Skip to main content
Log in

The Viscosity Method for Min–Max Free Boundary Minimal Surfaces

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

We adapt the viscosity method introduced by Rivière (Publ Math Inst Hautes Études Sci 126:177–246, 2017. https://doi.org/10.1007/s10240-017-0094-z) to the free boundary case. Namely, given a compact oriented surface \(\Sigma \), possibly with boundary, a closed ambient Riemannian manifold \(({\mathcal {M}}^m,g)\) and a closed embedded submanifold \({\mathcal {N}}^n\subset {\mathcal {M}}\), we study the asymptotic behavior of (almost) critical maps \(\Phi \) for the functional

$$\begin{aligned} E_\sigma (\Phi ):={\text {area}}(\Phi )+\sigma {\text {length}}(\Phi |_{\partial \Sigma })+\sigma ^4\int _\Sigma |\mathrm {I\!I}^\Phi |^4\,{\text {vol}}_\Phi \end{aligned}$$

on immersions \(\Phi :\Sigma \rightarrow {\mathcal {M}}\) with the constraint \(\Phi (\partial \Sigma )\subseteq {\mathcal {N}}\), as \(\sigma \rightarrow 0\), assuming an upper bound for the area and a suitable entropy condition. As a consequence, given any collection \({\mathcal {F}}\) of compact subsets of the space of smooth immersions \((\Sigma ,\partial \Sigma )\rightarrow ({\mathcal {M}},{\mathcal {N}})\), assuming \({\mathcal {F}}\) to be stable under isotopies of this space, we show that the min–max value

$$\begin{aligned} \inf _{A\in {\mathcal {F}}}\max _{\Phi \in A}{\text {area}}(\Phi ) \end{aligned}$$

is the sum of the areas of finitely many branched minimal immersions \(\Phi _{(i)}:\Sigma _{(i)}\rightarrow {\mathcal {M}}\) with \(\Phi _{(i)}(\partial \Sigma _{(i)})\subseteq {\mathcal {N}}\) and \(\partial _\nu \Phi _{(i)}\perp T{\mathcal {N}}\) along \(\partial \Sigma _{(i)}\), whose (connected) domains \(\Sigma _{(i)}\) can be different from \(\Sigma \) but cannot have a more complicated topology. Contrary to other min–max frameworks, the present one applies in an arbitrary codimension. We adopt a point of view which exploits extensively the diffeomorphism invariance of \(E_\sigma \) and, along the way, we simplify and streamline several arguments from the initial work (Rivière 2017). Some parts generalize to closed higher-dimensional domains, for which we get an integral stationary varifold in the limit.

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 is an easy consequence of the Riemannian uniformization theorem, applied to \((\Sigma ,g_k)\) if \(\partial \Sigma =\emptyset \), or to the doubled surface obtained by gluing two copies of \(\Sigma \) along \(\partial \Sigma \).

  2. The smoothness of f can be assumed by standard Schauder theory, since f satisfies an elliptic equation on a small ball.

References

  1. Allard, W.K.: On the first variation of a varifold. Ann. Math. (2) 95, 417–491, 1972. https://doi.org/10.2307/1970868

  2. Allard, W.K.: An integrality theorem and a regularity theorem for surfaces whose first variation with respect to a parametric elliptic integrand is controlled. Proc. Symp. Pure Math. 44, 1–28, 1986. https://doi.org/10.1090/pspum/044/840267

  3. Almgren, F.J.: The Theory of Varifolds. Mimeographed Notes. Princeton University Press, Princeton, 1965

  4. Ambrosetti, A., Malchiodi, A.: Nonlinear Analysis and Semilinear Elliptic Problems, Cambridge Studies in Advanced Mathematics, Vol. 104. Cambridge University Press, Cambridge, 2007. https://doi.org/10.1017/CBO9780511618260

