Skip to main content

Advertisement

Log in

Inverse mean curvature flow of rotationally symmetric hypersurfaces

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

We prove that the Inverse Mean Curvature Flow of a non-star-shaped, mean-convex embedded sphere in \(\mathbb {R}^{n+1}\) with symmetry about an axis and sufficiently long, thick necks exists for all time and homothetically converges to a round sphere as \(t \rightarrow \infty \). Our approach is based on a localized version of the parabolic maximum principle. We also present two applications of this result. The first is an extension of the Minkowski inequality to the corresponding non-star-shaped, mean-convex domains in \(\mathbb {R}^{n+1}\). The second is a connection between IMCF and minimal surface theory. Based on previous work by Meeks and Yau (Math Z 179:151–168, 1982) and using foliations by IMCF, we establish embeddedness of the solution to Plateau’s problem and a finiteness property of stable immersed minimal disks for certain Jordan curves in \(\mathbb {R}^{3}\).

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Almgren, F., Simon, L.: Existence of embedded solutions of Plateau’s problem. Annali Della Scuola Normale Superiore Di Pisa-classe Di Scienze 6, 447–495 (1979)

    MathSciNet  MATH  Google Scholar 

  2. Altschuler, S., Angenent, S., Giga, Y.: Mean curvature flow through singularities for surfaces of rotation. J. Geom. Anal. 5, 293–358 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  3. Angenent, S.B., Velázquenz, J.J.: Degenerate neckpinches in mean curvature flow. J. für Reine Angew. Math. 482, 15–66 (1997)

    MathSciNet  MATH  Google Scholar 

  4. Athanassenas, M.: Volume-preserving mean curvature flow of rotationally symmetric surfaces. Comment. Math. Helv. 72, 52–66 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  5. Athanassenas, M., Kandanaarachchi, S.: On the convergence of axially symmetric volume preserving mean curvature flow. Pac. J. Math. 259(1), 41–54 (2012). arXiv:1108.5849

    Article  MathSciNet  MATH  Google Scholar 

  6. Athanassenas, M., Kandanaarachchi, S.: Singularities of axially symmetric volume preserving mean curvature flow (2012)

  7. Beeson, M.: A real-analytic Jordan curve cannot bound infinitely many relative minima of area. In: Differential Geometry (2006)

  8. Beeson, M.: The \(6\pi \) theorem about minimal surfaces. Pac. J. Math. 117(1), 17–25 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  9. Choi, B., Daskalopolous, P.: Evolution of non-compact hypersurfaces by inverse mean curvature (Dec. 2018)

  10. Coskunuzer, B.: Embedded plateau problem. Trans. Am. Math. Soc. 364(3), 1211–1224 (2012). https://doi.org/10.1090/s0002-9947-2011-05486-3

    Article  MathSciNet  MATH  Google Scholar 

  11. Ecker, K., Huisken, G.: Interior estimates for hypersurfaces moving by mean curvature. Invent. Math. 105, 547–569 (1991). https://doi.org/10.1007/BF01232278

    Article  MathSciNet  MATH  Google Scholar 

  12. Ekholm, T., White, B., Wienholtz, D.: Embeddedness of minimal surfaces with total boundary curvature at most \(4\pi \). Ann. Math. 155, 209–234 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Gage, M., Hamilton, R.S.: The heat equation shrinking convex plane curves. J. Differ. Geom. 23(1), 69–96 (1986). https://doi.org/10.4310/jdg/1214439902

    Article  MathSciNet  MATH  Google Scholar 

  14. Gang, Z., Knopf, D., Sigal, I.M.: Neckpinch dynamics for asymmetric surfaces evolving by mean curvature flow (Nov. 2013)

  15. Gerhardt, C.: Flow of nonconvex hypersurfaces into spheres. J. Differ. Geom. 32, 299–314 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  16. Grayson, M.A.: A short note on the evolution of a surface by its mean curvature. Duke Math. 58(3), 555–558 (1989). https://doi.org/10.1215/S0012-7094-89-05825-0

