Hamiltonian Stationary Lagrangian Surfaces with Non-Negative Gaussian Curvature in Kähler-Einstein Surfaces
Abstract
In this paper, we give some simple conditions under which a Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold must have a Euclidean factor or be a fiber bundle over a circle. We also characterize the Hamiltonian stationary Lagrangian surfaces whose Gaussian curvature is non-negative and whose mean curvature vector is in some space when the ambient space is a simply connected complex space form.
1 Introduction
Let be a Kähler manifold of complex dimension . carries a natural symplectic structure given by the closed -form which is defined by for . We say that a Lagrangian submanifold is Hamiltonian stationary if it is a critical point of the volume functional under compactly supported Hamiltonian deformations, i.e. variations for which the variational vector field is of the form for some . In [14], Oh calculated the Euler-Lagrange equation of the variational problem and found that Hamiltonian stationary Lagrangian submanifolds are characterised by
or equivalently by
where denotes the mean curvature vector of , which we define as the trace of its second fundamental form , i.e. , is the differential -form on defined by and is the co-differential operator on induced by the metric .
By a theorem of Dazord (see, for example, Theorem 2.1 in [14]), in any Kähler manifold , the restriction of the Ricci form of to is given by . When is Kähler-Einstein, i.e. for some constant , then the differential -form is closed and thus defines a cohomology class in H on any Lagrangian submanifold. Therefore, is both closed and co-closed, hence harmonic, on any Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold.
In [2], Arsie proved that when is a Calabi-Yau manifold, then represents an integral cohomology class of called the Maslov class. Therefore, we will refer to as the Maslov form of .
Let denote the normal bundle of in and denote the collection of smooth sections of . Also, for any point and vector , let denote the projection of onto and let denote the Levi-Civita connection on . Then, there is a connection in that is given by
for any normal vector field and tangent vector . We say that a normal vector field is parallel if .
For any point and vector , let denote the projection of onto . Then, since is Lagrangian and , we have that
for any normal vector field and tangent vector . Therefore, is parallel if and only if is parallel.
We say that has parallel second fundamental form if
vanishes for all .
We present our results in two separate sections. In Section 2, we consider a complete, connected Hamiltonian stationary Lagrangian submanifold of arbitrary dimension inside a Kähler-Einstein manifold. We introduce a set of conditions, most of which consist of the non-negativity of the Ricci curvature of in the direction of and some pointwise or integral control over the absolute value of , that allows us to combine the Bochner formula for the harmonic -form and some Liouville-type theorems to deduce that must be parallel in the normal bundle of . The existence of a non-trivial global parallel vector field can restrict both the topology and the geometry of a manifold significantly. For example, if is simply connected, then it must be isometric to a Riemannian product of the form . As for a purely topological consequence, if is not diffeomorphic to such a product, then it must admit a circle action whose orbits are not homologous to zero. In Section 3, we restrict our attention to the case when . We also strengthen our assumptions by requiring that our surface has non-negative Gaussian curvature which allows us to describe explicitly all complete, connected Hamiltonian stationary Lagrangian surfaces in , and in that has non-negative Gaussian curvature and whose mean curvature vector is in some space.
The author would like to express his gratitude to Prof. Jingyi Chen for his invaluable suggestions and support, which were essential to the completion of this paper.
2 Hamiltonian Stationary Lagrangians in Kähler-Einstein Manifolds and the Bochner Method
Let denote any of the following sets of assumptions:
-
1.
and for some ;
-
2.
has non-negative Ricci curvature and for some ;
-
3.
is oriented, and as where is the distance function on relative to a fixed point ;
-
4.
and there exists a point , a non-decreasing function , constants and such that for all and
(1) whenever . Here, denotes the geodesic ball in of radius around the point ;
-
5.
has conformal Maslov form, i.e. the vector field is conformal.
If at least one of the sets of assumptions labelled (1)–(5) is satisfied, we say that is satisfied.
We can state the main result of this section as follows.
