Skip to main content
Log in

Volume preserving mean curvature flow for star-shaped sets

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

Abstract

We study the evolution of star-shaped sets in volume preserving mean curvature flow. Constructed by approximate minimizing movements, our solution preserves a strong version of star-shapedness. We also show that the solution converges to a ball as time goes to infinity. For asymptotic behavior of the solution we use the gradient flow structure of the problem, whereas a modified notion of viscosity solutions is introduced to study the geometric properties of the flow by moving planes method.

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. Antonopoulou, D.C., Karali, G., Sigal, I.M.: Stability of spheres under volume-preserving mean curvature flow. Dyn. PDE 7(4), 327–344 (2010)

    MathSciNet  MATH  Google Scholar 

  2. Andrews, B.: Volume-preserving anisotropic mean curvature flow. Indiana Univ. Math. J. 50(2), 783–827 (2001)

    MathSciNet  MATH  Google Scholar 

  3. Athanassenas, M.: Volume-preserving mean curvature flow of rotationally symmetric surfaces. Comment. Math. Helv. 72(1), 52–66 (1997)

    MathSciNet  MATH  Google Scholar 

  4. Barles, G.: An introduction to the theory of viscosity solutions for first-order Hamilton–Jacobi equations and applications. Hamilton–Jacobi Equations: Approximations, Numerical Analysis and Applications, pp. 49–109. Springer, Berlin (2013)

    MATH  Google Scholar 

  5. Bellettini, G., Caselles, V., Chambolle, A., Novaga, M.: The volume preserving crystalline mean curvature flow of convex sets in RN. J. Math. Pures Appl. 92(5), 499–527 (2009)

    MathSciNet  MATH  Google Scholar 

  6. Bourgoing, M.: Viscosity solutions of fully nonlinear second order parabolic equations with \(l^1\) dependence in time and neumann boundary conditions. Discrete Contin. Dyn. Syst. A 21(3), 763–800 (2008)

    MathSciNet  MATH  Google Scholar 

  7. Bourgoing, M.: Viscosity solutions of fully nonlinear second order parabolic equations with \(l^1\) dependence in time and neumann boundary conditions. existence and applications to the level-set approach. Discrete Contin. Dyn. Syst. A 21(4), 1047–1069 (2008)

    MathSciNet  MATH  Google Scholar 

  8. Barles, G., Soner, H.M., Souganidis, P.E.: Front propagation and phase field theory. SIAM J. Control Optim. 31(2), 439–469 (1993)

    MathSciNet  MATH  Google Scholar 

  9. Chen, Y.G., Giga, Y., Goto, S.: Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations. J. Differ. Geom. 33(3), 749–786 (1991)

    MathSciNet  MATH  Google Scholar 

  10. Caffarelli, L.A., Salsa, S.: A Geometric Approach to Free Boundary Problems, vol. 68. American Mathematical Society, Providence (2005)

    MATH  Google Scholar 

  11. Ecker, K.: Regularity Theory for Mean Curvature Flow. Springer, Boston (2004)

    MATH  Google Scholar 

  12. Evans, L.C., Gariepy, R.F.: Measure Theory and Fine Properties of Functions. Studies in Advanced Mathematics. CRC Press, Boca Raton (1992)

    MATH  Google Scholar 

  13. Ecker, K., Huisken, G.: Interior estimates for hypersurfaces moving by mean curvature. Invent. Math. 105(3), 547–569 (1991)

    MathSciNet  MATH  Google Scholar 

  14. Escher, J., Simonett, G.: The volume preserving mean curvature flow near spheres. Proc. Am. Math. Soc. 126(9), 2789–2796 (1998)

    MathSciNet  MATH  Google Scholar 

  15. Feldman, W.M., Kim, I.C.: Dynamic stability of equilibrium capillary drops. Arch. Rational Mech. Anal. 211(3), 819–878 (2014)

    MathSciNet  MATH  Google Scholar 

  16. Giga, Y.: Surface Evolution Equations, vol. 99. Birkhäuser, Basel (2006). Monographs in Mathematics

    MATH  Google Scholar 

  17. Gilbarg, D., Trudinger, N.S.: Elliptic Partial Differential Equations of Second Order. Springer, Berlin (2015)

    MATH  Google Scholar 

  18. Huisken, G.: The volume preserving mean curvature flow. J. Reine Angew. Math. 382(35–48), 78 (1987)

    MathSciNet  MATH  Google Scholar 

  19. Ishii, H.: Hamilton-Jacobi equations with discontinuous Hamiltonians on arbitrary open sets. 28, 33–77 (1985)

    Google Scholar 

  20. Kim, I., Kwon, D.: On mean curvature flow with forcing. Commun. Partial Differ. Equ. 45, 414–455 (2019)

    MathSciNet  Google Scholar 

  21. Li, H.: The volume-preserving mean curvature flow in Euclidean space. Pac. J. Math. 243(2), 331–355 (2009)

    MathSciNet  MATH  Google Scholar 

  22. Maggi, F.: Sets of Finite Perimeter and Geometric Variational Problems: An Introduction to Geometric Measure Theory. Cambridge University Press, Cambridge (2012)