    Article  MathSciNet  MATH  Google Scholar 

  17. Guan, P., Li, J.: The quermassintegral inequalities for k-convex starshaped domains. Adv. Math. 221, 1725–1732 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  18. Hardt, R., Simon, L.: Boundary regularity and embedded solutions for the oriented plateau problem. Bull. Am. Math. Soc. 1 (1979). https://doi.org/10.1090/S0273-0979-1979-14581

  19. Harvie, B.: Inverse mean curvature flow over non-star-shaped surfaces. Math. Res. Lett. (Aug. 2020). arXiv:1909.01328 [math.DG]

  20. Harvie, B.: The inverse mean curvature flow: singularities, dynamical stability, and applications to minimal surfaces. Ph.D Thesis. University of California, Davis (2021)

  21. Head, J., Kandanaarachchi, S.: Singularity formation in axially symmetric mean curvature flow with Neumann boundary (2019). arXiv:1908.02871 [math.DG]

  22. Huisken, G.: Asymptotic behavior for singularities of the mean curvature flow. J. Differ. Geom. 31(1), 285–299 (1990). https://doi.org/10.4310/jdg/1214444099

    Article  MathSciNet  MATH  Google Scholar 

  23. Huisken, G.: Flow by mean curvature of convex surfaces into spheres. J. Differ. Geom. 20(1), 237–266 (1984). https://doi.org/10.4310/jdg/1214438998

    Article  MathSciNet  MATH  Google Scholar 

  24. Huisken, G., Ilmanen, T.: Higher regularity of the inverse mean curvature flow. J. Differ. Geom 80 (Nov. 2008). https://doi.org/10.4310/jdg/1226090483

  25. Huisken, G., Polden, A.: Geometric evolution equations for hypersurfaces. In: Hildebrandt, S., Struwe, M. (eds) Calculus of Variations and Geometric Evolution Problems: Lectures given at the 2nd Session of the Centro Internazionale Matematico Estivo (C.I.M.E.) held in Cetraro, Italy, June 15–22 (1996), pp. 45–84. Springer, Berlin (1999). https://doi.org/10.1007/BFb0092669

  26. Lumer, G.: Principes du maximum paraboliques pour des domaines (x,t) noncylindriques. In: Hirsch, F., Mokobodzki, G. (eds) Séminaire de Théorie du Potentiel Paris, No. 8, pp. 105–113. Springer, Berlin (1987)

  27. Mantegazza, C.: Lecture Notes on Mean Curvature Flow (2011)

  28. Meeks, W., Yau, S.: The existence of embedded minimal surfaces and the problem of uniqueness. Math. Z. 179, 151–168 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  29. Morgan, F.: On finiteness of the number of stable minimal hypersurfaces with a fixed boundary. Indiana Univ. Math. J. 35 (Apr. 1985). https://doi.org/10.1090/S0273-0979-1985-15396-0

  30. Osserman, R.: A proof of the regularity everywhere of the classical solution to Plateau’s problem. Ann. Math. 91(3), 550–569 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  31. Rossman, W.: On embeddedness of area-minimizing disks, and an application to constructing complete minimal surfaces (2008). arXiv:0804.4208

  32. Urbas, J.I.E.: On the expansion of starshaped hypersurfaces by symmetric functions of their principal curvatures. Math. Z. 205(3), 355–372 (1990). https://doi.org/10.1007/BF02571249

    Article  MathSciNet  MATH  Google Scholar 

  33. White, B.: Lectures on Minimal Surface Theory (2016). arXiv:1308.3325 [math.DG]

Download references

Acknowledgements

I would like to thank the referee for valuable feedback and suggestions. I also would like to thank Professors Joel Hass and Adam Jacob of the University of California, Davis and Professor Mao-Pei Tsui of National Taiwan University for many helpful discussions. Finally, I would like to thank the University of California, Davis Department of Mathematics and the National Center for Theoretical Sciences, Mathematics Division at National Taiwan University for their financial support throughout my graduate and postdoctoral studies.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian Harvie.

Additional information

Communicated by Andrea Mondino.

