Skip to main content
Log in

Self-shrinkers for the mean curvature flow in arbitrary codimension

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

In this paper, we generalize Colding–Minicozzi’s recent results about codimension-1 self-shrinkers for the mean curvature flow to higher codimension. In particular, we prove that the sphere \({bf S}^{n}(\sqrt{2n})\) is the only complete embedded connected \(F\)-stable self-shrinker in \(\mathbf{R}^{n+k}\) with \(\mathbf{H}\ne 0\), polynomial volume growth, flat normal bundle and bounded geometry. We also discuss some properties of symplectic self-shrinkers, proving that any complete symplectic self-shrinker in \(\mathbf{R}^4\) with polynomial volume growth and bounded second fundamental form is a plane. As a corollary, we show that there is no finite time Type I singularity for symplectic mean curvature flow, which has been proved by Chen–Li using different method. We also study Lagrangian self-shrinkers and prove that for Lagrangian mean curvature flow, the blow-up limit of the singularity may be not \(F\)-stable.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Altschuler, S.J.: Singularities of the curve shrinking flow for space curves. J. Differ. Geom. 34, 491–514 (1991)

    MathSciNet  MATH  Google Scholar 

  2. Anciaux, H.: Construction of Lagrangian self-similar solutions to the mean curvature flow in \(C^n\). Geom. Dedicata 120, 37–48 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  3. Andrews, B., Li, H., Wei, Y.: \(F\)-stability for self-shrinking solutions to mean curvature flow. arXiv:1204.5010

  4. Arezzo, C., Sun, J.: Conformal solitons to the mean curvature flow and minimal submanifolds. Math. Nachr. (to appear)

  5. Cao, H., Li, H.: A gap theorem for self-shrinkers of the mean curvature flow in arbitrary codimension. Calc. Var. doi:10.1007/s00526-012-0508-1 (2012)

  6. Chen, J., Li, J.: Mean curvature flow of surfaces in 4-manifolds. Adv. Math. 163, 287–309 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chen, J., Li, J.: Singularity of mean curvature flow of Lagrangian submanifolds. Invent. Math. 156, 25–51 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  8. Colding, T.H., Minicozzi II, W.P.: Generic mean curvature glow I: generic singularities. Ann. Math. 175, 755–833 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  9. Colding, T., Minicozzi, W.: Minimal surfaces. Courant Lecture Notes in Mathematics, no. 4. New York University (1998)

  10. Ecker, K., Huisken, G.: Mean curvature evolution of entire graphs. Ann. Math. 130, 453–471 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  11. Groh, K., Schwarz, M., Smoczyk, K., Zehmisch, K.: Mean curvature flow of monotone Lagrangian submanifolds. Math. Z. 257(2), 295–327 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  12. Han, X., Li, J.: On symplectic curvature flows. Lecture Notes

  13. Han, X., Li, J., Sun, J.: The second type singularities of symplectic and Lagrangian mean curvature flows. Chi. Ann. Math. Ser. B 32, 223–240 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  14. Han, X., Sun, J.: Translating solitons to symplectic mean curvature flows. Ann. Glob. Anal. Geom. 38, 161–169 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  15. Huisken, G.: Flow by mean curvature of convex surfaces into spheres. J. Differ. Geom. 20, 237–266 (1984)

    MathSciNet  MATH  Google Scholar 

  16. Huisken, G.: Asymptotic behavior for singularities of the mean curvature flow. J. Differ. Geom. 31(1), 285–299 (1990)

    MathSciNet  MATH  Google Scholar 

  17. Huisken, G.: Local and global behaviour of hypersurfaces moving by mean curvature. In: Proceedings of Symposia in Pure Mathematics, vol. 54, Part I, pp. 175–191 (1993)

  18. Iimanen, T.: Singularities of mean curvature flow of surfaces. Preprint

  19. Le, N.Q., Sesum, N.: Blow-up rate of the mean curvature during the mean curvature flow and a gap theorem for self-shrinkers. Commun. Anal. Geom. 19(4), 633–659 (2011)