  5. Bellettini, C.: Multiplicity-1 minmax minimal hypersurfaces in manifolds with positive Ricci curvature. arXiv preprint arXiv:2004.10112, 2020

  6. Bernard, Y., Rivière, T.: Uniform regularity results for critical and subcritical surface energies. Calc. Var. Partial Differ. Equ. 58(1), art. 10, 2019. https://doi.org/10.1007/s00526-018-1457-0

  7. Carlotto, A., Franz, G., Schulz, M.B.: Free boundary minimal surfaces with connected boundary and arbitrary genus. arXiv preprint arXiv:2001.04920, 2020

  8. Cheng, D.R.: Asymptotics for the Ginzburg–Landau equation on manifolds with boundary under homogeneous Neumann condition. J. Funct. Anal. 278(4), 108364+93, 2020. https://doi.org/10.1016/j.jfa.2019.108364

  9. Chodosh, O., Mantoulidis, C.: Minimal surfaces and the Allen–Cahn equation on 3-manifolds: index, multiplicity, and curvature estimates. Ann. Math. (2) 191(1), 213–328, 2020. https://doi.org/10.4007/annals.2020.191.1.4

  10. De Lellis, C., Ramic, J.: Min–max theory for minimal hypersurfaces with boundary. Ann. Inst. Fourier (Grenoble) 68(5), 1909–1986, 2018. https://doi.org/10.5802/aif.3200

  11. Douglas, J.: Solution of the problem of Plateau. Trans. Am. Math. Soc. 33(1), 263–321, 1931. https://doi.org/10.2307/1989472

  12. Evans, L.C., Gariepy, R.F.: Measure Theory and Fine Properties of Functions (Revised Edition), Textbooks in Mathematics. CRC Press, Boca Raton, 2015. https://doi.org/10.1201/b18333

  13. Federer, H., Fleming, W.H.: Normal and integral currents. Ann. Math. (2) 72, 458–520, 1960. https://doi.org/10.2307/1970227

  14. Folha, A., Pacard, F., Zolotareva, T.: Free boundary minimal surfaces in the unit 3-ball. Manuscr. Math. 154(3–4), 359–409, 2017. https://doi.org/10.1007/s00229-017-0924-9

  15. Fraser, A., Schoen, R.: Sharp eigenvalue bounds and minimal surfaces in the ball. Invent. Math. 203(3), 823–890, 2016. https://doi.org/10.1007/s00222-015-0604-x

  16. Ghoussoub, N.: Duality and Perturbation Methods in Critical Point Theory, Cambridge Tracts in Mathematics, Vol. 107. Cambridge University Press, Cambridge, 1993. https://doi.org/10.1017/CBO9780511551703

  17. Giaquinta, M.: Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems, Annals of Mathematics Studies, Vol. 105. Princeton University Press, Princeton, 1983. https://doi.org/10.2307/j.ctt1b9s07q

  18. Guang, Q., Li, M., Wang, Z., Zhou, X.: Min–max theory for free boundary minimal hypersurfaces II—general Morse index bounds and applications. arXiv preprint arXiv:1907.12064, 2019

  19. Guaraco, M.A.M.: Min–max for phase transitions and the existence of embedded minimal hypersurfaces. J. Differ. Geom. 108(1), 91–133, 2018. https://doi.org/10.4310/jdg/1513998031

  20. Hummel, C.: Gromov’s Compactness Theorem for Pseudo-holomorphic Curves, Progress in Mathematics, Vol. 151. Birkhäuser, Basel, 1997. https://doi.org/10.1007/978-3-0348-8952-0

  21. Hutchinson, J.E., Tonegawa, Y.: Convergence of phase interfaces in the van der Waals–Cahn–Hilliard theory. Calc. Var. Partial Differ. Equ. 10(1), 49–84, 2000. https://doi.org/10.1007/PL00013453

  22. Imayoshi, Y., Taniguchi, M.: An Introduction to Teichmüller Spaces. Springer, Tokyo, 1992. https://doi.org/10.1007/978-4-431-68174-8

