Skip to main content
Log in

Min–max theory for constant mean curvature hypersurfaces

  • Published:
Inventiones mathematicae Aims and scope

Abstract

In this paper, we develop a min–max theory for the construction of constant mean curvature (CMC) hypersurfaces of prescribed mean curvature in an arbitrary closed manifold. As a corollary, we prove the existence of a nontrivial, smooth, closed, almost embedded, CMC hypersurface of any given mean curvature c. Moreover, if c is nonzero then our min–max solution always has multiplicity one.

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

Notes

  1. This states that if a sequence of currents \(T_i\) converges to some T under the flat topology and the masses do not drop, \(\limsup _{i\rightarrow \infty } \mathbf {M}(T_i)\le \mathbf {M}(T)\), then the sequence converges under the stronger \(\mathbf {F}\)-metric, \(\lim _{i\rightarrow \infty }\mathbf {F}(T_i, T)=0\). Moreover, this convergence is uniform in T when T ranges in a compact subset under the \(\mathbf {F}\)-metric topology.

References

  1. Agol, I., Marques, F.C., Neves, A.: Min–max theory and the energy of links. J. Am. Math. Soc. 29(2), 561–578 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  2. Almgren Jr., F.J.: The Theory of Varifolds, Mimeographed Notes. Princeton University, Princeton (1965)

    Google Scholar 

  3. Almgren Jr., F.J.: Existence and regularity almost everywhere of solutions to elliptic variational problems with constraints. Mem. Am. Math. Soc. 4(165), viii+199 (1976)

    MathSciNet  MATH  Google Scholar 

  4. Arnold, V.I.: Arnold’s problems. Springer-Verlag, Berlin; PHASIS, Moscow. Translated and revised edition of the 2000 Russian original. Philippov, A. Yakivchik and M. Peters, With a preface by V (2004)

  5. Almgren, J., Justin, F.: The homotopy groups of the integral cycle groups. Topology 1, 257–299 (1962)

    Article  MathSciNet  MATH  Google Scholar 

  6. Barbosa, J.L., do Carmoand, M., Eschenburg, J.: Stability of hypersurfaces of constant mean curvature in Riemannian manifolds. Math. Z. 197(1), 123–138 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  7. Bellettini, C., Wickramasekera, N.: Stable CMC integral varifolds of codimension 1: regularity and compactness. arXiv preprint arXiv:1802.00377 (2018)

  8. Bérard, P., Meyer, D.: Inégalités isopérimétriques et applications. Ann. Sci. École Norm. Sup. (4) 15(3), 513–541 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  9. Chambers, G.R., Liokumovich, Y.: Existence of minimal hypersurfaces in complete manifolds of finite volume. arXiv:1609.04058 (2016)

  10. Chruściel, P.T., Galloway, G.J., Pollack, D.: Mathematical general relativity: a sampler. Bull. Am. Math. Soc. (N.S.) 47(4), 567–638 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  11. Colding, T.H., De Lellis, C.: The min–max construction of minimal surfaces. Surv Differ Geom 8:75–107

    Article  MathSciNet  MATH  Google Scholar 

  12. Colding, T.H., Minicozzi, W.P.: II. A Course in Minimal Surfaces Volume 121 of Graduate Studies in Mathematics. American Mathematical Society, Providence (2011)

    Google Scholar 

  13. De Lellis, C., Ramic, J.: Min–max theory for minimal hypersurfaces with boundary. Ann. Inst. Fourier (Grenoble) 68(5), 1909–1986 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  14. De Lellis, C., Tasnady, D.: The existence of embedded minimal hypersurfaces. J. Differ. Geom. 95(3), 355–388 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  15. Duzaar, F., Steffen, K.: Existence of hypersurfaces with prescribed mean curvature in Riemannian manifolds. Indiana Univ. Math. J. 45(4), 1045–1093 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  16. Ginzburg, V.L.: On closed trajectories of a charge in a magnetic field. An application of symplectic geometry. In: Contact and symplectic geometry (Cambridge 1994). Publications of the Newton Institute, vol. 8, pp. 131–148. Cambridge University Press, Cambridge, UK (1996)

  17. Giusti, E.: Minimal Surfaces and Functions of Bounded Variation. Monographs in Mathematics, vol. 80. Birkhäuser Verlag, Basel (1984)