    MathSciNet  MATH  Google Scholar 

  20. Lee, Y.-I., Lue, Y.-K.: The stability of self-Shrinkers of mean curvature flow in higher codimension. arXiv:1204.6116

  21. Neves, A.: Singularities of Lagrangian mean curvature flow: zero-Maslov class case. Invent. Math. 168(3), 449–484 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  22. Oh, Y.-G.: Second variation and stabilities of minimal lagrangian submanifolds in Kähler manifolds. Invent. Math. 101, 501–519 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  23. Smoczyk, K.: Self-shrinkers of the mean curvature flow in arbitrary codimension. Int. Math. Res. Not. 48, 2983–3004 (2005)

    Article  MathSciNet  Google Scholar 

  24. Smoczyk, K.: A relation between mean curvature flow solitons and minimal submanifolds. Math. Nachr. 229, 175–186 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  25. Wang, L.: A Bernstein type theorem for self-similar shrinkers. Geom. Dedicata 151, 297–303 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  26. Wang, M.-T.: Mean curvature flow of surfaces in Einstein four manifolds. J. Differ. Geom. 57, 301–338 (2001)

    MATH  Google Scholar 

  27. Zhang, Y.B.: \(F\)-stability of self-shrinker solutions to harmonic map heat flow. Calc. Var. 45, 347–366 (2012)

Download references

Acknowledgments

We wish to thank Professor Jiayu Li for interesting and helpful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Claudio Arezzo.

Appendices

Appendix A

In this appendix, we will prove the variations of normal vector field and mean curvature we need in Sect. 3. The proof is standard. When the variation vector \(\mathbf{V }\) is the mean curvature vector \(\mathbf{H }\), they are proved in [12]. We will follow the computations in [12].

We begin with fixing our notation. In a normal coordinate around some point in \(\Sigma \), the induced metric on \(\Sigma \) is given by

$$\begin{aligned} g_{ij}=\langle \partial _i F,\partial _j F\rangle , \end{aligned}$$
(8.1)

where \(\partial _i(i=1,\ldots ,n)\) are the partial derivatives with respect to the local coordinate. Here, \(\langle \cdot ,\cdot \rangle \) is the inner product of \(\mathbf{R }^{n+k}\).

We choose a local field of orthonormal frames \(e_1, \cdots , e_n, e_{n+1}, \cdots , e_{n+k}\) of \(\mathbf{R }^{n+k}\) along \(\Sigma _s\) such that \(e_1,\ldots , e_n\) are tangent vectors of \(\Sigma _s\) and \(e_{n+1},\ldots ,e_{n+k}\) are in the normal bundle over \(\Sigma _s\). From now on, we will agree on the following index ranges:

$$\begin{aligned} 1\le i,j,k,l \le n, \ \ n+1\le \alpha ,\beta ,\gamma \le n+k, \ \ 1\le A,B,C \le n+k. \end{aligned}$$

We can write

$$\begin{aligned} \mathbf{A }=\mathbf{A }^{\alpha }e_{\alpha }, \ \ \ \mathbf{H }=-H^{\alpha }e_{\alpha }. \end{aligned}$$

Let \(\mathbf{A }^{\alpha }=(h^{\alpha }_{ij})\), where \((h^{\alpha }_{ij})\) is a matrix. By the Weingarten equation, we have

$$\begin{aligned} h^{\alpha }_{ij}=\langle \bar{\nabla }_{e_i}e_{\alpha },e_j\rangle =-\langle e_{\alpha },\bar{\nabla }_{e_i}e_j\rangle =h^{\alpha }_{ji}, \end{aligned}$$

where \(\bar{\nabla }\) is the Levi-Civita connection on \(\mathbf{R }^{n+k}\). Furthermore,

$$\begin{aligned} H^{\alpha }=g^{ij}h^{\alpha }_{ij}=h^{\alpha }_{ii}. \end{aligned}$$