  23. Irie, K., Marques, F.C., Neves, A.: Density of minimal hypersurfaces for generic metrics. Ann. Math. (2) 187(3), 963–972, 2018. https://doi.org/10.4007/annals.2018.187.3.8

  24. Jost, J.: On the existence of embedded minimal surfaces of higher genus with free boundaries in Riemannian manifolds. Variational Methods for Free Surface Interfaces (Eds. Concus P. and Finn R.) Springer, New York, 1987

  25. Kapouleas, N., Li, M.: Free boundary minimal surfaces in the unit three-ball via desingularization of the critical catenoid and the equatorial disk. arXiv preprint arXiv:1709.08556, 2017

  26. Kapouleas, N., Wiygul, D.: Free-boundary minimal surfaces with connected boundary in the 3-ball by tripling the equatorial disc. arXiv preprint arXiv:1711.00818, 2017

  27. Ketover, D.: Free boundary minimal surfaces of unbounded genus. arXiv preprint arXiv:1612.08691, 2016

  28. Kuwert, E., Lamm, T., Li, Y.: Two-dimensional curvature functionals with superquadratic growth. J. Eur. Math. Soc. (JEMS) 17(12), 3081–3111, 2015. https://doi.org/10.4171/JEMS/580

  29. Lehto, O., Virtanen, K.I.: Quasiconformal Mappings in the Plane, Die Grundlehren der mathematischen Wissenschaften, Vol. 126, 2nd edn. Springer, New York-Heidelberg-Berlin, 1973

  30. Li, M.: A general existence theorem for embedded minimal surfaces with free boundary. Commun. Pure Appl. Math. 68(2), 286–331, 2015. https://doi.org/10.1002/cpa.21513

  31. Li, M.: Free boundary minimal surfaces in the unit ball: recent advances and open questions. arXiv preprint arXiv:1907.05053, 2019

  32. Li, M., Zhou, X.: Min-max theory for free boundary minimal hypersurfaces I—regularity theory. arXiv preprint arXiv:1611.02612, 2016

  33. Liokumovich, Y., Marques, F.C., Neves, A.: Weyl law for the volume spectrum. Ann. Math. (2) 187(3), 933–961, 2018. https://doi.org/10.4007/annals.2018.187.3.7

  34. Marques, F.C., Neves, A.: Min–max theory and the Willmore conjecture. Ann. Math. (2) 179(2), 683–782, 2014. https://doi.org/10.4007/annals.2014.179.2.6

  35. Marques, F.C., Neves, A.: Morse index and multiplicity of min–max minimal hypersurfaces. Camb. J. Math. 4(4), 463–511, 2016. https://doi.org/10.4310/CJM.2016.v4.n4.a2

  36. Marques, F.C., Neves, A.: Existence of infinitely many minimal hypersurfaces in positive Ricci curvature. Invent. Math. 209(2), 577–616, 2017. https://doi.org/10.1007/s00222-017-0716-6

  37. Modica, L.: The gradient theory of phase transitions and the minimal interface criterion. Arch. Ration. Mech. Anal. 98(2), 123–142, 1987. https://doi.org/10.1007/BF00251230

  38. Pigati, A., Rivière, T.: The regularity of parameterized integer stationary varifolds in two dimensions. Commun. Pure Appl. Math. 73(9), 1981–2042, 2020. https://doi.org/10.1002/cpa.21927

  39. Pigati, A., Rivière, T.: A proof of the multiplicity one conjecture for min–max minimal surfaces in arbitrary codimension. Duke Math. J. 169(11), 2005–2044, 2020. https://doi.org/10.1215/00127094-2020-0002

  40. Pigati, A., Stern, D.: Minimal submanifolds from the abelian Higgs model. Invent. Math. 223(3), 1027–1095, 2021. https://doi.org/10.1007/s00222-020-01000-6