Theorem 2.1.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold. If satisfies , then
-
(a)
is parallel and thus has constant length;
-
(b)
vanishes identically, so if there exists a point such that is non-degenerate, then must be minimal;
-
(c)
and the scalar curvature of must be constant along the integral curves of .
The growth bound (1) from condition (4) is satisfied, for example, when has quadratic volume growth and does not grow faster than at infinity for some . In particular, it is satisfied when has quadratic volume growth and . Therefore, we have the following corollary.
Corollary 2.2.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold. If , and has quadratic volume growth, then the conclusions of Theorem 2.1 hold; in particular, must be constant.
Remark.
This phenomenon is related to the notion of parabolicity of a manifold. We say that a manifold is (strongly) parabolic if it does not admit a negative, non-constant subharmonic function, i.e. if and , then it must be constant. It is easy to see that a parabolic manifold does not admit a non-constant subharmonic function that is bounded from above. A sufficient condition111As it is discussed in [6] after Corollary 7.7, when the Ricci curvature is non-negative, then this condition is also necessary. for the parabolicity of a manifold was given by Karp in [9], which implies, for example, that every complete, non-compact manifold with quadratic volume growth is parabolic.
The main restriction imposed on by the conclusion of the Theorem 2.1 is that is parallel since the existence of a non-trivial global parallel vector field restricts the topology of a manifold significantly. For example, the following result of Welsh [19], states that the existence of a complete non-trivial global parallel vector field forces the existence of a circle action whose orbits are not real homologous to zero. By a complete vector field, we mean a vector field whose integral curves are defined for all time.
Theorem 2.3 (Welsh [19]).
Suppose that is a Riemannian manifold that admits a non-zero complete parallel vector field. Then either is diffeomorphic to the product of a Euclidean space with some other manifold, or else there is a circle action on whose orbits are not real homologous to zero. Moreover, if is not diffeomorphic to the product of a Euclidean space with some other manifold and its first integral homology class is finitely generated, then is a fiber bundle over a circle with finite structure group.
Corollary 2.4.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold. If is not minimal and it satisfies , then is diffeomorphic to the product of a Euclidean space with some other manifold or there is a circle action on whose orbits are not real homologous to zero. Moreover, it satisfies the conclusion of Theorem 2.1; and if it is not diffeomorphic to the product of a Euclidean space with some other manifold and its first integral homology class is finitely generated, then is a fiber bundle over a circle with finite structure group.
Proof.
Suppose that is a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold that satisfies . Then all the assumptions of Theorem 2.1 are satisfied thus all of its conclusions hold. In particular, is parallel so its integral curves are geodesics. Since is complete, all of its geodesics are defined for all and we see that is a complete vector field. Therefore, when is not minimal, is a non-zero complete parallel vector field on and we can apply Theorem 2.3 to finish the proof. ∎
When is not minimal, we can use Corollary 2.4 to establish the existence of a circle action on whose orbits are not real homologous to zero but only if is not diffeomorphic to a product of and some other manifold. It turns out that when is simply connected then the existence of such a splitting is guaranteed. Moreover, can be split in such a way isometrically.
Corollary 2.5.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold. If is not minimal and it satisfies , then its universal cover equipped with the pull-back metric is isometric to for some totally geodesic submanifold of .
Proof.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold that satisfies . Suppose that is not minimal. Then, by Theorem 2.1, is a non-zero parallel vector field.
Let denote the universal cover of which we equip with the pull-back metric. This makes into a Riemannian covering. We know that and since is a local isometry, we must also have for . Define . Then so must also be a parallel vector field. Since is simply connected and is closed, there exists a smooth function such that or equivalently .
In order to prove Theorem 2.1, we need the following lemma.
Lemma 2.6.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold. If satisfies , then is constant.
Proof.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold. Then is closed so we can apply the Bochner formula [15, p. 207] for to get
First, we observe that if is one of the conditions (1)–(4) , then and the function is clearly subharmonic. Since compact manifolds do not admit non-constant subharmonic functions, we may assume without the loss of generality that is non-compact when is one of the conditions (1)–(4).