Suppose the variation vector filed is \(\mathbf{V }=V^{\alpha }e_{\alpha }\), i.e.,

$$\begin{aligned} F(\cdot ,s):\Sigma \rightarrow \mathbf{R }^{n+k} \end{aligned}$$

satisfies

$$\begin{aligned} \frac{\partial F}{\partial s}=\mathbf{V }. \end{aligned}$$

Then we have

Lemma 8.1

The induced metric satisfies

$$\begin{aligned} \frac{\partial }{\partial s}g_{ij}=2V^{\alpha }h^{\alpha }_{ij}, \end{aligned}$$
(8.2)

and

$$\begin{aligned} \frac{\partial }{\partial s}g^{ij}=-2V^{\alpha }h^{\alpha }_{ij}. \end{aligned}$$
(8.3)

Proof

We prove it at a fixed point. We have

$$\begin{aligned} \frac{\partial }{\partial s}g_{ij}&= \bar{\nabla }_{\mathbf{V }}\langle \partial _i F,\partial _j F\rangle = \langle \bar{\nabla }_{\mathbf{V }} \partial _i F,e_j \rangle +\langle e_i,\bar{\nabla }_{\mathbf{V }} \partial _j F\rangle \nonumber \\&= \langle \bar{\nabla }_{e_i} \mathbf{V },e_j \rangle +\langle e_i,\bar{\nabla }_{e_j} \mathbf{V }\rangle \nonumber \\&= \langle \bar{\nabla }_{e_i} (V^{\alpha }e_{\alpha }),e_j \rangle +\langle e_i,\bar{\nabla }_{e_j} (V^{\alpha }e_{\alpha })\rangle \nonumber \\&= V^{\alpha }\langle \bar{\nabla }_{e_i} e_{\alpha },e_j \rangle +V^{\alpha }\langle e_i,\bar{\nabla }_{e_j} e_{\alpha }\rangle \nonumber \\&= 2V^{\alpha }h^{\alpha }_{ij}. \end{aligned}$$

Here we have used the fact that

$$\begin{aligned} \bar{\nabla }_{\mathbf{V }} \partial _i F=\frac{\partial ^2}{\partial s\partial x_i}F=\bar{\nabla }_{\partial _i F}\mathbf{V }. \end{aligned}$$

As at the fixed point \(p, g_{ij}(p)=\delta _{ij}\), we know that

$$\begin{aligned} \frac{\partial }{\partial s}g^{ij}=-2V^{\alpha }h^{\alpha }_{ij}. \end{aligned}$$

\(\square \)

Lemma 8.2

Denote \(\langle \frac{\partial }{\partial s}e_{\alpha },e_{\beta } \rangle =\langle \bar{\nabla }_{\mathbf{V }}e_{\alpha },e_{\beta } \rangle =b^{\beta }_{\alpha }\), then \(b^{\beta }_{\alpha }=-b_{\beta }^{\alpha }\), and we have

$$\begin{aligned} \frac{\partial }{\partial s}e_{\alpha }=-\nabla V^{\alpha }-V^{\beta }\langle \bar{\nabla }_{e_i}e_{\beta },e_{\alpha } \rangle e_i+b^{\beta }_{\alpha }e_{\beta }. \end{aligned}$$
(8.4)

Here, \(\nabla V^{\alpha }\) is the covariant differential for the induced metric on \(\Sigma _s\).