    Book  MATH  Google Scholar 

  18. Guaraco, M.A.M.: Min–max for phase transitions and the existence of embedded minimal hypersurfaces. J. Differ. Geom. 108(1), 91–133 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  19. Harvey, R., Lawson, B.: Extending minimal varieties. Invent. Math. 28, 209–226 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  20. Heinz, E.: über die Existenz einer Fläche konstanter mittlerer Krümmung bei vorgegebener Berandung. Math. Ann. 127, 258–287 (1954)

    Article  MathSciNet  MATH  Google Scholar 

  21. Hildebrandt, S.: On the Plateau problem for surfaces of constant mean curvature. Commun. Pure Appl. Math. 23, 97–114 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  22. Hoffman, D., Meeks, W.H.: III. The strong halfspace theorem for minimal surfaces. Invent. Math. 101(2), 373–377 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  23. Huisken, G., Yau, S.-T.: Definition of center of mass for isolated physical systems and unique foliations by stable spheres with constant mean curvature. Invent. Math. 124(1–3), 281–311 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  24. Kapouleas, N.: Complete constant mean curvature surfaces in Euclidean three-space. Ann. Math. (2) 131(2), 239–330 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  25. Ketover, D.: Equivariant min–max theory. arXiv:1612.08692 (2016)

  26. Ketover, D., Zhou, X.: Entropy of closed surfaces and min–max theory. J. Differ. Geom. 110(1), 31–71 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  27. Krantz, S.G., Parks, H.R.: A Primer of Real Analytic Functions, Volume 4 of Basler Lehrbücher [Basel Textbooks]. Birkhäuser Verlag, Basel (1992)

    Book  Google Scholar 

  28. Li, M., Zhou, X.: Min–max theory for free boundary minimal hypersurfaces I-regularity theory. J. Differ. Geom. arXiv:1611.02612 (2016)

  29. Liokumovich, Y., Marques, F.C., Neves, A.: Weyl law for the volume spectrum. Ann. Math. (2) 187(3), 933–961 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  30. Lobaton, E.J., Salamon, T.R.: Computation of constant mean curvature surfaces: application to the gas-liquid interface of a pressurized fluid on a superhydrophobic surface. J. Colloid Interface Sci. 314, 184–198 (2007)

    Article  Google Scholar 

  31. López, R.: Wetting phenomena and constant mean curvature surfaces with boundary. Rev. Math. Phys. 17(7), 769–792 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  32. Mahmoudi, F., Mazzeo, R., Pacard, F.: Constant mean curvature hypersurfaces condensing on a submanifold. Geom. Funct. Anal. 16(4), 924–958 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  33. Marques, F.C., Neves, A.: Min-max theory and the Willmore conjecture. Ann. Math. (2) 179(2), 683–782 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  34. Marques, F.C., Neves, A.: Morse index and multiplicity of min–max minimal hypersurfaces. Camb. J. Math. 4(4), 463–511 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  35. Marques, F.C., Neves, A.: Existence of infinitely many minimal hypersurfaces in positive ricci curvature. Invent. Math. (2017). https://doi.org/10.1007/s00222-017-0716-6

    Article  MathSciNet  MATH  Google Scholar 

  36. Meeks III, W., Simon, L., Yau, S.T.: Embedded minimal surfaces, exotic spheres, and manifolds with positive Ricci curvature. Ann. Math. (2) 116(3), 621–659 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  37. Meeks III, W.H., Mira, P., Perez, J., Ros, A.: Constant mean curvature spheres in homogeneous three-spheres. arXiv preprint arXiv:1308.2612 (2013)

  38. Meeks III, W.H., Mira, P., Perez, J., Ros, A.: Constant mean curvature spheres in homogeneous three-manifolds. arXiv preprint arXiv:1706.09394 (2017)

  39. Montezuma, R.: Min–max minimal hypersurfaces in non-compact manifolds. J. Differ. Geom. 103(3), 475–519 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  40. Morgan, F.: Regularity of isoperimetric hypersurfaces in Riemannian manifolds. Trans. Am. Math. Soc. 355(12), 5041–5052 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  41. Nardulli, S.: The isoperimetric profile of a smooth Riemannian manifold for small volumes. Ann. Glob. Anal. Geom. 36(2), 111–132 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  42. Novikov, S.P.: The Hamiltonian formalism and a multivalued analogue of Morse theory. Uspekhi Mat. Nauk 37(5(227)), 3–49 (1982)