Publisher's Note

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

A Appendices

A Appendices

1.1 A.1 Non-cylindrical maximum principle

We recall Theorem 4.4:

Theorem A.1

Let \(F: N^{n} \times [0,T) \rightarrow \mathbb {R}^{n+1}\) be a solution of the Inverse Mean Curvature Flow (1.1) over a closed manifold N. For a domain \(U \subset N \times [0,T)\) and \(f \in C^{2,1}(U) \cap C(\overline{U})\), suppose for a smooth vector field \(\eta \) over U we have

$$\begin{aligned} \left( \partial _{t} - \frac{1}{H^{2}} \Delta \right) f \le \langle \eta , \nabla f \rangle . \end{aligned}$$

(Resp. \(\ge \) at a minimum) Here \(\Delta \) and \(\nabla \) are the Laplacian and gradient operators over \(N_{t}\), respectively. Then

$$\begin{aligned} \sup _{U} f \le \sup _{\partial _{P} U} f \end{aligned}$$

(Resp. \(\inf _{U} f \ge \inf _{\partial _{P} U} f\)). Furthermore, suppose that f has a positive supremum over U and that each \((x_{0},t_{0}) \in \partial _{P} U {\setminus } \tilde{\partial }_P U\) is a limit point of \(U \cap \{ t < t_{0} \}\). Then

$$\begin{aligned} \sup _{U} f \le \sup _{\tilde{\partial }_{P} U} f \end{aligned}$$
(A.61)

(Resp. \(\inf _{U} f \ge \inf _{\tilde{\partial }_{P} U} f\) for a positive minimum).

Proof of Theorem 4.4

For the first part, we follow the proof in the Appendix of [6]. For a given smooth vector field \(\eta \) over U we have by hypothesis

$$\begin{aligned} \left( \partial _{t} - \frac{1}{H^{2}} \Delta \right) f \le \langle \eta , \nabla f \rangle . \end{aligned}$$

We argue by contradiction: define the function \(\tilde{f}(x,t)=f(x,t)- \epsilon t\) for some \(\epsilon >0\). Then

$$\begin{aligned} \partial _{t} \tilde{f}= \partial _{t} f- \epsilon , \quad \partial _{i} \tilde{f} = \partial _{i} f, \quad \partial _{ij} \tilde{f} = \partial _{ij} f. \end{aligned}$$

The operator over \(\tilde{f}\) must then obey

$$\begin{aligned} \left( \partial _{t} - \frac{1}{H^{2}} \Delta - \eta \cdot \nabla \right) \tilde{f} < 0. \end{aligned}$$
(A.62)

On the other hand, at any interior maximum \((x_{0},t_{0}) \in U\) of \(\tilde{f}\), the criteria for a local maximum dictate that at \((x_{0},t_{0})\)

$$\begin{aligned} \partial _{t} \tilde{f} \ge 0, \quad \partial _{i} \tilde{f}=0, \quad \partial _{ij} \tilde{f} \le 0, \end{aligned}$$

where the last inequality is in the operator-theoretic sense for the symmetric matrix \(\partial _{ij} \tilde{f}\). Writing

$$\begin{aligned} \Delta \tilde{f}= g^{ij} (\partial _{ij} \tilde{f} - \Gamma ^{k}_{ij} \partial _{k} \tilde{f}), \quad \nabla \tilde{f} = g^{ij} \partial _{j} \tilde{f} \partial _{i} \vec {F} \end{aligned}$$

in view of the positivity of \(g_{ij}\), we see \(\Delta \tilde{f} \le 0\) and \(\nabla f=0\). Hence

$$\begin{aligned} \left( \partial _{t} - \frac{1}{H^{2}} \Delta - \eta \cdot \nabla \right) \tilde{f}(x_{0},t_{0}) \ge 0, \end{aligned}$$

contradicting (A.62). So \(\tilde{f}\) has no interior maximum and thus

$$\begin{aligned} \sup _{U} f - \epsilon T \le \sup _{U} \tilde{f} \le \sup _{\partial _{P} U} f. \end{aligned}$$

Then \(\sup _{U} f \le \sup _{\partial _{P} U} f + \epsilon T\). For \(T< \infty \), letting \(\epsilon \rightarrow 0\) yields the result. To prove the statement for the infimum, take \(\tilde{f}=f- \epsilon t\), \(\epsilon >0\), and repeat this argument for a minimum.