    MATH  Google Scholar 

  23. Mugnai, L., Seis, C., Spadaro, E.: Global solutions to the volume-preserving mean-curvature flow. Calc. Var. Partial. Differ. Equ. 55(1), 18–23 (2016)

    MathSciNet  MATH  Google Scholar 

  24. Takasao, K.: Existence of weak solution for volume preserving mean curvature flow via phase field method. Indiana Univ. Math. J. 66, 2015–2035 (2017)

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dohyun Kwon.

Additional information

Communicated by Y. Giga.

Publisher's Note

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

Inwon Kim was partially supported by NSF DMS-1566578. Dohyun Kwon was partially supported by Kwanjeong Educational Foundation.

Appendices

Appendix A: Examples

1.1 A.1: Examples of unbounded total mean curvature

In this section, we present examples of unbounded total mean curvature for \({\mathbb R}^2\) in Example A.1 and \({\mathbb R}^n, n \ge 3\) in Example A.2. Here is an example of a non-convex domain in \({\mathbb R}^2\) such that the total mean curvature is unbounded although perimeter and volume are bounded.

Example A.1

Let \(\{O_{i}\}_{i \in {\mathbb N}}\) be a sequence of mutually disjoint balls in \({\mathbb R}^2\) with radius \(r_{i}:=\frac{1}{i^2}\) such that \(O_{i} \subset \subset B_{10}\) for all \(i \in {\mathbb N}\). Define a domain \(\Omega \) by

$$\begin{aligned} \Omega := B_{10} - \bigcup _{i=1}^{\infty } O_i. \end{aligned}$$
(A.1)

Gauss–Bonnet theorem implies \(\int _{\partial O_i} H = -2\pi \) for all \(i \in {\mathbb N}\). Thus, total mean curvature is unbounded although perimeter and volume are positive and bounded as follows,

$$\begin{aligned} |\Omega | = w_n \left( 10^2 - \sum _{i=1}^{\infty } \frac{1}{i^4} \right) , \quad \text {Per}(\Omega ) = n w_n \left( 10- \sum _{i=1}^{\infty } \frac{1}{i^2} \right) . \end{aligned}$$
(A.2)

Based on Example A.1, we can construct a simply connected and star-shaped set, whose total mean curvature is unbounded for \(n\ge 3\) (see Example A.2 and Lemma A.3).

Example A.2

Let \(\phi :{\mathbb R}^n \rightarrow {\mathbb R}^+\) be a smooth function such that \(\phi \) is concave,

$$\begin{aligned} 0< \phi <1 \hbox { in }B_1^{n-1}, \quad \phi = 0 \hbox { in }(B_1^{n-1})^c \hbox { and } |D\phi | \le 1 \hbox { in } {\mathbb R}^n. \end{aligned}$$
(A.3)

Denote

$$\begin{aligned} {\mathcal D}&:= \{ [0,\phi (x')] \times x' : x' \in B_1^{n-1} \} \subset {\mathbb R}^n, \quad \nonumber \\ \partial ^- D&:= B_1^{n-1} \text { and } \partial ^+ D:= \{ (\phi (x'),x') : x' \in B_1^{n-1} \}. \end{aligned}$$
(A.4)

Here, \(B_1^{n-1}\) is a ball of radius 1 and center 0 in \({\mathbb R}^{n-1}\).

As Example A.1, choose \(\{ x_i \}_{ i \in {\mathbb N}} \subset {\mathbb R}^{n-1}\) such that \(B_{r_i}^{n-1}(x_i)\) are a sequence of disjoint balls such that

$$\begin{aligned} r_{i}:=\frac{1}{i^{\frac{1}{n-2}}} \hbox { and } x_i \in B_{10}^{n-1}. \end{aligned}$$
(A.5)

Define a domain \(\Omega \subset {\mathbb R}^n\) by

$$\begin{aligned} \Omega := \Omega _1 \cup \Omega _2 \hbox { where } \Omega _1 := [-40,40]^n \quad \hbox { and } \quad \Omega _2 := \bigcup _{i=1}^{\infty } (r_i {\mathcal D}+ (40, x_i)) . \end{aligned}$$
(A.6)

where \(e_1\) is a unit vector of the first axis. From Lemma A.3, \(\Omega \) is star-shaped with respect to some ball.