If is in for some , then by a well-known result of Yau [21], is constant. Therefore, condition (1) implies that is constant.
If has non-negative Ricci curvature, then by a result of Li and Schoen (Theorem 2.2. in [12]) it does not admit any non-negative subharmonic function for all . Thus, condition (2) implies that is constant.
Now, assume that condition (3) is satisfied. In [1], Alías, Caminha and do Nascimento prove that every non-negative subharmonic function that converges to 0 at infinity on a connected, oriented, complete and non-compact Riemannian manifold must be identically zero. Applying this maximum principle to the function gives us that . Therefore, we conclude that condition (3) also implies that is constant.
Next, assume that condition (4) is satisfied. Since is a non-negative and non-decreasing function,
whenever . Therefore,
However, in [9], Karp showed that every non-negative non-constant subharmonic function on a complete non-compact Riemannian manifold satisfies
for all and center . Therefore, must be constant and we can conclude that condition (4) also implies that is constant.
Finally, suppose that is conformal. Then, since is divergence-free,
Therefore, the vector field is in fact Killing and the tensor is skew-symmetric. Since the dual -form is closed, we also know that is symmetric, and hence it must be zero. Therefore, is parallel which implies that it must also have constant length. We can conclude that if has conformal Maslov class, then must be constant, which completes the proof. ∎
Now, we can prove Theorem 2.1.
Proof of Theorem 2.1.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein manifold that satisfies . Then, as in the proof of Lemma 2.6, we have that
(2) |
By Lemma 2.6, is constant so the left-hand side of equation (2) vanishes identically. When is one of the conditions (1)–(4), then we have two non-negative terms on the right-hand side so they must each vanish identically, i.e. we must have that and . When is conformal, then by the same argument that we used in the proof of Lemma 2.6, is parallel. So which forces . Therefore, we can conclude that is parallel and is identically zero whenever is satisfied. This proves .
Recalling the Weitzenböck formula [15, p. 211], we have
where is the Hodge-Laplacian acting on differential -forms. Since is both harmonic and parallel, we have that
and thus must also vanish identically. Let us also assume that there exists a point such that is non-degenerate. Since , we must have that . However, we know that has constant length so must vanish identically and thus is minimal. This proves .
Let us also recall the contracted Bianchi identity (Proposition 7.18. [11])
(3) |
where is the scalar curvature of . The trace is taken on the first and the third indices, i.e. given a local orthonormal frame , equation (3) reads as
Therefore, plugging into equation (3) gives us that
The first and the third terms vanish since , while the second term is zero because is parallel. So we conclude that the scalar curvature must be constant along the integral curves of , which completes the proof of . ∎
3 Hamiltonian Stationary Lagrangian Surfaces with Non-Negative Gaussian Curvature in Kähler-Einstein surfaces
Let denote a complete connected Hamiltonian stationary Lagrangian surface isometrically immersed in a Kähler-Einstein surface .
In order to obtain a characterization that is more explicit than the one given by Theorem 2.1, we adjust our sets of assumptions from before. Let denote any of the following sets of assumptions:
-
1.
has non-negative Gaussian curvature and for some ;
-
2.
is oriented, it has non-negative Gaussian curvature and as ;
-
3.
has non-negative Gaussian curvature and the growth condition (1) is satisfied;
-
4.
has non-negative Gaussian curvature and conformal Maslov form.
If at least one of the sets of assumptions labelled (1)–(4) is satisfied, we say that is satisfied.
We will treat the cases when is compact and when it is non-compact separately.
Theorem 3.1.
Let be a Kähler-Einstein surface and let be a closed connected Hamiltonian stationary Lagrangian surface in that satisfies . If is orientable, then it is
-
•
a flat torus or
-
•
a minimal sphere.
If is not orientable, then it is
-
•
a flat Klein bottle or
-
•
a minimal projective plane.
In both cases, has parallel mean curvature.
Theorem 3.2.
