^{2}=C, we have f0f\leq 0.

Since C0C\geq 0 is arbitrary, we have f0f\leq 0 always, which gives the desired inequality. ∎

Remark 3.11.

Notice that in the above proposition, if n8n\leq 8, then the proof is significantly simpler: just use the trivial estimate |ProjL(xj)xi||xi||\operatorname{Proj}_{L(x_{j})}x_{i}|\leq|x_{i}| and apply Proposition 2.3.

By Lemma 3.9 and Proposition 3.10, at each point pΣp\in\Sigma, we have

A=1mNEA,JΣ(NEA)\displaystyle\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle
=\displaystyle= λ2(j=1ni=1n|ej1|2|ProjL(ej1)ei1|28j=1ni=1nej1,ei126j=1nk=1dej1,ηk12)\displaystyle{\lambda^{2}}\bigg{(}\sum_{j=1}^{n}\sum_{i=1}^{n}|e_{j}^{1}|^{2}|\operatorname{Proj}_{L(e_{j}^{1})}e_{i}^{1}|^{2}-8\sum_{j=1}^{n}\sum_{i=1}^{n}\langle e^{1}_{j},e_{i}^{1}\rangle^{2}-6\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}\bigg{)}
\displaystyle\leq 6λ2j=1nk=1dej1,ηk12\displaystyle-6{\lambda^{2}}\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}
\displaystyle\leq 0.\displaystyle 0.

Now recall Φ=(ϕ1,ϕ2):Σ𝕆P2×M\Phi=(\phi_{1},\phi_{2}):\Sigma\rightarrow\mathbb{O}P^{2}\times M is a stable compact minimal immersion. The stability of Φ\Phi requires

ΣA=1mNEA,JΣ(NEA)0.\int_{\Sigma}\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle\geq 0.

Combining the above two inequalites, we must have

0=A=1mNEA,JΣ(NEA)=6λ2j=1nk=1dej1,ηk12.0=\sum_{A=1}^{m}-\langle N_{E_{A}},J_{\Sigma}(N_{E_{A}})\rangle=-6{\lambda^{2}}\sum_{j=1}^{n}\sum_{k=1}^{d}\langle e_{j}^{1},\eta_{k}^{1}\rangle^{2}.

Thus for all j,kj,k,

ej1,ηk1=0.\langle e_{j}^{1},\eta_{k}^{1}\rangle=0.

But since we have 0=ej,ηk=ej1,ηk1+ej2,ηk20=\langle e_{j},\eta_{k}\rangle=\langle e_{j}^{1},\eta_{k}^{1}\rangle+\langle e_{j}^{2},\eta_{k}^{2}\rangle, this shows that we necessarily have

ej2,ηk2=0\langle e_{j}^{2},\eta_{k}^{2}\rangle=0

as well.

For the product manifold 𝕆P2×M\mathbb{O}P^{2}\times M, let

P:T(𝕆P2×M)T𝕆P20T(𝕆P2×M),P:T(\mathbb{O}P^{2}\times M)\to T\mathbb{O}P^{2}\oplus 0\subseteq T(\mathbb{O}P^{2}\times M),
Q:T(𝕆P2×M)0TMT(𝕆P2×M)Q:T(\mathbb{O}P^{2}\times M)\to 0\oplus TM\subseteq T(\mathbb{O}P^{2}\times M)

denote the projection maps. Let F=PQF=P-Q, then F2=IF^{2}=I. For pΣp\in\Sigma, we claim that F(dΦ(TpΣ))dΦ(TpΣ)F(d\Phi(T_{p}{\Sigma}))\subseteq d\Phi(T_{p}{\Sigma}). In fact, since F(ej)=F(ej1,ej2)=(ej1,ej2)F(e_{j})=F(e_{j}^{1},e_{j}^{2})=(e_{j}^{1},-e_{j}^{2}), for each ηk\eta_{k}, we have

F(ej),ηk=ej1,ηk1ej2,ηk2=0.\langle F(e_{j}),\eta_{k}\rangle=\langle e_{j}^{1},\eta_{k}^{1}\rangle-\langle e_{j}^{2},\eta_{k}^{2}\rangle=0.

