Stable submanifolds in the product of projective spaces II
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 -dimensional Euclidean space, and recently Chodosh and Li proved that a complete, two-sided, stable minimal hypersurface in the -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.
Theorem 1.2.
[RL22] Let be a compact stable minimal immersion of codimension and dimension , where is any Riemannian manifold of dimension . Then,
-
•
If , , where is a compact stable minimal immersion of codimension , and therefore . In particular, for , , is a constant function, and , for .
-
•
If , , where is a compact stable minimal immersion of codimension , and therefore . In particular, for , there are no compact stable minimal immersions of codimension in . And for , , is a constant function, and , for .
-
•
If , is a stable geodesic, is a constant function, and therefore with a point of .
-
•
If , is a stable minimal immersion of dimension , is a constant function, and therefore with a point of .
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:
thmmainone
There do not exist odd-dimensional stable compact minimal immersions in .
For even-dimensional stable compact minimal immersions in , 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 .
Theorem 1.2 shows that the stable minimal submanifolds of dimension or codimension or in are precisely the product of stable minimal submanifolds in and . In this paper, we generalize this result to arbitrary dimension:
thmmaintwo Let be a stable compact minimal immersion, where is a Riemannian manifold of dimension . Then , , where , are stable compact minimal immersions.
In this paper, we further consider the stable compact minimal immersions of arbitrary dimension in . 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 be a stable compact minimal immersion, where is a Riemannian manifold of dimension . Then , , where is either a point, or , and , 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 be an matrix. Then,
-
•
the row spaces of and are the same.
-
•
the symmetric matrices and have the same nonzero eigenvalues with the same multiplicity.
Proof.
-
•
Let us denote by the row space of a matrix . It is clear that . For the reverse direction, it is enough to show . Let . Then we have
so . Thus
showing , so . Thus as desired.
-
•
Let be a nonzero eigenvalue of with multiplicity . Then there exist orthonormal eigenvectors with eigenvalue . Then, , so are eigenvectors of with eigenvalue . Moreover, we claim they are linearly independent. Let be such that . Then . Since ’s are linearly independent, we must have for each . Thus ’s are linearly independent.
Reverting the roles of and shows the -eigenspaces of and have the same dimension. Thus and have the same nonzero eigenvalues with multiplicity as desired.
∎
For any two matrices , one can define the Frobenius or trace inner product as
which induces the norm
This is the usual Eulidean inner product by identifying with .
We thus have the familiar Cauchy-Schwarz inequality:
and we obtain equality if and only if and are nonnegative multiples of each other.
The particular nice thing about the trace operator is that we have
for any .
Proposition 2.2.
Let be an orthogonal matrix. Let be vectors in . Let be the associated matrix
Then
and we have equality if and only if commutes with . In particular, when we have equality, is -invariant.
Proof.
Notice that , and , so
, and .
Moreover, we also have that
.
We calculate
Thus
as desired. We have equality if and only if we obtain equality in the Cauchy-Schwarz inequality, which happens if and only if and are nonnegative multiples of each other. Since , we have equality if and only if , that is, when .
Lastly, suppose we have . Multiplying on the left to this equation yields . This shows that the row space of is contained in the the row space of . By Proposition 2.1, this means the row space of is contained in the row space of , so . Since they have the same dimension, we have , so is -invariant. Since is orthogonal, , -invariant spaces and -invariant spaces coincide. ∎
We also have the following nice result.
Proposition 2.3.
Let be vectors such that . Then
and we have equality if and only if all the ’s have the same length and are mutually orthogonal.
Proof.
Without loss of generality we can assume . Let be the associated matrix
Then , and .
Using Cauchy-Schwarz inequality, we have
To have equality, we need to be a scalar multiple of the identity. By Proposition 2.1, 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 ’s have the same length and are mutually orthogonal. ∎
3. Main results
3.1. Preliminaries
3.1.1. Second variation formula
Let be a minimal isometric immersion of an -dimensional compact Riemannian manifold . Let denote the space of all -vector fields along . For any in , let be a -one-parameter family of immersions of into such that and for each . Then we have the classical second variational formula
Theorem 3.1 (Second Variation formula).
Under the above assumptions,
,
where the normal
component of and is the elliptic Jacobi operator on the normal bundle defined by
,
and the normal Laplacian on the normal bundle is given by
.