    MathSciNet  Google Scholar 

  43. Pacard, F.: Constant mean curvature hypersurfaces in Riemannian manifolds. Riv. Mat. Univ. Parma (7) 4, 141–162 (2005)

    MathSciNet  MATH  Google Scholar 

  44. Pitts, J.T.: Existence and Regularity of Minimal Surfaces on Riemannian Manifolds. Volume 27 of Mathematical Notes. Princeton University Press, Princeton (1981)

    Google Scholar 

  45. Qing, J., Tian, G.: On the uniqueness of the foliation of spheres of constant mean curvature in asymptotically flat 3-manifolds. J. Am. Math. Soc. 20(4), 1091–1110 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  46. Ros, A.: The isoperimetric problem. Global Theory of Minimal Surfaces. In: Clay Mathematics Proceedings, vol. 2, pp. 175–209. American Mathematical Society, Providence, RI (2005)

  47. Rosenberg, H., Schneider, M.: Embedded constant-curvature curves on convex surfaces. Pacific J. Math. 253(1), 213–218 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  48. Rosenberg, H., Smith, G.: Degree theory of immersed hypersurfaces. arXiv:1010.1879v3 (2016)

  49. Schneider, M.: Closed magnetic geodesics on \(S^2\). J. Differ. Geom. 87(2), 343–388 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  50. Schoen, R., Simon, L.: Regularity of stable minimal hypersurfaces. Commun. Pure Appl. Math. 34(6), 741–797 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  51. Schoen, R., Simon, L., Yau, S.-T.: Curvature estimates for minimal hypersurfaces. Acta Math. 134(3–4), 275–288 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  52. Simon, L.: Lectures on geometric measure theory. In: Proceedings of the Centre for Mathematical Analysis, vol. 3. Centre for Mathematical Analysis, Australian National University, Canberra (1983)

  53. Smith, F.R: On the existence of embedded minimal 2-spheres in the 3-sphere, endowed with an arbitrary Riemannian metric. Ph.D. thesis, Ph.D. thesis, Supervisor: Leon Simon, University of Melbourne (1982)

  54. Song, A.: Local min–max surfaces and strongly irreducible minimal Heegaard splittings. arXiv:1706.01037 (2017)

  55. Struwe, M.: Large \(H\)-surfaces via the mountain-pass-lemma. Math. Ann. 270(3), 441–459 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  56. Struwe, M.: The existence of surfaces of constant mean curvature with free boundaries. Acta Math. 160(1–2), 19–64 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  57. Tamanini, I.: Boundaries of Caccioppoli sets with Hölder-continuous normal vector. J. Reine Angew. Math. 334, 27–39 (1982)

    MathSciNet  MATH  Google Scholar 

  58. Wente, H.C.: Counterexample to a conjecture of H. Hopf. Pacific J. Math. 121(1), 193–243 (1986)

    Article  MATH  Google Scholar 

  59. White, B.: The maximum principle for minimal varieties of arbitrary codimension. Commun. Anal. Geom. 18(3), 421–432 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  60. Wickramasekera, N.: A general regularity theory for stable codimension 1 integral varifolds. Ann. Math. (2) 179(3), 843–1007 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  61. Ye, R.: Foliation by constant mean curvature spheres. Pacific J. Math. 147(2), 381–396 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  62. Zhou, X.: Min-max hypersurface in manifold of positive Ricci curvature. J. Differ. Geom. 105(2), 291–343 (2017)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Both authors are grateful to Prof. Shing-Tung Yau for suggesting this problem and for his generous support. X. Zhou would also like to thank Prof. Richard Schoen and Prof. Neshan Wickramasekera for valuable comments. J. Zhu would also like to thank Prof. William Minicozzi for his invaluable guidance and encouragement. X. Zhou is partially supported by NSF grant DMS-1704393 and DMS-1811293. J. Zhu is partially supported by NSF grant DMS-1607871. Finally, both authors would like to thank the anonymous referees for their comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xin Zhou.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: an interpolation lemma

The first lemma below was essentially due to Pitts [44, Lemma 3.8], but the modification to find the interpolation sequence using boundaries of Caccioppoli sets was completed by the first author [62, Proposition 5.3].