The boundary is divided by two parts,

$$\begin{aligned} {\partial \Omega _t}&= \Gamma _1 \cup \Gamma _2 \hbox { where } \Gamma _1:= \left( \partial \Omega _1 {\setminus } \left( \bigcup _{i=1}^{\infty } r_i \partial {\mathcal D}^-+ (40,x_i) \right) \right) \hbox { and }\nonumber \\ \Gamma _2&:= \bigcup _{i=1}^{\infty } (r_i\partial {\mathcal D}^+ + (40,x_i)). \end{aligned}$$
(A.7)

Note that \(\Gamma _1\) has bounded total mean curvature, but total mean curvature of \(\Gamma _2\) is unbounded as follows. From (A.5) and the change of variables, it holds that

$$\begin{aligned} \int _{\Gamma _2} H d\sigma = \sum _{i=1}^{\infty } \int _{r_i\partial {\mathcal D}^+ + x_i} H d\sigma = \left( \int _{\partial {\mathcal D}^+} H d\sigma \right) \sum _{i=1}^{\infty } r_{i}^{n-2}. \end{aligned}$$
(A.8)

Then, (A.7) implies that

$$\begin{aligned} \int _{\Gamma _2} H d\sigma = \left( \int _{\partial {\mathcal D}^+} H d\sigma \right) \sum _{i=1}^{\infty } \frac{1}{i}, \end{aligned}$$
(A.9)

while volume and perimeter are bounded as follows.

$$\begin{aligned} |\Omega | = \left( 80^n + |{\mathcal D}| \sum _{i=1}^{\infty } r_i^{n} \right) , \quad \text {Per}(\Omega ) = \left( 80^{n-1} n - \sum _{i=1}^{\infty } w_{n-1} r_i^{n-1} + |\partial ^+ {\mathcal D}| \sum _{i=1}^{\infty } r_i^{n-1} \right) . \end{aligned}$$
(A.10)

Lemma A.3

There exists \(r>0\) such that \(\Omega \) given in Example A.2 is star-shaped with respect to \(B_r\).

Proof

As \(\Omega _1\) given in (A.6) is star-shaped, it is enough to consider a point in \(\Omega _2\). Note that \(\Gamma _2\) given in (A.7) is smooth. From Lemma C.1, it is enough to show that

$$\begin{aligned} x \cdot \mathbf {n}_x \ge r \end{aligned}$$
(A.11)

for some \(r>0\) and all \(x \in \Gamma _2\). Denote \(x = (x^1,x') \in {\mathbb R}^n\). For \(x \in r_i \partial {\mathcal D}^+ + (40,x_i)\), there exists \(y_i \in B_1^{n-1}\) such that

$$\begin{aligned} x = \left( r_i \phi (y_i), r_i y_i\right) + (40,x_i) \end{aligned}$$
(A.12)

and thus

$$\begin{aligned} \mathbf {n}_x = \frac{ (1, - D\phi (y_i))}{\sqrt{ 1 + |D\phi (y_i)|^2}}. \end{aligned}$$
(A.13)

From (A.3) and (A.5), it holds that

$$\begin{aligned} x \cdot \mathbf {n}_x \ge \frac{40}{\sqrt{ 1 + |D\phi (y_i)|^2}} - \frac{|D\phi (y_i) | (| r_i y_i| + |x_i|) }{\sqrt{ 1 + |D\phi (y_i)|^2}} \ge 10. \end{aligned}$$
(A.14)

\(\square \)

1.2 A.2: A non-convex set satisfying the \(\rho \textit{-reflection}\) property

Example A.4

Consider a non-convex set \(O_r\) for \(r> 2\rho \) defined by

$$\begin{aligned} O_r := (B_r(0) + \rho e_1) \cup (B_r(0) - \rho e_1) \end{aligned}$$
(A.15)

where \(e_1\) denotes the first unit vector. Then, \(O_r\) has \(\rho \textit{-reflection}\) for all \(r>2\rho \) (see Lemma A.5). In particular, we can find a non-convex set with unit volume satisfying the assumption (1.5) in Theorem 1 (see Lemma A.6).

Lemma A.5

For \(r> 2\rho \), \(O_r\) given in (A.15) has \(\rho \textit{-reflection}\).

Proof

First of all, as \(r> 2\rho \), we have \(\overline{B_\rho (0)} \subset B_r(0) \pm \rho e_1\) and thus \(O_r\) satisfies the first condition of \(\rho \textit{-reflection}\) in Definition 2.1.