For the second part, we show \(\sup _{U} f \le \sup _{\tilde{\partial }_{P} U} f\) if \(\sup _{U} f > 0\). Define \(Z= \partial _{P} U {\setminus } \tilde{\partial }_{P} U\), and for each \(t \in [0,T)\) let \(Z_{t}= Z \cap \{ t \}\) be the cross sections of Z. We argue by contradiction: suppose (A.61) does not hold. Then the maximum of f does not occur on \(\partial _{\tilde{P}} U\) nor does it occur at an interior point of U, so it must occur on the set Z. Call \(Z_{\max }\) the union of \(Z_{t}\)’s on which the maximum is achieved, and let \(t_{*}>0\) be the first time at which \(\sup _{U} f\) is achieved on \(Z_{\max }\). Pick \(\beta > 0\) such that

$$\begin{aligned} \sup _{U} f> \beta > \sup _{\partial _{P} U \setminus Z_{\max }} f, \end{aligned}$$

and define

$$\begin{aligned} Y_{\beta } = \{ (x,t) \in U| f(x,t) > \beta \}. \end{aligned}$$

\(\overline{Y_{\beta }}\) must intersect \(Z_{t_{*}}\). Consider \((x_{0},t_{*}) \in Z_{t_{*}} \cap \overline{Y_{\beta }}\). From the additional assumption in the proposition there is a sequence of points \((x_{n},t_{n}) \in U\) with \(t_{n} < t_{*}\) converging to \((x_{0},t_{*})\). By continuity of f, \(f(x_{n},t_{n}) > \beta \) for large enough n. This means that the set

$$\begin{aligned} X_{\beta }=\{ t < t_{*} \} \cap Y_{\beta } \end{aligned}$$
(A.63)

is nonempty and open. By openness, we pick a time \(t_{1} < t_{*}\) so that the set \(X_{\beta } \cap \{ t \le t_{1} \} \ne \varnothing \).

Fig. 6
figure 6

The cutoff function \(\phi \) is 1 for times less than \(t_{1}\) and 0 for times greater than \(t_{2}\). This guarantees that the supremum of \(\phi f\) occurs at an interior point of \(X_{\beta }'\)

Fix a time \(t_{2} \in (t_{1},t_{*})\) and choose a cutoff function \(\phi : [0,T] \rightarrow [0,1]\) such that \(\phi (t)=1\) when \(t \in [0,t_{1}]\), \(\phi '(t) < 0\) when \(t \in (t_{1}, t_{2})\), and \(\phi (t)=0\) when \(t \in [t_{2},T]\), see Fig. 6. Since \((\phi f) (x, t_{1})= f(x, t_{1}) > \beta \) for \((x,t_{1}) \in X_{\beta }\), we know \(\sup _{U} \phi f> \beta >0\).

We calculate

$$\begin{aligned} \partial _{t} (\phi f) = \phi \partial _{t} f + \phi ' f, \quad \Delta \phi f = \phi \Delta f, \quad \nabla (\phi f) = \phi \nabla f, \end{aligned}$$

so

$$\begin{aligned} \left( \partial _{t} - \frac{1}{H^{2}} \Delta - \eta \cdot \nabla \right) (\phi f) = \phi \left( \partial _{t} - \frac{1}{H^{2}} \Delta - \eta \cdot \nabla \right) f + \phi ' f. \end{aligned}$$
(A.64)

By hypothesis we have \((\partial _{t} - \frac{1}{H^{2}} \Delta - \eta \cdot \nabla ) f (x,t) \le 0\) and \(\phi ' \le 0\). Since \(\sup _{U} \phi f > 0 \), any interior point \((x_{0},t_{0}) \in U\) at which \(\phi f\) achieves this supremum would need to satisfy

$$\begin{aligned} \left( \partial _{t} - \frac{1}{H^{2}} \Delta - \eta \cdot \nabla \right) (\phi f) (x_{0},t_{0}) < 0. \end{aligned}$$