Lemma A.1

Suppose \(L>0\), \(\eta >0\), W is a compact subset of U, and \(\Omega \in {\mathcal {C}}(M)\). Then there exists \(\delta =\delta (L, \eta , U, W, \Omega )>0\), such that for any \(\Omega _1, \Omega _2\in {\mathcal {C}}(M)\) satisfying

(a):

\({\text {spt}}(\Omega _i-\Omega )\subset W\), \(i=1, 2\),

(b):

\(\mathbf {M}(\partial \Omega _i)\le L\), \(i=1, 2\),

(c):

\({\mathcal {F}}(\partial \Omega _1-\partial \Omega _2)\le \delta \),

there exist a sequence \(\Omega _1=\Lambda _0, \Lambda _1, \ldots , \Lambda _m=\Omega _2\in {\mathcal {C}}(M)\) such that for each \(j=0, \ldots , m-1\),

  1. (i)

    \({\text {spt}}(\Lambda _j-\Omega )\subset U\),

  2. (ii)

    \({\mathcal {A}}^c(\Lambda _j)\le \max \{{\mathcal {A}}^c(\Omega _1), {\mathcal {A}}^c(\Omega _2)\}+\eta \),

  3. (iii)

    \(\mathbf {M}(\partial \Lambda _j-\partial \Lambda _{j+1})\le \eta \).

  4. (iv)
    $$\begin{aligned} \mathbf {M}(\partial \Lambda _j)\le \max \{\mathbf {M}(\partial \Omega _1), \mathbf {M}(\partial \Omega _2)\}+\frac{\eta }{2}. \end{aligned}$$
  5. (v)
    $$\begin{aligned} \mathbf {M}(\Lambda _j-\Omega _i)\le \frac{\eta }{2c}, \text { for } i =1, 2. \end{aligned}$$

Proof

Note that by a covering argument, one only needs to prove the case when \(\partial \Omega _2\) is fixed, and in this case \(\delta =\delta (L, \eta , U, W, \Omega , \Omega _2)\).

Under our assumptions on \(\Omega _1, \Omega _2\), there are two issues for us that differ from the conclusions of Pitts [44, Lemma 3.8]:

  1. (1)

    we require the interpolating sequence \(\{\partial \Lambda _j\}\) to consist of boundaries of Caccioppoli sets, while in [44] the interpolating sequence consists of integral currents \(\{T_j\in {\mathcal {Z}}_n(M)\}\) (using notations therein);

  2. (2)

    for point (ii), we require that \({\mathcal {A}}^c(\Lambda _j)\) does not increase much from \({\mathcal {A}}^c(\Omega _1), {\mathcal {A}}^c(\Omega _2)\), while in [44] it was only proven that \(\mathbf {M}(T_j)\le \max \{\mathbf {M}(\partial \Omega _1), \mathbf {M}(\partial \Omega _2)\}+\eta \).

These two minor points can be easily deduced from the mentioned work of Zhou [62], and now we point out the necessary details. In fact, with regards to point (1), the proof of [62, Proposition 5.3] already proceeds using boundaries, and we will first justify properties (i, iii, iv).

Specifically, given \(L, \eta , \Omega , \Omega _2\) satisfying the assumptions in the Lemma, [62, Proposition 5.3] (applied to the case when \(m=1, l=0\) therein) gives the desired \(\delta >0\), such that if \(\Omega _1\) satisfies the assumptions (a–c), then there exists a sequence \(\Omega _1=\Lambda _0, \Lambda _1, \ldots , \Lambda _m=\Omega _2\in {\mathcal {C}}(M)\) such that for each \(j=0, \ldots , m-1\), properties (iii) and (iv) are satisfied.

Although in [62, Proposition 5.3] \(\Omega _i\) were not assumed to satisfy the additional condition \({\text {spt}}(\Omega _i-\Omega )\subset W\). It can be seen that the construction in [62, Proposition 5.3] indeed satisfies the required property (i) under this additional condition.