Proof

We have

$$\begin{aligned} \frac{\partial }{\partial s}e_{\alpha }&= \langle \frac{\partial }{\partial s}e_{\alpha },e_i \rangle e_i +\langle \frac{\partial }{\partial s}e_{\alpha },e_{\beta } \rangle e_{\beta } \nonumber \\&= \langle \bar{\nabla }_{\mathbf{V }} e_{\alpha },e_i \rangle e_i +b^{\beta }_{\alpha }e_{\beta }\nonumber \\&= -\langle e_{\alpha },\bar{\nabla }_{\mathbf{V }}\partial _i F \rangle e_i +b^{\beta }_{\alpha }e_{\beta }\nonumber \\&= -\langle e_{\alpha },\bar{\nabla }_{e_i}\mathbf{V }\rangle e_i +b^{\beta }_{\alpha }e_{\beta }\nonumber \\&= -\langle e_{\alpha },\bar{\nabla }_{e_i}(V^{\beta }e_{\beta })\rangle e_i +b^{\beta }_{\alpha }e_{\beta }\nonumber \\&= -\langle e_{\alpha },(\bar{\nabla }_{e_i}V^{\beta })e_{\beta } +V^{\beta }\bar{\nabla }_{e_i}e_{\beta }\rangle e_i +b^{\beta }_{\alpha }e_{\beta }\nonumber \\&= -(\bar{\nabla }_{e_i}V^{\alpha })e_i- V^{\beta } \langle \bar{\nabla }_{e_i}e_{\beta },e_{\alpha }\rangle e_i +b^{\beta }_{\alpha }e_{\beta }\nonumber \\&= -\nabla V^{\alpha }-V^{\beta }\langle \bar{\nabla }_{e_i} e_{\beta },e_{\alpha } \rangle e_i+b^{\beta }_{\alpha }e_{\beta }. \end{aligned}$$

\(\square \)

Lemma 8.3

The second fundamental form satisfies

$$\begin{aligned} \frac{\partial }{\partial s}h_{ij}^{\alpha }=-V^{\alpha }_{,ji}+V^{\beta }h_{ik}^{\alpha }h_{jk}^{\beta } +h_{ij}^{\beta }\langle e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle . \end{aligned}$$
(8.5)

Here, \(V^{\alpha }_{,ji}\) denotes the second covariant derivative for the connection on the normal bundle.

Proof

We compute at a fixed point \(p\in \Sigma \). We can choose a frame \(e_i\) so that \(\bar{\nabla }^T_{e_i}e_j(p)=0\), i.e., at \(p, \bar{\nabla }_{e_i}e_j=-h^{\beta }_{ij}e_{\beta }\). From

$$\begin{aligned} h^{\alpha }_{ij}=-\langle \bar{\nabla }_{e_i}e_j, e_{\alpha }\rangle , \end{aligned}$$

and the fact that \(\mathbf{R }^{n+k}\) is flat, we have

$$\begin{aligned} \frac{\partial }{\partial s}h^{\alpha }_{ij}&= -\langle \bar{\nabla }_{\mathbf{V }}\bar{\nabla }_{e_i}e_j, e_{\alpha }\rangle -\langle \bar{\nabla }_{e_i}e_j, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= -\langle \bar{\nabla }_{e_i}\bar{\nabla }_{\mathbf{V }}e_j, e_{\alpha }\rangle -\langle \bar{\nabla }_{[\mathbf{V },e_i]}e_j, e_{\alpha }\rangle -\langle \bar{\nabla }_{e_i}e_j, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= -\langle \bar{\nabla }_{e_i}\bar{\nabla }_{e_j}\mathbf{V }, e_{\alpha }\rangle -\langle -h^{\beta }_{ij}e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= -\langle \bar{\nabla }_{e_i}(\bar{\nabla }_{e_j}^T\mathbf{V }+\bar{\nabla }_{e_j}^{\perp }\mathbf{V }), e_{\alpha }\rangle +\langle h^{\beta }_{ij}e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= -\langle \bar{\nabla }_{e_i}(\bar{\nabla }_{e_j}^T\mathbf{V }), e_{\alpha }\rangle -\langle \bar{\nabla }_{e_i}^{\perp }\bar{\nabla }_{e_j}^{\perp }\mathbf{V }, e_{\alpha }\rangle +\langle h^{\beta }_{ij}e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= \langle \bar{\nabla }_{e_j}^T\mathbf{V }, \bar{\nabla }_{e_i}e_{\alpha }\rangle -\langle V^{\beta }_{,ji}e_{\beta }, e_{\alpha }\rangle +\langle h^{\beta }_{ij}e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= \langle \bar{\nabla }_{e_j}^T(V^{\beta }e_{\beta }), h^{\alpha }_{ik}e_k\rangle -V^{\beta }_{,ji}\langle e_{\beta }, e_{\alpha }\rangle +\langle h^{\beta }_{ij}e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= -V^{\alpha }_{,ji}+V^{\beta }\langle h^{\beta }_{jl}e_l, h^{\alpha }_{ik}e_k\rangle +\langle h^{\beta }_{ij}e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle \nonumber \\&= -V^{\alpha }_{,ji}+V^{\beta }h^{\alpha }_{ik}h^{\beta }_{jk} +h^{\beta }_{ij}\langle e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle . \end{aligned}$$