  41. Pitts, J.T.: Existence and Regularity of Minimal Surfaces on Riemannian manifolds, Mathematical Notes, Vol. 27. Princeton University Press, Princeton and University of Tokyo Press, Tokyo, 1981. https://doi.org/10.2307/j.ctt7zv66w

  42. Radó, T.: On Plateau’s problem. Ann. Math. (2) 31(3), 457–469, 1930. https://doi.org/10.2307/1968237

  43. Rivière, T.: Lecture 3. A Viscosity Approach to Minmax, Lecture 3 in Minmax Methods in the Calculus of Variations of Curves and Surfaces. Notes available online at https://people.math.ethz.ch/~riviere/minimax, 2016

  44. Rivière, T.: A viscosity method in the min–max theory of minimal surfaces. Publ. Math. Inst. Hautes Études Sci. 126, 177–246, 2017. https://doi.org/10.1007/s10240-017-0094-z

  45. Rivière, T.: The regularity of conformal target harmonic maps. Calc. Var. Partial Differ. Equ. 56(4), art. 117, 2017. https://doi.org/10.1007/s00526-017-1215-8

  46. Sacks, J., Uhlenbeck, K.: The existence of minimal immersions of 2-spheres. Ann. Math. (2) 113(1), 1–24, 1981. https://doi.org/10.2307/1971131

  47. Schoen, R., Simon, L.: Regularity of stable minimal hypersurfaces. Commun. Pure Appl. Math. 34(6), 741–797, 1981. https://doi.org/10.1002/cpa.3160340603

  48. Simon, L.: Lectures on Geometric Measure Theory, Proceedings of the Centre for Mathematical Analysis, Vol. 3. Australian National University, Canberra, 1984

  49. Song, A.: Existence of infinitely many minimal hypersurfaces in closed manifolds. arXiv preprint arXiv:1806.08816, 2018

  50. Stern, D.: Existence and limiting behavior of min–max solutions of the Ginzburg–Landau equations on compact manifolds. To appear on J. Differ. Geom. arXiv preprints arXiv:1612.00544, 2016, and arXiv:1704.00712, 2017

  51. Struwe, M.: On a free boundary problem for minimal surfaces. Invent. Math. 75(3), 547–560, 1984. https://doi.org/10.1007/BF01388643

  52. Wickramasekera, N.: A general regularity theory for stable codimension 1 integral varifolds. Ann. Math. (2) 179(3), 843–1007, 2014. https://doi.org/10.4007/annals.2014.179.3.2

  53. Yau, S.T.: Seminar on Differential Geometry, Annals of Mathematics Studies, vol. 102. Princeton University Press, Princeton (1982)

  54. Zhou, X.: On the multiplicity one conjecture in min-max theory. arXiv preprint arXiv:1901.01173, 2019

Download references

Acknowledgements

The author is grateful to the anonymous referees for their careful reading of an earlier version of the manuscript and for their helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alessandro Pigati.

Additional information

Communicated by A. Figalli

Publisher's Note

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

Appendix

Appendix

Proposition A.1

A continuous, \(W^{1,2}\) map \(u:B_1^2\rightarrow {\mathbb {R}}^m\) solving a linear system of the form

$$\begin{aligned} -\partial _i(g_{jk}\partial _i u^j)+b_{kpq}\partial _i u^p\partial _i u^q=0, \end{aligned}$$
(A.1)

with \(g\geqq \lambda >0\) symmetric and continuous and b bounded, is \(W^{1,r}_{loc}\) for all \(r<\infty \).

The same holds for u defined on the half-ball \(U':=B_1^2\cap \{\Im (z)\geqq 0\}\), if in addition we have

$$\begin{aligned} \partial _\nu u^k=0\text { for }k\leqq n,\quad u^k=0\text { for }k>n, \end{aligned}$$

as well as \(g_{ij}=0\) for \(i\leqq n\), \(j>n\), on the boundary \(\partial U'\), for some \(0\leqq n\leqq m\).