Note that properties (iv, v) immediately imply (ii), so it remains only to show (v). Indeed, from the construction in [62, Proposition 5.3], one knows that the symmetric difference \(\Lambda _j\triangle \Omega _i\) is a subset of the union of the symmetric difference \(\Omega _1\triangle \Omega _2\) together with finitely many (depending only on L and \(\eta \)) balls of arbitrarily small radii in U. Since \({\text {Vol}}(\Omega _1\triangle \Omega _2)=\mathbf {M}(\Omega _1-\Omega _2)={\mathcal {F}}(\partial \Omega _1, \partial \Omega _2)\) by the isoperimetric lemma [62, Lemma 7.3], one thus obtains (v) by taking \(\delta \) small enough. \(\square \)

Next we explain how to use Lemma A.1 to finish the proof of \((c)\Longrightarrow (d)\) in Proposition 5.3.

Proof of Proposition 5.3\((c)\Longrightarrow (d)\)continued It suffices to prove that for any given \(\epsilon , \delta \) and \(\Omega \in {\mathscr {A}}^c_n(U; \epsilon , \delta ; \mathbf {M})\), there exists \(\delta '=\delta '(\epsilon , \delta , U, W, \Omega )>0\), such that \(\Omega \in {\mathscr {A}}^c_n(W; \epsilon , \delta '; {\mathcal {F}})\). Indeed, suppose so, then by (c), we can find a sequence \(\Omega _i\in {\mathscr {A}}^c_n(U; \epsilon _i, \delta _i; \mathbf {M})\) with \(\lim _{i\rightarrow \infty }|\partial \Omega _i|=V\) as varifods. As \(\Omega _i\in {\mathscr {A}}^c_n(W; \epsilon _i, \delta _i'; {\mathcal {F}})\) for some \(\delta _i'\), this immediately implies that V is c-almost minimizing in W.

Given some \(\delta '>0\), to show that \(\Omega \in {\mathscr {A}}^c_n(W; \epsilon , \delta '; {\mathcal {F}})\), consider an arbitrary sequence \(\Omega =\Omega _0, \Omega _1, \Omega _2, \ldots , \Omega _m\in {\mathcal {C}}(M)\) satisfying that:

  1. (i)

    \({\text {spt}}(\Omega _i-\Omega )\subset W\);

  2. (ii)

    \({\mathcal {F}}(\partial \Omega _{i+1}, \partial \Omega _i)\le \delta '\);

  3. (iii)

    \({\mathcal {A}}^c(\Omega _i)\le {\mathcal {A}}^c(\Omega )+\delta '\), for \(i=1, \ldots , m\).

Note that the mass \(\mathbf {M}(\partial \Omega _i)\le \mathbf {M}(\partial \Omega )+c{\text {Vol}}(M)+1\); in Lemma A.1 we let \(L=\mathbf {M}(\partial \Omega )+c{\text {Vol}}(M)+1\), \(\eta =\delta /2\), and hence there exists \(\delta '=\delta (L, \eta , U, W, \Omega )\), and for each i there exist a sequence \(\Omega _i=\Omega _{i, 1}, \ldots , \Omega _{i, l_i}=\Omega _{i+1}\), such that

  1. (i)

    \({\text {spt}}(\Omega _{i, l}-\Omega )\subset U\),

  2. (ii)

    \({\mathcal {A}}^c(\Omega _{i, l})\le \max \{{\mathcal {A}}^c(\Omega _i), {\mathcal {A}}^c(\Omega _{i+1})\}+\delta /2\le {\mathcal {A}}^c(\Omega )+\delta /2+\delta '\),

  3. (iii)

    \(\mathbf {M}(\partial \Omega _{i, l}-\partial \Omega _{i, l+1})\le \delta /2\).

Now consider the new sequence by putting all these interpolations together: \(\Omega _0=\Omega , \Omega _{0, 1}, \ldots , \Omega _{0, l_0}=\Omega _1, \Omega _{1, 1}, \ldots , \Omega _{1, l_1}=\Omega _2, \ldots , \Omega _3, \ldots , \cdots , \Omega _m\). It is evident that this new sequence satisfies assumptions (i), (ii), (iii) in Definition 5.1 for \({\mathscr {A}}^c_n(U; \epsilon , \delta ; \mathbf {M})\) if we further shrink \(\delta '\) to make \(\delta '\le \delta /2\), and hence we get \({\mathcal {A}}^c(\Omega _m)\ge {\mathcal {A}}^c(\Omega )-\epsilon \). This proves that \(\Omega \in {\mathscr {A}}^c_n(W; \epsilon , \delta '; {\mathcal {F}})\). \(\square \)

Appendix B: interpolation process

Proof of Claim 2 in Proposition 4.4

Here we describe the construction of \(\{\phi _i\}\) by interpolating \(\{\phi ^1_i\}\).