\(\square \)

Lemma 8.4

The mean curvature vector satisfies

$$\begin{aligned} \frac{\partial }{\partial s}\mathbf{H }=(\Delta V^{\alpha }+V^{\beta }h^{\beta }_{ij}h^{\alpha }_{ij})e_{\alpha }+H^{\alpha }\nabla V^{\alpha } +H^{\alpha }V^{\beta }\langle \bar{\nabla }_{e_i}e_{\beta },e_{\alpha }\rangle e_i. \end{aligned}$$
(8.6)

Proof

By Lemma 8.1 and Lemma 8.3, we have

$$\begin{aligned} \frac{\partial }{\partial s}H^{\alpha }&= \frac{\partial }{\partial s}\left(g^{ij}h^{\alpha }_{ij}\right) = \left(\frac{\partial }{\partial s}g^{ij}\right)h^{\alpha }_{ij}+g^{ij}\frac{\partial }{\partial s}h^{\alpha }_{ij}\nonumber \\&= -2V^{\beta }h^{\beta }_{ij}h^{\alpha }_{ij}+(-\Delta V^{\alpha }+V^{\beta }h^{\beta }_{ij}h^{\alpha }_{ij} +H^{\beta }\langle e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle )\nonumber \\&= -\Delta V^{\alpha }-V^{\beta }h^{\beta }_{ij}h^{\alpha }_{ij} +H^{\beta }\langle e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle . \end{aligned}$$
(8.7)

Combining with Lemma 8.2, we obtain

$$\begin{aligned} \frac{\partial }{\partial s}\mathbf{H }&= \frac{\partial }{\partial s}\left(-H^{\alpha }e_{\alpha }\right) =-\left(\frac{\partial }{\partial s}H^{\alpha }\right) e_{\alpha }-H^{\alpha }\frac{\partial }{\partial s}e_{\alpha } \nonumber \\&= \left(\Delta V^{\alpha }+V^{\beta }h^{\beta }_{ij}h^{\alpha }_{ij} -H^{\beta }\langle e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha } \rangle \right)e_{\alpha } \nonumber \\&+H^{\alpha }\nabla V^{\alpha }+H^{\alpha }V^{\beta }\langle \bar{\nabla }_{e_i}e_{\beta },e_{\alpha } \rangle e_i -H^{\alpha }b^{\beta }_{\alpha }e_{\beta }. \end{aligned}$$

Note that

$$\begin{aligned} -H^{\beta }\langle e_{\beta }, \bar{\nabla }_{\mathbf{V }}e_{\alpha }\rangle e_{\alpha } =-H^{\beta }b^{\beta }_{\alpha } e_{\alpha }=-H^{\alpha }b^{\alpha }_{\beta } e_{\beta }=H^{\alpha }b^{\beta }_{\alpha }e_{\beta }. \end{aligned}$$