Note that a ball \(B_r(x_0)\) satisfies (2.1) if the half space \(\Pi ^-_{\nu }(s)\) contains \(x_0\). For \(\nu \in S^{n-1}\) and all \(s>\rho \), \(\Pi ^-_{\nu }(s)\) contains \(\pm \rho e_1\) and \(\pm \rho e_1\) are the centers of balls \(B_r(0) \pm \rho e_1\). Using this fact and the aforementioned reflection property of the balls, we have

$$\begin{aligned} \Psi _{\Pi _\nu (s)} ((B_r(0) + \rho e_1) \cap \Pi ^+_{\nu }(s)) \subset (B_r(0) + \rho e_1) \cap \Pi ^-_{\nu }(s) \end{aligned}$$

and

$$\begin{aligned} \Psi _{\Pi _\nu (s)} ((B_r(0) - \rho e_1) \cap \Pi ^+_{\nu }(s)) \subset (B_r(0) - \rho e_1) \cap \Pi ^-_{\nu }(s). \end{aligned}$$

Therefore, \(O_r\) satisfies the second condition of \(\rho \textit{-reflection}\) in Definition 2.1. \(\square \)

Lemma A.6

For all \(\rho \in [0, (5 c_n)^{-1})\), there exists \(r > 2\rho \) such that \(|O_r| = 1\). Here, \(c_n = |B_1(0)|^{1/n}\).

Proof

Note that \(r \mapsto |O_r|\) is a increasing function in \(r>2\rho \). Also, as \(|O_r| > |B_r(0)|\), \(|O_r|\) goes to infinity as \(r \rightarrow \infty \). Furthermore, it holds that

$$\begin{aligned} |O_{2\rho } | \le 2 |B_{2\rho }(0)| = 2 (2\rho c_n)^n \le \frac{2^{n+1}}{5^n} < 1 \end{aligned}$$
(A.16)

and thus we conclude. \(\square \)

Appendix B: Proof of Proposition 3.5

In this section, we prove Proposition 3.5. First, in Lemma B.3, we show short-time star-shapedness based on the Hölder continuity in Lemma B.1. The remaining arguments are parallel to Theorem 6.5 in [20]. For simplicity, we fix \(r_0\) and \(R_0\) given in (3.11) and \(\delta \in (0,\delta _0)\) for \(\delta _0\) given in (3.4). Also, let \({\Omega _t}\) be a viscosity solution of \(V = -H + \gamma _{\delta }(|\Xi _t|)\) where \(\Xi _t\) is a flat flow given in Definition B.4.

First, let us recall several properties of solutions from [20].

Lemma B.1

[20, Corollary 3.7] Assume that \(\Omega _{0} \in S_{r,R}\). Then, there exists

$$\begin{aligned} {\mathcal M}_1={\mathcal M}_1(r,R,\Vert \gamma _{\delta }(|\Xi _t|)\Vert _{L^\infty ([0,T])}) \end{aligned}$$

such that we have

$$\begin{aligned} \sup _{x \in {\partial \Omega _t}} d(x, {\partial \Omega _0}) \le {\mathcal M}_1|t|^{\frac{1}{2}} \hbox { for } t \in [0,T]. \end{aligned}$$
(B.1)

Theorem B.2

[20, Theorem 3.6] Let \(I = [0,t_0)\) be the maximal interval satisfying \(\overline{B_\rho } \subset \Omega _t\). Then, \(\Omega _t\) has \(\rho \textit{-reflection}\) in I.

Lemma B.1 and Theorem B.2 imply the following lemma.

Lemma B.3

(Short-time star-shapedness) For \(r>r_0>0\) and \(0<R<R_0\), suppose that \({\overline{B}}_{(1+\beta )\rho } \subset \Omega _0\) and \({\Omega _0}\in S_{r,R}\) for \(r = \rho (\beta ^2 + 2\beta )\). Then, for all \(t \in [0,t_1]\), it holds that for some \({\hat{r}} > r_0\) and \({\hat{R}} < R_0\)

$$\begin{aligned} {\Omega _t}\in S_{{\hat{r}},{\hat{R}}}. \end{aligned}$$
(B.2)

where

$$\begin{aligned} t_1=t_1(r,R,\Vert \gamma _{\delta }(|\Xi _t|)\Vert _{L^\infty ([0,T])}) := \frac{1}{2} \left( \min \left\{ \frac{ \sqrt{r^2 + \rho ^2} - \sqrt{r_0^2 + \rho ^2}}{{\mathcal M}_1}, \frac{R_0 - R}{{\mathcal M}_1} \right\} \right) ^2. \end{aligned}$$
(B.3)

Here, \({\mathcal M}_1\) is given in Lemma B.1.

Proof

From Lemma B.1, it holds that in \([0,t_1]\)

$$\begin{aligned} {\overline{B}}_{(1+\beta )\rho - {\mathcal M}_1 t_1^{\frac{1}{2}}} \subset \Omega _t \subset B_{R + {\mathcal M}_1 t_1^{\frac{1}{2}}}. \end{aligned}$$
(B.4)

Theorem B.2 implies that \(\Omega _t\) has \(\rho \textit{-reflection}\) for all \(t \in [0,t_1]\)

By (B.3), it holds that

$$\begin{aligned} R + {\mathcal M}_1 t_1^{\frac{1}{2}}< R_0 \text{ and } r_0 < \left( ((1+\beta )\rho - {\mathcal M}_1 t_1^{\frac{1}{2}})^2 - \rho ^2 \right) ^{\frac{1}{2}}. \end{aligned}$$
(B.5)

From (C.9), we conclude. \(\square \)

Recall definition of flat flows and comparison principle from [20].