For fixed \(i\in {{\mathbb {N}}}\) and consider a 1-cell \(\alpha \in I(1, k_i)\). We only need to show how to interpolate \(\phi ^1_i\) when restricted to \(\alpha _0\). For notational simplicity we write \(\alpha =[0, 1]\). For \(x\in \alpha \) let \(\tilde{X}_i(x)\) be the linear interpolation between \(\tilde{X}_i(0)=\tilde{X}(|\partial \phi ^*_i(0)|)\) and \(\tilde{X}_i(1)=\tilde{X}(|\partial \phi ^*_i(1)|)\), where \(\tilde{X}\) is defined in (4.6)). That is, \(\tilde{X}_i(x)=(1-x)\tilde{X}_i(0)+x\tilde{X}_i(1)\).

The continuity of the map \(V\rightarrow \tilde{X}(V)\) implies that \(\Vert \tilde{X}_i(x)-\tilde{X}_i(0)\Vert _{C^1(M)}\rightarrow 0\) uniformly as \(i\rightarrow \infty \), (since under the varifold \(\mathbf {F}\)-metric, \(\mathbf {F}(|\partial \phi ^*_i(0)|, |\partial \phi ^*_i(1)|)\le \mathbf {f}(\phi ^*_i)\rightarrow 0\) uniformly in i). Define \({\bar{Q}}_i(x)\) to be the push-forward of \(\phi ^*_i(0)\) by the flow of \(\tilde{X}_i(x)\) up to time 1; this gives a map \({\bar{Q}}_i: \alpha \rightarrow \mathcal {C}(M)\). Note that \(\partial {\bar{Q}}_i: \alpha \rightarrow {\mathcal {Z}}_n(M)\) is continuous under the \(\mathbf {F}\)-metric.

Since \({\bar{Q}}_i(x)\) and \(\phi ^1_i(0)\) are the push-forwards of the same initial set \(\phi ^*_i(0)\) under the flows of \(\tilde{X}_i(x)\) and \(\tilde{X}_i(0)\) respectively, we have

$$\begin{aligned} \left. \begin{array}{cl} \mathbf {F}(\partial {\bar{Q}}_i(x), \partial \phi ^1_i(0))\rightarrow 0 \\ \mathbf {M}({\bar{Q}}_i(x)- \phi ^1_i(0))\rightarrow 0 \end{array} \right. , \text { uniformly in }x,\alpha \hbox { as }i\rightarrow \infty . \end{aligned}$$
(B.1)

Note that the uniformity follows from a simple contradiction argument, using the uniform smallness between \(\tilde{X}_i(x)\) and \(\tilde{X}_i(0)\) together with the fact that the set of images of all the \(\{\phi ^*_i\}\) is compact.

As \({\bar{Q}}_i(1)\) and \(\phi ^1_i(1)\) are the respective push-forwards of \(\phi ^*_i(0)\) and \(\phi ^*_i(1)\) under the same flow of \(\tilde{X}_i(1)\) (and since these flows have uniformly bounded Lipschitz constant, independent of i), we have

$$\begin{aligned}&\mathbf {M}(\partial {\bar{Q}}_i(1)-\partial \phi ^1_i(1)) \le C\, \mathbf {M}(\partial \phi ^*_i(1)-\partial \phi ^*_i(0))\nonumber \\&\quad \rightarrow 0, \text { uniformly in }\alpha \text { as }i\rightarrow \infty . \end{aligned}$$
(B.2)

Now we can apply the discretization-interpolation theorem [62, Theorem 5.1] (c.f. [33, Theorem 13.1]) to \({\bar{Q}}_i\). In fact, given any continuous map \({\bar{\phi }}\) from a 1-cell \(\alpha \) to \({\mathcal {C}}(M)\), the discretization-interpolation theorem implies that for any \(\eta >0\), there exists a large integer \(l_\eta >0\), and a discrete map \(\phi \) defined on the vertices of some refinement \(\alpha (l_\eta )\), so that: (i) the masses of \(\phi \) do not go up much, (ii) the fineness of \(\phi \) is controlled by \(\eta \), and (iii) \(\phi \) stays close to \({\bar{\phi }}\) under the flat norm. In particular, the triple \((\eta , l_\eta , \phi )\) can be chosen to be any triple \((\delta _j, k_j, \phi _j)\) given by [62, Theorem 5.1] (here the latter triple uses the notation of that theorem) satisfying \(\delta _j\le \eta \).