Remark A.2

The condition \(\partial _\nu u^k=0\) could be written more faithfully as \(g_{jk}\partial _\nu \Phi ^j=0\) and is of course meant in a weak sense, coupled with the equation: namely, we require \(\int _{U'}(g_{jk}\partial _i f\partial _i u^j+b_{kpq}f\partial _i u^p \partial _i u^q)=0\) for all \(f\in C^\infty _c(U')\) and \(k\leqq n\), allowing f to be nonzero on \(\partial U'\).

Proof

Assume u is a solution on the unit ball. Then, for any ball \(B_{2r}^2(x)\subseteq B_1^2\), we can integrate the equation against \(\eta ^2(u-(u)_{B_{2r}^2(x)})\), where \(\eta \in C^\infty _c(B_{2r}^2(x))\) is a cut-off function satisfying \(\eta =1\) on \(B_r^2(x)\) and \(|\mathrm{d}\eta |\leqq \frac{2}{r}\). Recall that the notation \((u)_S\) indicates the average of u on a set S. This gives

$$\begin{aligned} \lambda \int \eta ^2|\mathrm{d}u|^2 \leqq C\int \eta |\mathrm{d}u|\,|\mathrm{d}\eta |\,|u-(u)_{B_{2r}^2(x)}| +C\int \eta ^2|\mathrm{d}u|^2{\text {osc}}(u,B_{2r}^2(x)) \end{aligned}$$

and, applying Young’s inequality, it follows that

$$\begin{aligned} \int _{B_r^2(x)}|\mathrm{d}u|^2\leqq Cr^{-2}\int _{B_{2r}^2(x)}|u-(u)_{B_{2r}^2(x)}|^2 \leqq Cr^{-2}\left( \int _{B_{2r}^2(x)}|\mathrm{d}u|\right) ^2 \end{aligned}$$

whenever \({\text {osc}}(u,B_{2r}^2(x))\) is small enough. The classical Gehring’s lemma (see, for example, [17, Theorem V.1.2]) then implies that \(\mathrm{d}u\in L^r(B)\) for some \(r>2\) and any fixed ball \(B\subset \subset B_1^2\) (with r depending on B). Then the nonlinear term \(b_{kpq}\partial _i u^p\partial _i u^q\) is \(L^{r/2}(B)\) and standard elliptic regularity theory gives \(\mathrm{d}u\in L^s_{loc}(B)\), with \(\frac{1}{s}=\frac{2}{r}-\frac{1}{2}\), so that \(s>r\); iterating, we get \(\mathrm{d}u\in L^t_{loc}\) for any \(t<\infty \).

If we are in the half-ball case, then we can reduce to the previous case by reflection. We extend g and u to \({\tilde{g}}\) and \({\tilde{u}}\) on the ball \(B_1^2\), by means of the formula

$$\begin{aligned} g(s,-t):=Ug(s,t)U,\quad \begin{pmatrix}{\tilde{u}}^1 \\ \vdots \\ {\tilde{u}}^m\end{pmatrix}(s,-t):=U\begin{pmatrix}u^1 \\ \vdots \\ u^m\end{pmatrix}(s,t) \end{aligned}$$

for \((s,-t)\) in the lower half-ball, with \(U:=\begin{pmatrix}I_n &{}\quad \\ &{}\quad -I_{m-n}\end{pmatrix}\). Note that, by our hypotheses on g, \({\tilde{g}}\) is still continuous. Also, it is straightforward to check that \({\tilde{u}}\) solves

$$\begin{aligned} -\partial _i({\tilde{g}}_{jk}\partial _i {\tilde{u}}^j)+{\tilde{b}}_{kpq}\partial _i {\tilde{u}}^p\partial _i {\tilde{u}}^q=0, \end{aligned}$$