Therefore, F(ej)NpΣF(e_{j})\perp N_{p}\Sigma. This shows F(ej)dΦ(TpΣ)F(e_{j})\subset d\Phi(T_{p}\Sigma), so F(dΦ(TpΣ))dΦ(TpΣ)F(d\Phi(T_{p}{\Sigma}))\subseteq d\Phi(T_{p}{\Sigma}). Thus Σ\Sigma is an invariant submanifold of 𝕆P2×M\mathbb{O}P^{2}\times M.

Analogously, as in the quaternionic case, by Theorem 1 in [XN00], Σ\Sigma is isometric to a product manifold Σ1×Σ2\Sigma_{1}\times\Sigma_{2}, where Σ1\Sigma_{1} is an immersion into 𝕆P2\mathbb{O}P^{2} and Σ2\Sigma_{2} is an immersion into MM. Since Φ:Σ𝕆P2×M\Phi:\Sigma\to\mathbb{O}P^{2}\times M is a stable compact minimal immersion, we further have that Σ1\Sigma_{1} and Σ2\Sigma_{2} are stable compact minimal immersions into 𝕆P2\mathbb{O}P^{2} and MM, respectively. However, Ohnita [Ohn86, Theorem E] showed that the only nontrivial stable minimal immersed submanifolds of 𝕆P2\mathbb{O}P^{2} are precisely the octonion protective lines 𝕆P1S8\mathbb{O}P^{1}\cong S^{8}. Using this result, we have thus established Theorem 1.2.

References

  • [BG72] Robert Brown and Alfred Gray. Riemannian manifolds with holonomy group Spin(9). Differential Geometry (in honor of Kentaro Yano), Kinokunia, Tokyo, pages 41–59, 1972.
  • [CL21] Otis Chodosh and Chao Li. Stable minimal hypersurfaces in 𝐑4\mathbf{R}^{4}. https://arxiv.org/abs/2108.11462, 2021.
  • [CW13] Hang Chen and Xianfeng Wang. On stable compact minimal submanifolds of Riemannian product manifolds. Journal of Mathematical Analysis and Applications, 2013.
  • [dCP79] Manfredo do Carmo and Chia-Kuei Peng. Stable complete minimal surfaces in 3\mathbb{R}^{3} are planes. Bulletin (New Series) of the American Mathematical Society, 1(6):903–906, 1979.
  • [FCS80] Doris Fischer-Colbrie and Richard Schoen. The structure of complete stable minimal surfaces in 3-manifolds of non-negative scalar curvature. Communications on Pure and Applied Mathematics, 33(2):199–211, 1980.
  • [HSV09] Rowena Held, Iva Stavrov, and Brian VanKoten. (Semi-)Riemannian geometry of (para-)octonionic projective planes. Differential Geometry and its Applications, 27(4):464–481, 2009.
  • [Kot20] Jan Kotrbatỳ. Integral geometry on the octonionic plane. PhD thesis, Friedrich-Schiller-Universität Jena, 2020.
  • [LS73] Blaine Lawson and James Simons. On stable currents and their application to global problems in real and complex geometry. Annals of Mathematics, 1973.
  • [Ohn86] Yoshihiro Ohnita. Stable minimal submanifolds in compact rank one symmetric spaces. Tohoku Mathematical Journal, Second Series, 1986.
  • [OPPV13] Liviu Ornea, Maurizio Parton, Paolo Piccinni, and Victor Vuletescu. Spin(9) geometry of the octonionic Hopf fibration. Transformation Groups, 18(3):845–864, 2013.
  • [Pog81] Aleksei V. Pogorelov. On the stability of minimal surfaces. Doklady Akademii Nauk, 260(2):293–295, 1981.
  • [RL22] Alejandra Ramirez-Luna. Stable submanifolds in the product of projective spaces. https://arxiv.org/abs/2202.00083, 2022.
  • [Sak77] Kunio Sakamoto. Planar geodesic immersions. Tohoku Mathematical Journal, Second Series, 29(1):25–56, 1977.
  • [Sim68] James Simons. Minimal varieties in Riemannian manifolds. Annals of Mathematics, 1968.
  • [TU14] Francisco Torralbo and Francisco Urbano. On stable compact minimal submanifolds. Proceedings of the American Mathematical Society, 2014.
  • [XN00] Senlin Xu and Yilong Ni. Submanifolds of product Riemannian manifold. Acta Mathematica Scientia, 20(2):213–218, 2000.