Definition B.4

[20, Definition 5.2] Let \((\Xi _t)_{t \ge 0}\) be a flat flow of (1.3) if there exists a sequence \(h_k \rightarrow 0\) such that

$$\begin{aligned} d_H(\Xi _t, E_{t}(h_k)) \rightarrow 0. \end{aligned}$$

Here, \(E_t=E_t(h)\) is given in (3.12).

Lemma B.5

[20, Proposition 6.1] Suppose that \({\Omega _t}\in S_{r,R}\) in [0, T] for some \(r>r_0\) and \(R<R_0\).

$$\begin{aligned} \hbox { If } \Omega _0 \subset \subset \Xi _0, \hbox { then } \Omega _t \subset \subset \Xi _t \hbox { in } [0,T]. \hbox { Also, if } \Xi _0 \subset \subset \Omega _0, \hbox { then }\Xi _t \subset \subset \Omega _t \hbox { in } [0,T]. \end{aligned}$$
(B.6)

Proof of Proposition 3.5

Let \({\Omega _t}^{\varepsilon ,+}\) and \({\Omega _t}^{\varepsilon ,-}\) be viscosity solutions of \(V = -H + \gamma _{\delta }(|\Xi _t|)\) starting from \({\Omega _0}^{\varepsilon ,-} := (1-\varepsilon ){\Omega _0}\) and \({\Omega _0}^{\varepsilon ,+}:= (1+\varepsilon ){\Omega _0}\), respectively. Note that

$$\begin{aligned} {\overline{B}}_{(1+\beta _1)\rho } \subset {\Omega _0}\in S_{r_1,R_1} \end{aligned}$$
(B.7)

where \(r_1 (= \rho (\beta _1^2 + 2\beta _1)^{\frac{1}{2}})\) and \(\beta _1\) are given in (3.7) and \(R_1\) is given in (3.8). From (B.7), there exist \(\varepsilon _0>0\), \(\beta \in (0, \beta _1) \) and \(R \in (R_1,R_0)\) such that

$$\begin{aligned} {\overline{B}}_{(1+\beta )\rho } \subset {\Omega _0}^{\varepsilon ,\pm } \in S_{r,R} \hbox { for all } \varepsilon \in [0,\varepsilon _0], \hbox { and } r := \rho (\beta ^2 + 2\beta )^{\frac{1}{2}}>r_0. \end{aligned}$$
(B.8)

Here, \(r_0\) and \(R_0\) is given in (3.11). \(\square \)

Let us show that \(\Xi _t = {\Omega _t}\) in \([0,t_1]\) for \(t_1=t_1(r,R,K_1)\) given in (B.3) and \(K_1 := \frac{1}{\delta }\max \{ 1, |B_{R_0}|\}\). As \(|\gamma _{\delta }(|\Xi _t|) | \le K_1\), we can apply Lemma B.3 combining with (B.8) and conclude that for some \({\hat{r}} > r_0\) and \({\hat{R}} < R_0\)

$$\begin{aligned} {\Omega _t}^{\varepsilon ,\pm } \in S_{{\hat{r}},{\hat{R}}} \hbox { in } [0,t_1]. \end{aligned}$$
(B.9)

By Lemma B.5, we conclude that

$$\begin{aligned} {\Omega _t}^{\varepsilon ,-} \subset \Xi _t \subset {\Omega _t}^{\varepsilon ,+}. \end{aligned}$$
(B.10)

Note that \({\Omega _t}^{\varepsilon ,+}\) and \({\Omega _t}^{\varepsilon ,-}\) converges to \({\Omega _t}\) as \(\varepsilon \rightarrow 0\) from the uniqueness in [20, Theorem 4.3]. We conclude that \(\Xi _t = {\Omega _t}\) in \([0,t_1]\). As the proof of Theorem 6.5 in [20], we can iterate this step to conclude that \(\Xi _t = {\Omega _t}\) for all \(t\ge 0\). Lastly, from Proposition 3.3, \({({\Omega ^\delta _t})_{t\ge 0}}\) is uniquely determined by its initial data. Now we conclude (3.13). \(\Box \)

Appendix C: Geometric properties

In this section, we consider geometric properties of \(\rho \textit{-reflection}\) and \(S_{r}\). First, let us recall a local property of \(S_{r}\) from [20].