By the same argument used for the first part of Theorem 4.4, this is impossible, and so \(\sup _{U} (\phi f) = \sup _{\partial _{P} U} (\phi f)\). However, \(\sup _{\partial _{P} U {\setminus } Z_{\max }} \phi f \le \sup _{\partial _{P} U {\setminus } Z_{\max }} f < \beta \) by hypothesis, and \(\phi f|_{Z_{\max }} \equiv 0\) as \(\phi =0\) on \([t_{*},T)\). This would altogether yield

$$\begin{aligned} \sup _{\partial _{P} U} \phi f< \beta < \sup _{U} \phi f, \end{aligned}$$

contradicting the first part of the non-cylindrical maximum principle. Conclude then that

$$\begin{aligned} f(x,t) \le \sup _{\tilde{\partial }_{P}U} f \end{aligned}$$

on U. The statement may be shown for a minimum by choosing \(\beta >0\) with \(\inf _{U} f< \beta < \inf _{\partial _{P} U {\setminus } Z_{\max }} f\).

Remark A.2

The version of this principle used in [6, 21] does not include the hypothesis that U approaches \(Z_{t_{*}}\) from below in time. However, if U only touches \(Z_{t_{*}}\) from above in time, \(Y_{\beta } \cap \{ t < t_{*}\}\) may be empty. The corresponding cutoff function would then need to be chosen to increase with t, so that the last term in (A.64) is possibly non-negative. Therefore, this additional hypothesis seems to be necessary.

1.2 A.2 Non-star-shaped admissible initial data

Proposition A.3

For any \(n \ge 2\), there is an admissible surface \(N_{0}^{n} \subset \mathbb {R}^{n+1}\) which is not star-shaped.

Proof

We will begin with an example of a non-star-shaped admissible surface which is of class \(C^{2}\). Define the domain \(E_{0} \subset \mathbb {R}^{n+1}\) by

$$\begin{aligned} E_{0}= \{ -10< x_{1}< 10 | (x_{2}^{2} + \cdots x_{n+1}^{2})^{\frac{1}{2}} < y(x_{1}) \}, \end{aligned}$$
(A.65)

where \(y: (-10,10) \rightarrow \mathbb {R}^{+}\) is the even extension of the function

$$\begin{aligned} y(x) = {\left\{ \begin{array}{ll} \frac{3}{4} &{} x \in [0,8) \\ \frac{3}{4} + 2 (x-8)^{3} - \frac{11}{4}(x-8)^{4} + (x-8)^{5} &{} x \in [8,9) \\ \sqrt{1-(x-9)^{2}} &{} x \in [9,10), \end{array}\right. } \end{aligned}$$
(A.66)

across \(x=0\). Direct computation shows that \(y \in C^{2}((-10,10))\). One can also verify through either computing maxima or graphing that

$$\begin{aligned} \frac{3}{4} \le y(x) (1 + |y'(x)|^{2})^{\frac{1}{2}}\le & {} \frac{5}{4} \end{aligned}$$
(A.67)
$$\begin{aligned} \frac{y''(x)y(x)}{1+|y'(x)|^{2}}< & {} 1 , \end{aligned}$$
(A.68)

for each \(x \in (-10,10)\). The left-hand side of (A.68) corresponds to the ratio \(-\frac{k}{p}\) of the principal curvatures of \(N_{0}\), and so this bound implies

$$\begin{aligned} \min _{N_{0}} H >0, \end{aligned}$$
(A.69)

The function in (A.67) corresponds to \(p^{-1}\) on \(N_{0}\), meaning

$$\begin{aligned} \frac{\max _{N_{0}} p}{\min _{N_{0}} p} < 2 \le n^{\frac{n}{2(n-1)}}, \quad n \ge 2, \end{aligned}$$
(A.70)

so altogether \(N_{0}\) is mean-convex and admissible.