with \({\tilde{b}}_{kpq}\) extending \(b_{kpq}\) according to the following rule: if \(k\leqq n\) then \({\tilde{b}}_{kpq}(s,-t):=b_{kpq}(s,t)\) if p and q belong to the same set in the partition \(\{\{1,\dots ,n\},\{n+1,\dots ,m\}\}\), and \({\tilde{b}}_{kpq}(s,-t):=-b_{kpq}(s,t)\) otherwise; if \(k>n\) then the opposite holds. Then from the case of the full ball we deduce \(\mathrm{d}{\tilde{u}}\in L^t_{loc}\) for any \(t<\infty \). \(\square \)

Remark A.3

If the coefficients in (A.1) are smooth functions of u, then u is smooth. To check this, note that in the full ball case u is \(C^{0,\alpha }_{loc}\) for any \(\alpha <1\). The same is then true for the coefficients \(g_{jk}(u)\). Since the nonlinearity \(b_{kpq}\partial _i u^p\partial _i u^q\) belongs to \(L^r_{loc}\) for all \(r<\infty \), classical Schauder theory then gives \(\mathrm{d}u\in C^{0,\alpha }_{loc}\) for all \(\alpha <1\) and bootstrapping we reach \(u\in C^\infty \).

In the half-ball case, we can still argue in the same way that \(\mathrm{d}{\tilde{u}}\in C^{0,\alpha }_{loc}\) for all \(\alpha <1\). So \({\tilde{g}}\) is locally Lipschitz and we deduce \({\tilde{u}}\in W^{2,r}_{loc}\) for all \(r<\infty \). Differentiating the original equation in the first variable preserves the boundary conditions and leads to an equation of the form

$$\begin{aligned} \partial _i(g_{jk}\partial _i(\partial _1 u^j))+f_k=0 \end{aligned}$$

with \(f_k\in L^r_{loc}\) for all \(r<\infty \), and the same reflection trick (applied to \(w:=\partial _1 u\)) gives \(\partial _1 u\in W^{2,r}_{loc}\) for all \(r<\infty \). Iterating we get the same for all derivatives \(\partial _1^k u\). Now the equation allows to deduce inductively that \(u\in W^{k,r}_{loc}\) for all k, since \(g_{jk}(u)\Delta u^j=-\partial _i(g_{jk}(u))\partial _i u^j+b_{kpq}(u)\partial _i u^p\partial _i u^q\); this expresses \(\partial _{22}u\) in terms of \(\partial _{11}u\) and lower order derivatives and hence, for any multi-index \(\alpha =(\alpha _1,\alpha _2)\) with \(\alpha _2\geqq 2\), we deduce that \(\partial ^\alpha u=\partial _1^{\alpha _1}\partial _2^{\alpha _2}u\in L^r_{loc}\) for all \(r<\infty \) from the same property enjoyed by \(\partial _1^{\alpha _1+2}\partial _2^{\alpha _2-2}u\) and lower order derivatives of u.

The recent statements deal with general varifolds. It is clear that we can assume the smallness constant \(c_V\) appearing in all of them to be always the same.

Lemma A.4

There exists \(c_V({\mathcal {M}},{\mathcal {N}})>0\) with the following property. Given \(p\in {\mathcal {N}}\) and \(0<s<c_V\), for any 2-varifold \({\mathbf {v}}\) on \({\mathcal {M}}\) which is free boundary stationary outside \({\bar{B}}_s(p)\) and has density \(\theta \geqq {\bar{\theta }}\) on \({\text {spt}}(|{\mathbf {v}}|){\setminus }{\bar{B}}_s(p)\), either \({\text {spt}}(|{\mathbf {v}}|)\subseteq B_{2s}(p)\) or \(|{\mathbf {v}}|({\mathcal {M}}{\setminus }{\bar{B}}_s(p))\geqq c_V{\bar{\theta }}\).