Thus we have

$$\begin{aligned} \frac{\partial }{\partial s}\mathbf{H }=(\Delta V^{\alpha }+V^{\beta }h^{\beta }_{ij}h^{\alpha }_{ij})e_{\alpha }+H^{\alpha }\nabla V^{\alpha } +H^{\alpha }V^{\beta }\langle \bar{\nabla }_{e_i}e_{\beta },e_{\alpha }\rangle e_i. \end{aligned}$$

\(\square \)

Appendix B

In this appendix, we will give another two geometric identities satisfied on self-shrinkers with arbitrary dimension and codimension. These results generalized Theorem 5.2 and Lemma 10.8 of [8].

Suppose \(\Sigma ^n\subset \mathbf{R }^{n+k}\) is a self-shrinker. We choose a frame \(\{e_A\}_{A=1}^{n+k}\) on \(\mathbf{R }^{n+k}\) along \(\Sigma \) such that \(\{e_i\}_{i=1}^{n}\) are tangent to \(\Sigma \) and \(\{e_{\alpha }\}_{\alpha =n+1}^{n+p}\) are in the normal bundle. We will compute pointwise. So we will always choose the frame \(\{e_i\}_{i=1}^{n}\) such that \(\bar{\nabla }_{e_i}^Te_j(p)=0\), i.e., at \(p, \bar{\nabla }_{e_i}e_j=-h^{\alpha }_{ij}e_{\alpha }\).

Lemma 9.1

Let \(L\) be defined by (3.9). Suppose \(\mathbf{w }\in \mathbf{R }^{n+k}\) is a fixed vector. Then on a self-shrinker \(\Sigma ^n\) in \(\mathbf{R }^{n+k}\), we have

$$\begin{aligned} L\mathbf{w }^{\perp }=\frac{1}{2}\mathbf{w }^{\perp }. \end{aligned}$$
(9.1)

Proof

By definition,

$$\begin{aligned} \mathbf{w }^{\perp }=\langle \mathbf{w },e_{\alpha }\rangle e_{\alpha }\equiv f^{\alpha }e_{\alpha }, \end{aligned}$$

where

$$\begin{aligned} f^{\alpha }=\langle \mathbf{w },e_{\alpha }\rangle . \end{aligned}$$
(9.2)

Then computing at \(p\) using the above chosen frame, we have

$$\begin{aligned} f^{\alpha }_{,i}&= \langle \bar{\nabla }^N_{e_i}(f^{\gamma }e_{\gamma }),e_{\alpha }\rangle =e_i(f^{\gamma })\langle e_{\gamma },e_{\alpha }\rangle +f^{\gamma }\langle \bar{\nabla }^N_{e_i}e_{\gamma },e_{\alpha }\rangle \nonumber \\&= \bar{\nabla }_{e_i}f^{\alpha }+\langle \mathbf{w } ,e_{\gamma }\rangle \langle \bar{\nabla }^N_{e_i}e_{\gamma },e_{\alpha }\rangle \nonumber \\&= \bar{\nabla }_{e_i}\langle \mathbf{w } ,e_{\alpha }\rangle -\langle \mathbf{w } ,e_{\gamma }\rangle \langle e_{\gamma },\bar{\nabla }^N_{e_i}e_{\alpha }\rangle \nonumber \\&= \langle \mathbf{w },\bar{\nabla }_{e_i}e_{\alpha }\rangle -\langle \mathbf{w },\bar{\nabla }^N_{e_i}e_{\alpha }\rangle =h^{\alpha }_{ij}\langle \mathbf{w } ,e_j\rangle \end{aligned}$$
(9.3)