Here, is an orthonormal basis of , is the connection of , is the normal connection of in , is the second fundamental form of in , and is the curvature tensor of .
We say a minimal immersion is stable if for all ,
3.1.2. Immersion into a Riemannian product
The general setting of our problem is as follows. Let be an isometric immersion, where is an -dimensional compact Riemannian manifold, is an -dimensional compact Riemannian submanifold, and is an -dimensional Riemannian manifold. Let denote the codimension of in . We further let be an isometric embedding of in an Euclidean space, and let be defined by . Let denote the second fundamental form of the immersion . We can then consider the chain of immersions .
Let be a point. We choose a local orthonormal frame near such that is an orthonormal frame in and is an orthonormal frame in . For any tangent vector at , we decompose as , where is tangent to at . We will thus write and .
For a fixed vector , by identifying with , we can decompose as
where , and . For , by identifying it with its preimage in and considering the immersion , we can further decompose it as
where and .
The case that will be interesting to us is when is a stable minimal submanifold of . To study this case, we need to use some appropriate normal sections in the second variation formula. To this end, let be an orthonormal basis of . Then , will be the normal sections we use. Computation shows that
Lemma 3.2.
[CW13, Equation (2.8)]
In the next three subsections, we will let be , , and respectively, and study the stable compact minimal immersions . For these projective spaces, there exist isotropic embeddings into Euclidean space so that is constant for any unit vector . In all these cases, we will show , and therefore obtain geometric information about .
3.2. Stable minimal submanifolds in .
We equip each , with the Fubini-Study metric of constant holomorphic sectional curvature .
Let be a stable compact minimal immersion of codimension and dimension .
Let 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 denote its second fundamental form, then at each , we have
(see Proposition 3.6 of [Ohn86]).
For each , we identify with . Same as in Section 3.1.2, we denote by the projection of onto the normal space in . We are now going to use the normal sections in the second variation formula, where runs over an orthonormal basis of the Euclidean space .
Let . Let be an orthonormal basis of . Let be the projection of onto and be the projection of onto . Let be the complex structure of .
Lemma 3.3.
[RL22, Lemma 3.1] Same assumptions as above. Let be the usual canonical basis of . Then
(1) |
The stability of requires
so
On the other hand, combining Lemma 3.3 and Proposition 2.2, we have
The above two equations together show that the integrand must be exactly zero, namely,
(2) |
Now we specialize to the case where , and consider the stable compact minimal immersion of dimension . For , let be the complex structure of . Then letting play the role of in the above arguments, we have
(3) |
as well.
For , choose a local orthonormal basis for and write under this basis as . Let denote the matrix . Then and , and since forms an orthonormal basis of we have
Notice that and are positive semi-definite since they are Gram matrices. Moreover, if is an eigenvalue of with eigenvector , then we have , so , showing is an eigenvalue of with eigenvector .
Thus if are eigenvalues of in non-increasing order with corresponding orthonormal eigenbasis , then are eigenvalues of in non-decreasing order with corresponding orthonormal eigenbasis . In particular, the multiplicity of as an eigenvalue of equals the multiplicity of as an eigenvalue of .
By Proposition 2.2, Equation (2) implies that commutes with , and Equation (3) implies that commutes with . Thus the eigenspaces of are -invariant, therefore even dimensional. Thus all the eigenvalues of have even multiplicity. By Proposition 2.1, this shows all the nonzero eigenvalues of have even multiplicity. Similarly, all the nonzero eigenvalues of have even multiplicity.
In particular, as an eigenvalue of has even multiplicity. Using the fact that the multiplicity of as an eigenvalue of equals the multiplicity of as an eigenvalue of , we get that the multiplicity of as an eigenvalue of is also even. Thus all the eigenvalues of have even multiplicity. However, sum of the multiplicities of the eigenvalues of equals its dimension, which is . This shows must be an even number.
We have thus proved Theorem 1.2.
3.3. Stable minimal submanifolds in .
Equip with its standard metric as a Riemannian symmetric space, with the maximum of the sectional curvatures given by . Let be a stable compact minimal immersion of codimension and dimension , where is any Riemannian manifold of dimension .
Let 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 denote its the second fundamental form, then at each , we have
(see Proposition 3.6 of [Ohn86]).
For each , we identify with . Same as in Section 3.1.2, we denote by the projection of onto the normal space in . As in the complex case, we are going to use the normal sections in the second variation formula, where runs over an orthonormal basis of .