Let be a Kähler-Einstein surface. If is a complete, connected non-compact Hamiltonian stationary Lagrangian surface in satisfying , then it has parallel mean curvature and it is
-
•
isometric to ,
-
•
diffeomorphic to and is minimal or
-
•
it is flat and its fundamental group is isomorphic to .
We start by proving the following lemma.
Lemma 3.3.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein surface . If satisfies then it has parallel mean curvature and it is also flat or minimal.
Proof.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein surface . Since dim, its curvature is entirely determined by its Gaussian curvature and, in particular,
Suppose that satisfies . It is easy to see that is stronger than so we can apply Theorem 2.1 which tells us that is parallel and that
Since is parallel, it has constant norm and therefore must be minimal or flat. ∎
Proof of Theorem 3.1.
Let be a closed, connected Hamiltonian stationary Lagrangian submanifold of a Kähler-Einstein surface . Also assume that satisfies . Then, by Lemma 3.3, has parallel mean curvature and it must also be minimal or flat.
First, Suppose that is orientable. Then, by the Gauss-Bonnet theorem,
(4) |
where is the Euler characteristic of . Since , where is the genus of , and is non-negative, we see that the genus must be or . Therefore, is diffeomorphic either to a sphere or to a torus respectively. Equation (4) also tells us that is flat if and only if it has genus , i.e. it is a torus. So, if is not flat then it is not just minimal but it must also have genus and thus it must be a minimal sphere. This completes the proof of the case when is orientable.
Now, suppose that is not orientable. In this case, , where is the non-orientable genus of which can be defined as the number of copies of appearing when the surface is represented as a connected sum of projective planes. Also, equation (4) still holds if we interpret the left-hand side as an integral of a density. One can easily see this by passing to the orientable double cover equipped with the pull-back metric. So, similarly to the orientable case, we have that must be or and hence is diffeomorphic either to a real projective plane or to a Klein bottle respectively. We also see that is flat if and only if it is a Klein bottle. Therefore, when is not flat, then it is not just minimal but must also have non-orientable genus and thus it is a minimal real projective plane. This completes the proof of the non-orientable case. ∎
Proof of Theorem 3.2.
It is known that the fundamental group of a non-compact surface is free (see, for example, [17, p. 142]). The first singular homology group of with coefficients in is the abelianization of its fundamental group so is the free abelian group on the generator set of . Since is free, we know that , where is the first Betti number of . Therefore, the cardinality of the generator set of is equal to .
First, assume that is orientable. Since , a result of Huber (Theorem 13. in [7]) tells us that is finitely connected, i.e. it is homeomorphic to a closed surface with finitely many punctures. Therefore, must be finite. Moreover, since the top homology group of a non-compact manifold vanishes identically, we have that and thus . Also, by Theorem 10. in [7],
(5) |
so we have
If is not orientable, then applying the same argument but to the orientable double cover of also yields .
3.1 Hamiltonian Stationary Lagrangian Surfaces with Non-Negative Gaussian Curvature in Complex Space Forms
Let be a complete, connected complex space form of complex dimension and constant holomorphic sectional curvature . Let be a Lagrangian submanifold.
We have the Wintgen-type inequality (Lemma 2.4. in [13]),
(6) |
where is a normalized (partial) normal scalar curvature222A similar inequality can be obtained using the Chen-Ricci inequality presented, for example, in [5]..
First, we look at the case .
Theorem 3.4.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold. If is satisfied, then has parallel second fundamental form. Moreover, when the ambient manifold is , then is either
-
•
a Lagrangian plane,
-
•
a Riemannian product of a circle and a line (a Lagrangian cylinder),
-
•
or a Riemannian product of two circles (possibly of different radii).
Proof.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold. Then, by Lemma 3.3, has parallel mean curvature and it must also be minimal or flat. Also, by (6),
(7) |
so we see that if is minimal, then it must also be flat. Therefore, we may assume, without a loss of generality, that is flat. By Theorem 2.6. in [10, p. 207], so we can conclude that has parallel second fundamental form.