Lemma C.1

[20, Lemma 3.2] For a continuously differentiable and bounded function \(\phi : {\mathbb {R}}^n \rightarrow {\mathbb {R}}\), let us denote the positive set of \(\phi \) by \(\Omega (\phi )\). Let us assume that \(\Omega (\phi )\) contains \(B_r(0)\) and \(D\phi \ne 0\) on \(\partial \Omega (\phi )\). Then the set \(\Omega (\phi )\) is in \(S_r\) if and only if

$$\begin{aligned} x \cdot \mathbf {n}_x = x \cdot \left( - \frac{D\phi }{|D\phi |}(x)\right) \ge r \hbox { for all } x\in \partial \Omega (\phi ), \end{aligned}$$

where \(\mathbf {n}_x\) denotes the outward normal of \(\partial \Omega (\phi )\) at x.

Here are several properties of \(\rho \textit{-reflection}\) and \(S_{r}\) from [15].

Lemma C.2

[15, Lemma 20] Suppose that \(\Omega \) has \(\rho \textit{-reflection}\). Then, we have

$$\begin{aligned} \sup _{x \in \partial \Omega } |x| - \inf _{x \in \partial \Omega } |x| \le 4\rho . \end{aligned}$$

Lemma 4.3 can be shown by Lemma C.3

Lemma C.3

[15, Lemma 3] For a bounded domain \(\Omega \) containing \(B_r(0)\), the followings are equivalent:

  1. (1)

    \(\Omega \in S_r\).

  2. (2)

    For all \(x\in \partial \Omega \), there is an interior cone to \(\Omega \):

    $$\begin{aligned} IC(x,r):=\left( (x+C(-x,\theta _{x,r})) \cap C(x,\frac{\pi }{2} - \theta _x ) \right) \cup B_r(0) \subset \Omega \end{aligned}$$
    (C.1)

    where where \(C(x,\theta )\) is a cone in the direction x with opening angle \(\theta \) for \(x \in {\mathbb R}^n\) and \(\theta \in [0, \pi ]\),

    $$\begin{aligned} C(x,\theta ):=\{ y \mid \langle x,y \rangle > \cos \theta |x| |y| \} \hbox { and }\theta _{x,r} := \arcsin \frac{r}{|x|} \in \left[ 0,\frac{\pi }{2}\right] . \end{aligned}$$
  3. (3)

    For all \(x\in \partial \Omega \), there is an exterior cone to \(\Omega \):

    $$\begin{aligned} EC(x,r):=\left( x+C(x,\theta _{x,r})\right) \cap B_\epsilon (x) \subset \Omega ^c \text { where } \theta _{x,r} := \arcsin \frac{r}{|x|} \in \left[ 0,\frac{\pi }{2}\right] . \end{aligned}$$
    (C.2)

Proof of Lemma 4.3

From Lemma C.3, it holds that for all \(x \in \partial E\),

$$\begin{aligned} IC(x,r) \subset E \hbox { and } EC(x,r) \subset E^c \end{aligned}$$
(C.3)

where IC is an interior cone given in (C.1), and EC is an exterior cone given in (C.2). Note that as \(|x| \le R\), the angle of both the interior cone and exterior cone, \(\theta _x\), is bounded from below as follows,

$$\begin{aligned} \theta _x := \arcsin \frac{r}{|x|} \ge \arcsin \frac{r}{R}. \end{aligned}$$
(C.4)

Thus, for \(\eta _1(r,R):= |IC(Re_1,r) \cap B_\varepsilon (Re_1)|\), it holds that for \(\varepsilon \in (0,r)\)

$$\begin{aligned} \eta _1 \varepsilon ^{n} \le |IC(x,r) \cap B_\varepsilon (x)| \le |E \cap B_\varepsilon (x)|. \end{aligned}$$
(C.5)

Here, \(e_1\) is a unit vector in the positive \(x_1\) direction. Similarly, it holds that

$$\begin{aligned} |B_\varepsilon (x) {\setminus } E| \ge |B_\varepsilon (x) \cap EC(x,r)| \ge \eta _1 \varepsilon ^{n}. \end{aligned}$$
(C.6)

As \(E \in S_{r,R}\), there exists \(\varepsilon _0 = \varepsilon (r,R) < r\) such that for all \(\varepsilon \in (0,\varepsilon _0)\)

$$\begin{aligned} B_\varepsilon (x) \cap \partial E = (U, f(U)) \end{aligned}$$
(C.7)

up to rotation for some Lipschitz function \(f = f_{x,\varepsilon } : U \subset B_\varepsilon ^{n-1}(x) \rightarrow {\mathbb R}\). Note that as \(E \in S_{r,R}\), the Lipschitz constant of f is uniformly bounded by some constant \({\mathcal L}= {\mathcal L}(r,R)\).