and

$$\begin{aligned} f^{\alpha }_{,ik}&= \langle \bar{\nabla }^N_{e_k}\bar{\nabla }^N_{e_i}(f^{\gamma }e_{\gamma }), e_{\alpha }\rangle =\langle \bar{\nabla }^N_{e_k}(f^{\gamma }_{,i}e_{\gamma }),e_{\alpha }\rangle \nonumber \\&= e_k(f^{\gamma }_{,i})\langle e_{\gamma },e_{\alpha }\rangle +f^{\gamma }_{,i}\langle \bar{\nabla }^N_{e_k}e_{\gamma },e_{\alpha } \rangle \nonumber \\&= e_k(f^{\alpha }_{,i}) +h^{\gamma }_{ij}\langle \mathbf{w } ,e_j\rangle \langle \bar{\nabla }^N_{e_k}e_{\gamma },e_{\alpha }\rangle \nonumber \\&= e_{k}(h^{\alpha }_{ij})\langle \mathbf{w } ,e_j\rangle +h^{\alpha }_{ij}\bar{\nabla }_{e_k}\langle \mathbf{w } ,e_j\rangle +h^{\gamma }_{ij}\langle \mathbf{w } ,e_j\rangle \langle \bar{\nabla }^N_{e_k}e_{\gamma },e_{\alpha }\rangle \nonumber \\&= \left(e_{k}(h^{\alpha }_{ij}) +h^{\gamma }_{ij}\langle \bar{\nabla }^N_{e_k}e_{\gamma },e_{\alpha }\rangle \right)\langle \mathbf{w } ,e_j\rangle -h^{\alpha }_{ij}\langle \mathbf{w } ,\bar{\nabla }_{e_k}e_j\rangle \nonumber \\&= h^{\alpha }_{ik,j}\langle \mathbf{w } ,e_j\rangle -h^{\alpha }_{ij}h^{\beta }_{kj}\langle \mathbf{w },e_{\beta }\rangle . \end{aligned}$$
(9.4)

Here, we have used (6.23) and Codazzi equation. Taking trace of (9.4) and using (6.25), we obtain

$$\begin{aligned} \Delta f^{\alpha } = H^{\alpha }_{,j}\langle \mathbf{w } ,e_j\rangle -h^{\alpha }_{ij}h^{\beta }_{ij}\langle \mathbf{w },e_{\beta }\rangle = \langle \mathbf{w } ,\nabla H^{\alpha }\rangle -f^{\beta }h^{\alpha }_{ij}h^{\beta }_{ij}. \end{aligned}$$
(9.5)

By (6.21) and (9.3), we have

$$\begin{aligned} \langle \mathbf{w } ,\nabla H^{\alpha }\rangle = H^{\alpha }_{,i}\langle \mathbf{w } ,e_i\rangle =\frac{1}{2}h^{\alpha }_{ij}\langle \mathbf{x } ,e_j\rangle \langle \mathbf{w } ,e_i\rangle = \frac{1}{2}f^{\alpha }_{,j}\langle \mathbf{x } ,e_j\rangle =\frac{1}{2}\langle \mathbf{x }, \nabla f^{\alpha }\rangle . \end{aligned}$$
(9.6)

Putting (9.6) into (9.5), we obtain

$$\begin{aligned} \Delta f^{\alpha }+f^{\beta }h^{\alpha }_{ij}h^{\beta }_{ij}-\frac{1}{2}\langle \mathbf{x }, \nabla f^{\alpha }\rangle +\frac{1}{2}f^{\alpha } =\frac{1}{2}f^{\alpha }. \end{aligned}$$
(9.7)

By definition of the operator \(L\), this is equivalent to (6.24). \(\square \)

The following result needs “flat normal bundle” assumption on the self-shrinker.