Proof

Pick \(\gamma >0\) small, to be fixed along the proof; we will choose \(c_V\leqq \gamma \), so that the varifold is free boundary stationary outside \({\bar{B}}_\gamma (p)\) Possibly multiplying \({\mathbf {v}}\) by \({\bar{\theta }}^{-1}\), we can assume \({\bar{\theta }}=1\). Note that if \(q\in {\text {spt}}(|{\mathbf {v}}|){\setminus }B_{2\gamma }(p)\), then, by (4.8), we have

$$\begin{aligned} |{\mathbf {v}}|(B_\gamma (q)) \geqq c({\mathcal {M}},{\mathcal {N}})\gamma ^2\theta (|{\mathbf {v}}|,q) \geqq c({\mathcal {M}},{\mathcal {N}})\gamma ^2. \end{aligned}$$
(A.2)

Otherwise, \(|{\mathbf {v}}|\) is supported in \(B_{2\gamma }(p)\). Assume we are in this second case and pick a set of coordinates \((x_1,\dots ,x_m):B_{5\gamma }(p)\rightarrow {\mathbb {R}}^m\) centered at p, with \({\mathcal {N}}\) corresponding to \(\{x_{n+1}=\dots =x_m=0\}\). We can impose that \(\Vert g_{ij}-\delta _{ij}\Vert _{C^1}\leqq \gamma \) (in coordinates), for \(\gamma \) small, independently of \(p\in {\mathcal {N}}\).

On this ball, we define the vector field X to be \(X(x):=\chi (|x|)x_i\frac{\partial }{\partial x_i}\), where \(\chi :[0,\infty )\rightarrow [0,1]\) is smooth and such that \(\chi '\geqq 0\) on \([0,3\gamma ]\), \(\chi =1\) on \([\frac{5}{3}s,3\gamma ]\), \(\chi =0\) on \([0,\frac{4}{3}s]\cup [4\gamma ,\infty )\). Assuming \(\{|x|\leqq 4\gamma \}\subset \subset B_{5\gamma }(p)\), we can smoothly extend X to all of \({\mathcal {M}}\), with \(X=0\) outside the ball. For \(\gamma \) small enough (independently of p and \(s<\gamma \)), the \(C^1\) closeness of \(g_{ij}\) to \(\delta _{ij}\) guarantees

$$\begin{aligned} {\text {div}}_\Pi X\geqq 0 \end{aligned}$$

for all \((p,\Pi )\in {\text {Gr}}_2({\mathcal {M}})\) in the support of \({\mathbf {v}}\), since we can assume \({\text {spt}}(|{\mathbf {v}}|)\subseteq \{|x|<3\gamma \}\): indeed, here the contribution of \(\chi '\) is nonnegative, while the one of the position vector \(x_i\frac{\partial }{\partial x_i}\) is close to 2 (multiplied by \(\chi (|x|)\)). Also, the inequality is strict if \(|x(p)|\geqq \frac{5}{3}s\). Moreover, X is tangent to \({\mathcal {N}}\). We can also assume that \({\bar{B}}_s(p)\subset \subset \{|x|\leqq \frac{4}{3}s\}\); hence, we can test the stationarity of \({\mathbf {v}}\) against X and reach the contradiction

$$\begin{aligned} 0=\int _{(p,\Pi )\in {\text {Gr}}_2({\mathcal {M}})}{\text {div}}_\Pi X\,\mathrm{d}{\mathbf {v}}(p,\Pi )>0, \end{aligned}$$

unless \({\text {spt}}(|{\mathbf {v}}|)\) is contained in \(\{|x|\leqq \frac{5}{3}s\}\). Since the latter can be assumed to be included in \(B_{2s}(p)\), the statement follows from (A.2). \(\square \)

Remark A.5

The same statement holds if \({\mathbf {v}}\) is stationary outside \({\bar{B}}_s(p)\), without the assumption \(p\in {\mathcal {N}}\). The proof is analogous (but simpler, in that we do not need coordinates adapted to \({\mathcal {N}}\)).