To demonstrate that \(N_{0}\) is not star-shaped, it is sufficient to consider graph of the function y in the \(x-y\) plane. In this plane, the line segment connecting (0, 0) to (9, 1), parametrized by \(L(x)=\frac{1}{9}x\), satisfies

$$\begin{aligned} L(8) = \frac{8}{9} > \frac{3}{4}= y(8), \end{aligned}$$

meaning this segment must intersect the graph of y at another point by the intermediate value property. Furthermore, for any \(x_{0} \le 0\) and \(0 \le y_{0} < y(x_{0})\), the line segment connecting \((x_{0},y_{0})\) to (9, 1) satisfies

$$\begin{aligned} L_{(x_{0},y_{0})} (x) \ge L(x), \quad x \in [0,9], \end{aligned}$$

for the L(x) defined above. So \(L_{(x_{0},y_{0})}(8) > \frac{3}{4}\), and once again by the intermediate value property this segment intersects the graph somewhere else. If we choose \(x_{0} \ge 0\), \(0 \le y_{0} < y(x_{0})\), the segment connecting \((x_{0},y_{0})\) to \((-9,1)\) must also intersect since y is even (Fig. 7).

Fig. 7
figure 7

A graph of the function y(x) defined in (A.66). The admissible surface generated by y(x) is not star-shaped because of its long neck

For any \(x_{0} \in E_{0}\), let \(P \subset \mathbb {R}^{n+1}\) be the 2-plane containing \(x_{0}\) and \(X_{1}\) (if \(x_{0}\) lies on this axis, let P be any 2-plane containing \(X_{1}\)). Choose Cartesian coordinates (xy) on P so that x-axis is \(X_{1}\) and the y-coordinate of \(x_{0}\) is non-negative. Then either the line segment connecting \(x_{0}\) to \((9,1) \in P\) or the line segment connecting \(x_{0}\) to \((-9,1) \in P\) is not contained in \(E_{0}\). Since these points lie on \(N_{0}\), \(N_{0}\) cannot be star-shaped with respect to \(x_{0}\).

To conclude, we must show that \(N_{0}\) is approximated in a \(C^{2}\) sense by smooth rotationally symmetric hypersurfaces. For P as above, \(N_{0} \cap P= \gamma \) is a closed plane curve parametrized over \({[0,1]}/{\mathbb {Z}} \simeq \mathbb {S}^{1}\) by

$$\begin{aligned} \gamma (t)= (x_{1}(t),x_{2}(t))= {\left\{ \begin{array}{ll} (10(4t-1), y(10(4t-1))) &{} 0 \le t \le \frac{1}{2}, \\ (10(3-4t), -y(10(3-4t)) &{} \frac{1}{2} \le t \le 1. \end{array}\right. } \end{aligned}$$
(A.71)

By convolving with a symmetric mollifier \(\psi _{\epsilon }\) over \(\mathbb {S}^{1}\), we obtain smooth functions \(x^{\epsilon }_{1}(t)\) and \(x^{\epsilon }_{2}(t)\) on \(\mathbb {S}^{1}\) which are respectively even and odd about \(t=\frac{1}{2}\). For \(\epsilon \) small enough, the corresponding smooth curve \(\gamma ^{\epsilon }(t)=(x^{\epsilon }_{1}(t),x^{\epsilon }_{2}(t))\) in P is uniformally convex over small intervals containing \(t=0\) and \(t=\frac{1}{2}\), and away from these intervals \(x^{\epsilon }_{1}(t)\) and \(x^{\epsilon }_{2}(t)\) converge in uniform \(C^{2}\) topology to \(x_{1}(t)\) and \(x_{2}(t)\) as \(\epsilon \rightarrow 0\). Then \(\gamma _{\epsilon }\) is the cross section \(N^{\epsilon }_{0} \cap P\) of a smooth, rotationally symmetric hypersurface \(N^{\epsilon }_{0}\) which inherits properties (A.69) and (A.70) and fails to be star-shaped. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Harvie, B. Inverse mean curvature flow of rotationally symmetric hypersurfaces. Calc. Var. 62, 125 (2023). https://doi.org/10.1007/s00526-023-02469-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-023-02469-8

Mathematics Subject Classification

Navigation