For the rest of the proof, we assume that is . Let be a local orthonormal frame on . Then is a local orthonormal frame on and the components of the second fundamental form of in are given by
Let denote the matrix . Then, since is flat, by Lemma 2.5. in [10, p. 206], its second fundamental form commutes, i.e. for all . Therefore, by Theorem 2.9. in [10, p. 210], is congruent to one of the following standard Lagrangian submanifolds:
-
1.
(a Lagrangian plane),
-
2.
for some (a Lagrangian cylinder),
-
3.
for some (a product of two circles).
∎
Before looking at the cases and , we state some simple corollaries of Theorem 3.4.
Corollary 3.5.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold that has non-negative Gaussian curvature and for some . Then it is either a Lagrangian plane, a Lagrangian cylinder or a product of two circles. Moreover, it can only be a cylinder when .
Proof.
Let be as stated in the corollary. When , then is clearly satisfied. Since , we know by the Bishop-Gromov volume comparison theorem that and thus has quadratic volume growth. Therefore, as discussed in the previous section after Theorem 2.1, the growth condition (1) is satisfied whenever . So is satisfied when as well and we can use Theorem 3.4 to conclude that must be a Lagrangian plane, a Lagrangian cylinder or a product of two circles for any .
Finally, we note that when is non-compact, then it has infinite volume [21] so it must be minimal if it has a constant mean curvature that is in for some . The standard Lagrangian cylinder in has constant mean curvature but it is neither compact nor minimal so it can only occur when . ∎
We say that a complete non-compact submanifold is asymptoticaly minimal if its mean curvature vector converges to at infinty, i.e. as .
Corollary 3.6.
The only complete, connected, oriented and asymptotically minimal Hamiltonian stationary Lagrangian submanifolds of with non-negative Gaussian curvature are Lagrangian planes.
Since a complete Kähler manifold of positive holomorphic sectional curvature is necessarily simply connected (see, for example, [18]), we may assume that is equipped with the standard Fubini-Study metric which has constant holomorphic sectional curvature . Let be the unit sphere in equipped with induced metric. Then the map given by , which is usually referred to as the Hopf fibration, can be used to construct Lagrangian immersions into . For more details, see, for example, §3. in [3].
Theorem 3.7.
Let be a closed connected Hamiltonian stationary Lagrangian submanifold with non-negative Gaussian curvature. Then is
-
•
a totally geodesic or
-
•
flat and is locally congruent to the image of where is the Hopf fibration and is given by with
for some real constants and .
Proof.
Let be a closed connected Hamiltonian stationary Lagrangian submanifold with non-negative Gaussian curvature. Then is satisfied so by Theorem 3.1, has parallel mean curvature and it is a minimal sphere, a minimal real projective plane, a flat Klein bottle or a flat torus. If is a minimal sphere or a minimal real projective plane then, by Theorem 7. in [20], is immersed in such a way that its image is a totally geodesic . If is a flat Klein bottle or a flat torus then has parallel second fundamental form by Theorem 2.6. in [10, p. 207]. Therefore, the result follows from the classification of submanifolds with parallel second fundamental forms in given by Theorem 7.1. in [4].
∎
Finally, we consider the case . Let denote the complex hyperbolic space of constant holomorphic sectional curvature , let denote equipped with the psuedo-Euclidean metric and set . Then the map given by , which we will also refer to as the Hopf fibration, can be used to construct Lagrangian immersions into . For more details, see, for example, §3. in [3].
Theorem 3.8.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold for some . If is satisfied, then is flat and has parallel second fundamental form. Moreover, when the ambient manifold is , then is locally congruent to the image of where is the Hopf fibration and is given by where
-
1.
,
and
for some real constants and satisfying and ; -
2.
,
and
for a real number ; -
3.
,
and
for some real constants and satisfying and ; -
4.
,
and
for some real constants and satisfying ; -
5.
,
and
for a real number ; or -
6.
,
and
.
Proof.