Lemma A.6

There exist \(c_V>0\) and \(\delta :(0,\infty )^2\rightarrow (0,\infty )\), with \(\lim _{s\rightarrow 0}\delta (s,t)=0\) for every t, satisfying the following property. Given two points \(p_1,p_2\in {\mathcal {M}}\) and a radius \(s>0\), let \(B:={\bar{B}}_s(p_1)\cup {\bar{B}}_s(p_2)\); if a 2-varifold \({\mathbf {v}}\) on \({\mathcal {M}}\) is free boundary stationary outside B, has density \(\theta \geqq {\bar{\theta }}\) on \({\text {spt}}(|{\mathbf {v}}|){\setminus }B\) and satisfies the bound

$$\begin{aligned} |{\mathbf {v}}|(B_r(q))\leqq c'r^2\quad \text {for all }q\in {\mathcal {M}},\ r>0, \end{aligned}$$

then either \(|{\mathbf {v}}|({\mathcal {M}})\leqq {\bar{\theta }}\delta (s,c'/{\bar{\theta }})\) or \(|{\mathbf {v}}|({\mathcal {M}})\geqq c_V{\bar{\theta }}\). The constant \(c_V\) and the function \(\delta \) depend only on \({\mathcal {M}}\) and \({\mathcal {N}}\).

Proof

We can assume \({\bar{\theta }}=1\). From (4.8) it follows that any nontrivial free boundary stationary varifold \({\mathbf {v}}'\) with density at least 1 on \({\text {spt}}(|{\mathbf {v}}'|)\) has \(|{\mathbf {v}}'|({\mathcal {M}})\geqq \lambda ({\mathcal {M}},{\mathcal {N}})\). Let \(\delta (s,c')\) be the supremum of all possible masses \(|{\mathbf {v}}|({\mathcal {M}})\) which are smaller than \(c_V\), for \({\mathbf {v}}\) as in the statement, with \(c_V\) to be specified below. Take a sequence \(s_k\rightarrow 0\) of positive numbers and a sequence \({\mathbf {v}}_k\) satisfying the assumptions with \(s=s_k\), as well as \(\delta (s_k,c')-2^{-k}<|{\mathbf {v}}_k|({\mathcal {M}})<c_V\).

Up to subsequences we get a limit varifold \({\mathbf {v}}_\infty \) which is free boundary stationary on the complement of two points \({\bar{p}}_1\) and \({\bar{p}}_2\). We still have \(|{\mathbf {v}}_\infty |(B_r(q))\leqq c'r^2\) for all centers q and all radii r. This upper bound implies easily that actually \({\mathbf {v}}_\infty \) is free boundary stationary on the full manifold see the proof of Theorem 5.12 for the details; also, by (4.8), it has a lower bound \(c\leqq 1\) for its density on \({\text {spt}}(|{\mathbf {v}}_\infty |)\). Hence, \(|{\mathbf {v}}_\infty |({\mathcal {M}})\geqq c\lambda \) unless \({\mathbf {v}}_\infty =0\).

Since \(|{\mathbf {v}}_\infty |({\mathcal {M}})=\lim _{k\rightarrow \infty }|{\mathbf {v}}_k|({\mathcal {M}})\leqq c_V\), choosing any \(c_V<c\lambda \) forces \({\mathbf {v}}_\infty =0\), so that \(\delta (s_k,c')\rightarrow 0\). This shows that \(\delta (s,c')\rightarrow 0\) as \(s\rightarrow 0\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pigati, A. The Viscosity Method for Min–Max Free Boundary Minimal Surfaces. Arch Rational Mech Anal 244, 391–441 (2022). https://doi.org/10.1007/s00205-022-01761-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-022-01761-9

Navigation