Lemma 9.2

If we extend the operator \(L\) to tensors, then on a self-shrinker \(\Sigma ^n\) in \({{\varvec{R}}}^{n+k}\) with flat normal bundle, we have

$$\begin{aligned} L{{\varvec{A}}}={{\varvec{A}}}. \end{aligned}$$
(9.8)

Proof

We will show that

$$\begin{aligned} (L\mathbf{A })^{\alpha }_{ij}=h^{\alpha }_{ij}. \end{aligned}$$
(9.9)

In general, we have the following Simons’ equality for the second fundamental form [12, 26]:

$$\begin{aligned} \Delta h^{\alpha }_{ij} = \nabla _i\nabla _jH^{\alpha }+H^{\beta }h^{\beta }_{il} h^{\alpha }_{lj}-h^{\beta }_{ik}h^{\beta }_{kl}h^{\alpha }_{lj} +2h^{\beta }_{ik}h^{\alpha }_{kl}h^{\beta }_{lj} -h^{\beta }_{ij}h^{\beta }_{kl}h^{\alpha }_{kl}-h^{\alpha }_{ik} h^{\beta }_{kl}h^{\beta }_{lj}.\qquad \quad \end{aligned}$$
(9.10)

Combining with (9.1), we have

$$\begin{aligned} (L\mathbf{A })^{\alpha }_{ij}&= \Delta h^{\alpha }_{ij}+h^{\beta }_{ij}h^{\beta }_{kl}h^{\alpha }_{kl} -\frac{1}{2}\langle \mathbf{x },\nabla h^{\alpha }_{ij}\rangle +\frac{1}{2}h^{\alpha }_{ij}\nonumber \\&= \frac{1}{2}\left(h^{\alpha }_{ij,k}\langle \mathbf{x } ,e_k\rangle +h^{\alpha }_{ij} -h^{\alpha }_{jk}h^{\beta }_{ki}\langle \mathbf{x },e_{\beta }\rangle \right) +H^{\beta }h^{\beta }_{il}h^{\alpha }_{lj}\nonumber \\&-h^{\beta }_{ik}h^{\beta }_{kl}h^{\alpha }_{lj}\ +2h^{\beta }_{ik}h^{\alpha }_{kl}h^{\beta }_{lj} -h^{\beta }_{ij}h^{\beta }_{kl}h^{\alpha }_{kl}-h^{\alpha }_{ik} h^{\beta }_{kl}h^{\beta }_{lj}\nonumber \\&+h^{\beta }_{ij}h^{\beta }_{kl}h^{\alpha }_{kl} -\frac{1}{2}\langle \mathbf{x },\nabla h^{\alpha }_{ij}\rangle +\frac{1}{2}h^{\alpha }_{ij}\nonumber \\&= h^{\alpha }_{ij}+2h^{\beta }_{ik}h^{\alpha }_{kl}h^{\beta }_{lj} -h^{\beta }_{ik}h^{\beta }_{kl}h^{\alpha }_{lj} -h^{\alpha }_{ik}h^{\beta }_{kl}h^{\beta }_{lj}. \end{aligned}$$
(9.11)

By Ricci equation,

$$\begin{aligned} 2h^{\beta }_{ik}h^{\alpha }_{kl}h^{\beta }_{lj}-h^{\beta }_{ik} h^{\beta }_{kl}h^{\alpha }_{lj}-h^{\alpha }_{ik}h^{\beta }_{kl}h^{\beta }_{lj}&= h^{\beta }_{ik}(h^{\alpha }_{kl}h^{\beta }_{lj}-h^{\beta }_{kl}h^{\alpha }_{lj}) +(h^{\beta }_{ik}h^{\alpha }_{kl}-h^{\alpha }_{ik}h^{\beta }_{kl}) h^{\beta }_{lj}\nonumber \\&= h^{\beta }_{ik}R_{\alpha \beta kj}+R_{\beta \alpha il}h^{\beta }_{lj}=0. \end{aligned}$$

The last equality follows from our assumption that the normal curvature is zero. Thus we obtain (9.9) from (9.11), This proves the lemma. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Arezzo, C., Sun, J. Self-shrinkers for the mean curvature flow in arbitrary codimension. Math. Z. 274, 993–1027 (2013). https://doi.org/10.1007/s00209-012-1104-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00209-012-1104-y

Keywords

Mathematics Subject Classification (2000)

Navigation