Let . Let be an orthonormal basis of , and let be an orthonormal basis of . Let be the projection of onto and be the projection of onto . Similarly, let be the projection of onto and be the projection of onto . Let be the quaternionic structure on .
Lemma 3.5.
[RL22, Lemma 4.1] Same assumptions as above. Let be the usual canonical basis of . Then, for
(4) |
The stability of requires
so
(5) |
On the other hand, combining Lemma 3.5 and Proposition 2.2, we have for
(6) | ||||
Equations (5) and (6) together show that both inequalities in (6) are equalities. Hence for ,
(7) |
and consequently
(8) |
From Equation (8) we have that for , , and
Thus for fixed, . By Proposition 2.2, Equation (7) also implies that is -invariant, so for each . Thus
(9) |
But since we have , Equation (9) shows that we necessarily have
(10) |
as well.
For the product manifold , let
denote the projection maps. Let , then . For , we claim that . In fact, since , for each , we have
where we have used Equations (9) and (10) in the second equality. Therefore, . This shows , so . Thus is an invariant submanifold of .
By Theorem 1 in [XN00], is isometric to a product manifold , where is an immersion into and is an immersion into . Since is a stable compact minimal immersion, we further have that and are stable compact minimal immersions into and , 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 is not stable.
3.4. Stable minimal submanifolds in .
Let denote the octonionic (Cayley) projective plane, endowed with the standard metric as a Riemannian symmetric space. Classically, can be seen as a 16-dimensional quotient manifold , where the isometry group of is isomophic to the exceptional Lie group and the isotropy group at any point is isomorphic to .
Alternatively, we can also define in terms of equivalence classes (cf. [HSV09]). Consider the relation on , where if and only if there exists such that . 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 :
and their union . The relation on is then an equivalence relation, and we can define the octonionic projective plane as the set of equivalence classes of by . That is,
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 denote the usual octonion basis. For , , let
denote its conjugate. Then the inner product on is given by
Moreover, for , we have the identities
(11) | ||||
(12) | ||||
(13) |
The 16-dimensional real vector space decomposes into octonionic lines of the form
that intersect each other only at the origin . Each line is an 8-dimensional real vector space. Here parametrizes the set of octonionic lines. Be cautious here that the octonionic line through and is not the set due to the non-associativity of octonion multiplication; the correct line is given by .
Proposition 3.7 (see e.g.,[Kot20, Section 2.2]).
Here we collect the properties of that we will use later.
-
•
The group acts transitively and effectively on the unit sphere , and we can view as a matrix group acting on .
-
•
maps octonionic lines to octonionic lines.
-
•
The setwise stabilizer of an octonionic line is isomorphic to , and the stabilizer of a point is isomorphic to .
Define the map by mapping to the unique octonionic line that contains . The restriction of the map to defines the octonionic Hopf fibration
with as the fiber. is the symmetry group of this fibration. (cf. [OPPV13])
Lemma 3.8.
Let , be two octonionic lines defined as above. Let be an orthonormal basis of and be an orthonormal basis of . Let , be the associated matrices
Then , where is an orthogonal matrix, and is some constant. We have if and only if , and if and only if .
Proof.
By Proposition 3.7, by an action of we can assume that and consequently . For , since are orthogonal complements, we have , for and . We can thus assume .
Let , , where are the usual octonion basis. Then form an orthogonal basis for the 8-dimensional vector space . Let
Therefore, there exists an orthogonal matrix such that .
Let , . Then form an orthogonal basis for the 8-dimensional vector space . Let
Therefore, there exists an orthogonal matrix such that .
It is clear to see that
Thus
where is an orthogonal matrix, and . We have precisely when and consequently .
∎
3.4.2. Geometry of
Here we review the basic Riemannian geometry of . We refer the reader to [BG72], [Ohn86], and [HSV09] for details. In each of the chart defined at the beginning of this subsection, let with denote the coordinate functions on the chart. Then they give rise to octonion valued 1-forms and . The standard metric of is then given by (see Equation (3.1) in [HSV09])
where is a scaling factor that denotes the maximum of the sectional curvatures of . Notice that this expression resembles the expression of the Fubini–Study metric of a complex projective plane.