Applying this to \({\bar{Q}}_i\), for any \(\eta >0\), we obtain a large integer \(l_\eta >0\) and a map \(Q_i: \alpha (l_\eta )_0\rightarrow {\mathcal {C}}(M)\), such that

  1. (i)

    given \(x\in \alpha (l_\eta )_0\),

    $$\begin{aligned} \mathbf {M}(\partial Q_i(x)) \le \mathbf {M}(\partial {\bar{Q}}_i(x))+\eta /2, \end{aligned}$$

    and also by the same argument as in the proof of point (v) of Lemma A.1,

    $$\begin{aligned} \mathbf {M}(Q_i(x)-{\bar{Q}}_i(x)) \le \eta /(2c), \end{aligned}$$

    and hence

    $$\begin{aligned} {\mathcal {A}}^c(Q_i(x))\le {\mathcal {A}}^c({\bar{Q}}_i(x))+\eta ; \end{aligned}$$
  2. (ii)

    \(\mathbf {f}(Q_i)\le \eta \);

  3. (iii)

    \(\sup \{{\mathcal {F}}(\partial Q_i(x)-\partial {\bar{Q}}_i(x)): x\in \alpha (l_\eta )_0\}<\eta \).

When \(\eta \rightarrow 0\), by (i, iii) and [44, 2.1(20)]Footnote 1 (see also [33, Lemma 4.1]), we have

$$\begin{aligned} \lim _{\eta \rightarrow 0}\sup \{\mathbf {F}(\partial Q_i(x), \partial {\bar{Q}}_i(x)): x\in \alpha (l_\eta )_0\}=0. \end{aligned}$$

To finish the interpolation, take a sequence \(\eta _i\rightarrow 0\), and denote \(l_i=k_i+l_{\eta _i}+1\), then we construct \(\phi _i: I(1, k_i+l_{\eta _i}+1)\rightarrow {\mathcal {C}}(M)\) by defining \(\phi _i\) on each \(\alpha (l_{\eta _i}+1)_0\) by

$$\begin{aligned} \phi _i(x)= \left\{ \begin{array}{cl} Q_i(3x) &{} \text { for }x\in [0, 1/3]\cap \alpha (l_{\eta _i}+1)_0 \\ \phi _i^1(1) &{} \text { otherwise.} \end{array} \right. \end{aligned}$$

The desired properties (a, b, c, d) of \(\phi _i\) can be read off from (B.1), (B.2) and the properties of \(Q_j\). Since \({\bar{Q}}_i\) is obtained from a continuous deformation from \(\phi ^*_i\), a further interpolation argument shows that S is homotopic to \(S^*\), and hence we finish the proof of Claim 2.

Appendix C: good replacement property and regularity

Here we record the notions of good replacements and the good replacement property. Recall the following definitions by Colding–De Lellis [11]. Consider two open subsets \(W\subset \subset U\subset M^{n+1}\).

Definition C.1

[11, Definition 6.1]. Let \(V \in V_n(U)\) be stationary in U. A stationary varifold \(V' \in V_n(U)\) is said to be a replacement for V in W if

Definition C.2

[11, Definition 6.2]. Let \(V \in V_n(U)\) be stationary in U. V is said to have the good replacement property in W if

  1. (a)

    there is a positive function \(r: W\rightarrow {{\mathbb {R}}}\) such that for every annulus \(A_{s, t}(x)\cap M\subset W\) with \(0<s<t<r(x)\), there is a replacement \(V'\) for V in \(A_{s, t}(x)\cap M\);

  2. (b)

    the replacement \(V'\) has a replacement \(V''\) in every annulus \(A_{s, t}(y)\cap M\subset W\) with \(0<s<t<r(y)\);

  3. (c)

    \(V''\) has a replacement \(V'''\) in every annulus \(A_{s, t}(z)\cap M\subset W\) with \(0<s<t<r(z)\).