From Theorem 9.1 in [22],

$$\begin{aligned} {\mathcal H}^{n-1}(B_\varepsilon (x) \cap \partial E) = \int _{U} \sqrt{ 1+ | \nabla f |^2 } dx \le |U| \sqrt{ 1+ {\mathcal L}^2} \le n w_n \varepsilon ^{n-1} \sqrt{ 1+ {\mathcal L}^2}. \end{aligned}$$
(C.8)

Thus, (4.3) holds with \(\eta _2(r,R):=n w_n \sqrt{ 1+ {\mathcal L}^2}\). Here, \(w_n\) is a volume of a unit ball in \({\mathbb R}^n\). On the other hand, from the isoperimetric inequality in [22, Proposition 12.37] and (4.2), we get the lower bound of (4.3). \(\square \)

Lemma C.4

[15, Lemma 21] Suppose that \(\Omega \) has \(\rho \textit{-reflection}\). Moreover, \(\Omega \in S_r\) with

$$\begin{aligned} r=(\inf _{x\in \partial \Omega } |x|^2 - \rho ^2 )^{1/2}. \end{aligned}$$
(C.9)

For \(E, F \subset {\mathbb R}^n\), define the Hausdorff distance by

$$\begin{aligned} d_H(E,F) := \max \left\{ \sup _{x \in E} d(x,F), \sup _{x \in F} d(x,E) \right\} . \end{aligned}$$
(C.10)

We say that \(A \lesssim _{r,R} B\) if there exists a constant \(C=C(r,R)>0\) depending on rR such that \(A \le CB\).

Lemma C.5

[15, Lemmas 23, 24] Consider sets \(\Omega _1\) and \(\Omega _2\) in \(S_{r,R}\) for \(R>r>0\). Then the following holds:

$$\begin{aligned} d_H(\partial \Omega _1, \partial \Omega _2) \lesssim _{r,R} d_H(\Omega _1, \Omega _2), \quad |\Omega _1 \Delta \Omega _2| \lesssim _{r,R} d_H(\Omega _1, \Omega _2). \end{aligned}$$

Lemma C.6

[15, Lemma 24] The metric space \((\partial S_{r,R}, d_H)\) is compact:

  1. (1)

    Suppose that \(\Gamma _j \in (\partial S_{r,R}, d_H)\) for some \(r,R>0\) and all \(j \in {\mathbb N}\). Then \(\{\Gamma _j\}_{j \in {\mathbb N}}\) has a subsequence that converges and any subsequential limit is also in \(\partial S_{r,R}\).

  2. (2)

    Let I be a compact interval in \({\mathbb R}\) and \(\Gamma _j : I \rightarrow \partial S_{r,R}\) for \(j \in {\mathbb N}\) is an equicontinuous sequence of paths in \((\partial S_{r,R}, d_H)\). Then, there is a subsequence of the \(\Gamma _j(\cdot )\) that converges uniformly on I on a path \(\Gamma : I \rightarrow (\partial S_{r,R}, d_H)\).

Lastly, let us show the following property of characteristic functions.

Lemma C.7

Let \(\{ {({\Omega ^{k}_{t}})_{t\ge 0}}\}_{k \in {\mathbb N}}\) be a sequence of sets in \(S_{r,R}\) for \(0<r<R\). Suppose that \({\Omega ^{k}_{t}}\) converges locally uniformly to \({\Omega ^\infty _t}\) on \({[0, \infty )}\). For a sequence of functions \(\{ u_k \}_{k \in {\mathbb N}\cup \{ + \infty \} }\) defined by

$$\begin{aligned} u_k := \chi _{{\Omega ^{k}_{t}}} - \chi _{({\Omega ^{k}_{t}})^C} \hbox { for } k \in {\mathbb N}\cup \{ + \infty \}, \end{aligned}$$
(C.11)

it holds that

$$\begin{aligned} u_\infty ^* = \mathop {\lim \,\sup {}^*}\limits _{k \rightarrow \infty }u_k \hbox { and } (u_\infty )_* = \mathop {\lim \,\inf {}_*}\limits _{k \rightarrow \infty } u_k. \end{aligned}$$
(C.12)

Here, \(\mathop {\lim \,\sup {}^*}\limits \) and \(\mathop {\lim \,\inf {}_*}\limits \) are given in (2.4).

Proof

Let us show the first equation in (C.12) only. The second one can be shown by the parallel arguments.