Let be a complete, connected Hamiltonian stationary Lagrangian submanifold for some . Suppose also that is satisfied. Then by Lemma 3.3, has parallel mean curvature and it is also flat or minimal. However, since , it is clear from (6) that cannot be minimal. Therefore, must be flat and thus it has parallel second fundamental form by Theorem 2.6. in [10, p. 207]. When is , the result follows from the classification of submanifolds with parallel second fundamental forms in given by Theorem 7.2. in [4]. ∎
References
- Alías et al. [2019] Luis José Alías, Antonio Caminha, and FY do Nascimento. A maximum principle at infinity with applications to geometric vector fields. Journal of Mathematical Analysis and Applications, 474(1):242–247, 2019.
- Arsie [2000] Alessandro Arsie. Maslov class and minimality in Calabi–Yau manifolds. Journal of Geometry and Physics, 35(2-3):145–156, 2000.
- Chen [1997] Bang-Yen Chen. Interaction of Legendre curves and Lagrangian submanifolds. Israel Journal of Mathematics, 99:69–108, 1997.
- Chen et al. [2010] Bang-Yen Chen, Franki Dillen, and Joeri Van der Veken. Complete classification of parallel Lorentzian surfaces in Lorentzian complex space forms. International Journal of Mathematics, 21(05):665–686, 2010.
- Deng [2009] Shangrong Deng. An improved Chen-Ricci inequality. Int. Electron. J. Geom, 2(2):39–45, 2009.
- Grigor’Yan [1999] Alexander Grigor’Yan. Analytic and geometric background of recurrence and non-explosion of the Brownian motion on Riemannian manifolds. Bulletin of the American Mathematical Society, 36(2):135–249, 1999.
- Huber [1958] Alfred Huber. On subharmonic functions and differential geometry in the large. Commentarii Mathematici Helvetici, 32(1):13–72, 1958.
- Innami [1982] Nobuhiro Innami. Splitting theorems of Riemannian manifolds. Compositio Mathematica, 47(3):237–247, 1982.
- Karp [1982] Leon Karp. Subharmonic functions on real and complex manifolds. Mathematische Zeitschrift, 179(4):535–554, 1982.
- Kon and Yano [1985] Masahiro Kon and Kentaro Yano. Structures on manifolds, volume 3. World scientific, 1985.
- Lee [2018] John M Lee. Introduction to Riemannian manifolds, volume 176. Springer, 2018.
- Li and Schoen [1984] Peter Li and Richard Schoen. and mean value properties of subharmonic functions on Riemannian manifolds. Acta Mathematica, 153(1):279–301, 1984.
- Mihai [2014] Ion Mihai. On the generalized Wintgen inequality for Lagrangian submanifolds in complex space forms. Nonlinear Analysis: Theory, Methods & Applications, 95:714–720, 2014.
- Oh [1993] Yong-Geun Oh. Volume minimization of Lagrangian submanifolds under Hamiltonian deformations. Mathematische Zeitschrift, 212(1):175–192, 1993.
- Petersen [2006] Peter Petersen. Riemannian geometry, volume 171. Springer, 2006.
- Sakai [1996] Takashi Sakai. On Riemannian manifolds admitting a function whose gradient is of constant norm. Kodai Mathematical Journal, 19(1):39–51, 1996.
- Stillwell [2012] John Stillwell. Classical topology and combinatorial group theory, volume 72. Springer Science & Business Media, 2012.
- Tsukamoto [1957] Yôtarô Tsukamoto. On Kählerian manifolds with positive holomorphic sectional curvature. Proceedings of the Japan Academy, 33(6):333–335, 1957.
- Welsh [1986] David J Welsh. On the existence of complete parallel vector fields. Proceedings of the American Mathematical Society, 97(2):311–314, 1986.
- Yau [1974] Shing-Tung Yau. Submanifolds with constant mean curvature. American Journal of Mathematics, 96(2):346–366, 1974.
- Yau [1976] Shing-Tung Yau. Some function-theoretic properties of complete Riemannian manifold and their applications to geometry. Indiana University Mathematics Journal, 25(7):659–670, 1976.