Note that our formulations are local compared to those in [11]. Indeed, the proofs of [11, Proposition 6.3] and [14, Theorem 2.8] are purely local, so the following proposition still holds:

Proposition C.3

[11, Proposition 6.3], [14, Proposition 2.8] When \(2\le n\le 6\), if \(V\in V_n(U)\) has the good replacement property in W, then is an integer multiple of some smooth embedded minimal hypersurface \(\Sigma \).

Appendix D: Almgren–Pitts deformation process

Here we sketch the deformation process of Almgren–Pitts [44, 4.10] that we used in the proof of Theorem 5.6. Recall that we assumed for the sake of contradiction that there is no \(V\in C(S)\) that is c-almost minimizing on small annuli.

As the critical set C(S) (Definition 3.6) is compact under the varifold \(\mathbf {F}\)-norm, there exists some uniform \(\epsilon >0\), such that for each \(V\in C(S)\), there exist \(p_V\in M\) and some \(\tilde{r}>0\), so that given any \(\delta >0\) and \(\tilde{r}>r+2s>r-2s>0\), and \(\Omega \in {\mathcal {C}}(M)\) with \(\mathbf {F}(|\partial \Omega |, V)<\epsilon \), we must have \(\Omega \notin {\mathscr {A}}^c_n(A_{r-2s, r+2s}(p_V)\cap M; \epsilon , \delta ; \mathbf {M})\). Moreover, we can pick finitely many \(\{V_1, \ldots , V_m\}\subset C(S)\), such that the \(\mathbf {F}\)-balls \(\{B^{\mathbf {F}}_{\epsilon /4}(V_j)\}\) form a cover of C(S). Given j, denote \(p_j=p_{V_j}\).

Now consider \(\phi _i\in S\) for i sufficiently large. Denote by \(I_i\) the subset of \(\text {domain}(\phi _i)\), such that if \(x\in I_i\), then \({\mathcal {A}}^c(\phi _i(x))\) is close enough to \(\mathbf {L}^c\). Then for each \(x\in I_i\), \(\phi _i(x)\) must be close to some \(V_j\) under the \(\mathbf {F}\)-norm, and so for any \(\delta >0\),

$$\begin{aligned} \phi _i(x)\notin {\mathscr {A}}^c_n(A_{r-2s, r+2s}(p_j)\cap M; \epsilon , \delta ; \mathbf {M}), \end{aligned}$$

where \(r, s>0\) can be as small as we want. By definition of \({\mathscr {A}}^c_n\) there exists a discrete deformation sequence, supported in \(A_{r-2s, r+2s}(p_j)\cap M\) and starting from \(\phi _i(x)\), which is almost continuous at the scale of \(\delta \) under the \(\mathbf {M}\)-norm and eventually deforms \({\mathcal {A}}^c(\phi _i(x))\) down by at least \(\epsilon \).

We can then choose \(\delta \) to be the fineness of \(\phi _i\), and use the corresponding deformation sequence to deform \(\phi _i\). The choice of \(\delta \) will make this deformation a homotopy. Note that to ensure \(\phi _i\) is a sweepout, we not only need to deform \(\phi _i(x)\) but also its adjacent slices. The strategy is then to allow two deformations, taking place in disjoint annuli, for each \(x\in I_i\). A covering argument then shows that any slice is indeed affected by at most two such deformations. Finally, the deformations are chosen so that at least in one small annulus, the \({\mathcal {A}}^c\)-value of \(\phi _i(\cdot )\) is deformed down by \(\epsilon \), while the \({\mathcal {A}}^c\)-value of \(\phi _i(\cdot )\) in the other deformation annulus is increased no more than \(\delta \). Choosing \(\delta <\epsilon /2\), the \({\mathcal {A}}^c\)-values of \(\phi _i(\cdot )\) will therefore be deformed down by \(\epsilon /2\) everywhere in \(I_i\), and this implies \(\mathbf {L}^c(\tilde{S})<\mathbf {L}^c(S)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, X., Zhu, J.J. Min–max theory for constant mean curvature hypersurfaces. Invent. math. 218, 441–490 (2019). https://doi.org/10.1007/s00222-019-00886-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-019-00886-1

Navigation