Let denote the Riemannian curvature tensor of . Let be a point in . We can identify with in a natural manner. Using the structure of the octonionic algebra, the curvature tensor is given by
(14) | ||||
where . 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 from the one in [Ohn86] ( versus ) and there is a typo in Equation (1.3) of [Ohn86], where at the end of the second line should be . Here we use this identification so as to be consistent with out definition of octonionic lines earlier.
In particular, we have
(15) | ||||
3.4.3. Proof of Theorem 1.2
Let be a stable compact minimal immersion of codimension and dimension , where is any Riemannian manifold of dimension .
Let , be an orthonormal basis of , and be an orthonormal basis of . Let be the projection of onto and be the projection of onto . Similarly, let be the projection of onto and be the projection of onto . We identify with .
For each , we identify with . Same as in Section 3.1.2, we denote by the projection of onto the normal space in . As in the complex and quaternionic cases, we are going to use the normal sections in the second variation formula, where runs over an orthonormal basis of .
Lemma 3.9.
Same assumptions as above. Let be the usual canonical basis of . Then,
(17) | ||||
(18) |
where is the octonionic line containing as defined above when . When , we just let be any octonionic line. Same for .
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) |
Using Equation (16), we have
and
.
Now applying the last two equalities in Equation (19), we obtain
(20) |
We look at the summand. Notice that each term in the summand is invariant under isometries. Since is isomorphic to the isotropy group at and acts transitively and effectively on , there exists some such that . We then have for some , and for some .
Also note that the setwise stabilizer of the 8-dimensional subspace is isomorphic to . For the vectors , , let be an element of such that and we take .
Now using Equation (15) for the curvature tensor, we have
Since form a basis of ), we have
so looking at the projection onto yields
and taking inner product of with itself gives
(22) |
Similarly, using instead of gives
(23) |
Equation (18) follows from symmetry between and in Equation (20) and Equation (15) for the curvature tensor.
∎
Proposition 3.10.
Let be nonzero vectors in . Then
Proof.
Group the ’s according to the octonionic lines they belong to. Let be the resulting distinct octonionic lines. For each , let be the vectors among that are on the line . Let be the matrix
Since is a Gram matrix, it is symmetric positive semi-definite. There thus exists some orthogonal matrix of size such that is diagonal. Further, since the columns of are from the same octonionic line, has at most rank . Thus we can assume that , where are the eigenvalues of . The matrix identity
tells us that the matrix has orthogonal columns, and all but the first eight columns are zero vectors. Since and have the same column span, there exist orthonormal vectors such that if we define an matrix and an matrix as
then . This gives .
The strategy for the proof is to rewrite
in terms of the eigenvalues so that we get a nicer expression, for which we can do maximization. In doing so, we need the ’s to be octonionic lines instead of aribitrary 8-dimensional spaces, and Lemma 3.8 plays a crucial role.
We first compute
Lemma 3.8 shows that for some orthogonal matrix of size , and some constant . Also, and . In particular, when , we have and . Thus
(24) |
We then look at the quantity
.
Notice that since , we have
.
On the other hand, since is a symmetric matrix, its trace equals to the sum of its eigenvalues. Thus . Using this and the identities and , we get
(25) |
Since
summing over yields
(26) | ||||
where in the second to last step we used the fact that is an orthogonal matrix.
We maximize subject to the constraint
for some constant . Using Lagrange multipliers, at the maximum of subject to the constraint, there exists some such that for all ,
(27) |
Using (26) and summing over yields
Thus either , or for all .
If , then by Equation (27), we have for all . Thus
If instead for all , then
In either case, at the maximum of subject to , we have .
Since is arbitrary, we have always, which gives the desired inequality. ∎
Remark 3.11.
Notice that in the above proposition, if , then the proof is significantly simpler: just use the trivial estimate and apply Proposition 2.3.
Now recall is a stable compact minimal immersion. The stability of requires
Combining the above two inequalites, we must have
Thus for all ,
But since we have , this shows that we necessarily have
as well.
For the product manifold , let
denote the projection maps. Let , then . For , we claim that . In fact, since , for each , we have
Therefore, . This shows , so . Thus is an invariant submanifold of .
Analogously, as in the quaternionic case, by Theorem 1 in [XN00], is isometric to a product manifold , where is an immersion into and is an immersion into . Since is a stable compact minimal immersion, we further have that and are stable compact minimal immersions into and , respectively. However, Ohnita [Ohn86, Theorem E] showed that the only nontrivial stable minimal immersed submanifolds of are precisely the octonion protective lines . 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 . 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 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.