By uniform convergence of \({\Omega ^{k}_{t}}\) in a finite interval, for any \(j \in {\mathbb N}\), there exists \(k_1>0\) such that for all \(k>k_1\)

$$\begin{aligned} d_H({\Omega ^{k}_{t}}, {\Omega ^\infty _t}) < \frac{1}{j}. \end{aligned}$$
(C.13)

Thus, for any \(x \in {\Omega ^\infty _t}\) and \(k>k_1\), there exists \(y \in {\Omega ^{k}_{t}}\) such that \(|x-y| < \frac{1}{j}\). Thus, we conclude that

$$\begin{aligned} \mathop {\lim \,\sup {}^*}\limits _{k \rightarrow \infty }u_k(x,t) = \lim \limits _{j \rightarrow \infty } \sup \left\{ u_k(y,s) : k \ge j,\quad |y-x| \le \frac{1}{j}, \quad |s-t| \le \frac{1}{j} \right\} = 1 \end{aligned}$$
(C.14)

and \(u_\infty ^*(x) = \mathop {\lim \,\sup {}^*}\limits _{k \rightarrow \infty }u_k(x)\) for \(x \in {\Omega ^\infty _t}\).

Note that we have for any sets \(\Omega _1, \Omega _2 \in S_{r,R}\)

$$\begin{aligned} d_H(\Omega _1^C, \Omega _2^C) \le d_H(\partial \Omega _1, \partial \Omega _2). \end{aligned}$$
(C.15)

Combining this with Lemma C.5, we conclude that \(({\Omega ^{k}_{t}})^C\) converges locally uniformly to \(({\Omega ^\infty _t})^C\). By parallel arguments, for any \(x \in ({\Omega ^\infty _t})^C\), we conclude that \(\mathop {\lim \,\sup {}^*}\limits _{k \rightarrow \infty }u_k(x,t) = -1\). As \(\mathop {\lim \,\sup {}^*}\limits _{k \rightarrow \infty }u_k\) is upper semicontinuous, we conclude (C.12). \(\square \)

Lemma C.8

For any function \(u : Q \rightarrow {\mathbb R}\) and \(\Theta \in C({[0, \infty )})\), it holds that

$$\begin{aligned} \widehat{u^*}(\cdot ;\Theta ) = {\widehat{u}}(\cdot ;\Theta )^* \end{aligned}$$
(C.16)

and

$$\begin{aligned} \widetilde{u_*}(\cdot ;\Theta ) = {\widetilde{u}}(\cdot ;\Theta )_*. \end{aligned}$$
(C.17)

Proof

Let us only show (C.16). The parallel arguments imply (C.17).

Let us assume that both sides are finite at \((x_0,t_0) \in Q\). We claim that

$$\begin{aligned} \widehat{u^*}(x_0,t_0;\Theta ) \le {\widehat{u}}(x_0,t_0;\Theta )^*. \end{aligned}$$
(C.18)

By the upper semicontinuity of \(({\widehat{u}})^*\), for \(\varepsilon >0\) there exists \(\delta \in (0, t_0)\) such that

$$\begin{aligned} {\widehat{u}}(\cdot , \cdot ;\Theta ) < {\widehat{u}}(x_0,t_0;\Theta )^* + \varepsilon \hbox { in } B_{\delta }(x_0) \times (t_0 - \delta , t_0 + \delta ). \end{aligned}$$
(C.19)

From the definition of \({\widehat{u}}\) in (2.14), it holds that

$$\begin{aligned} u(\cdot , \cdot )< {\widehat{u}}(x_0,t_0;\Theta )^* + \varepsilon \hbox { in } {\mathcal N}_1:= \{ (x,t) \in Q : |t - t_0|< \delta \hbox { and } |x - x_0 | < \delta + \Theta (t)\}. \end{aligned}$$
(C.20)

Furthermore, by the continuity of \(\Theta \), for any \(y \in {\overline{B}}_{\Theta (t_0)}(x_0)\) there exists a small neighborhood \({\mathcal N}_2\) of \((y,t_0)\) in Q such that \({\mathcal N}_2 \subset {\mathcal N}_1\). From (C.20), we have

$$\begin{aligned} u(y,t_0)^* \le {\widehat{u}}(x_0,t_0;\Theta )^* + \varepsilon \hbox { for all } y \in {\overline{B}}_{\Theta (t_0)}(x_0). \end{aligned}$$
(C.21)

As \(\varepsilon \) is arbitrary, we conclude (C.18).

Next, let us show the opposite inequality,

$$\begin{aligned} \widehat{u^*}(x_0,t_0;\Theta ) \ge {\widehat{u}}(x_0,t_0;\Theta )^*. \end{aligned}$$
(C.22)

For any \(\varepsilon >0\), let us show that there exists \(\delta \in (0, t_0)\) such that

$$\begin{aligned} \widehat{u^*}(x_0,t_0;\Theta ) > u(\cdot , \cdot ) - \varepsilon \hbox { in } B_{\delta + \Theta (t_0)}(x_0) \times (t_0 - \delta , t_0 + \delta ). \end{aligned}$$
(C.23)

If not, then there exists \(\{ (y_k,s_k) \}_{k \in {\mathbb N}} \subset Q\) such that

$$\begin{aligned} |s_k - t_0 | \le \frac{1}{k}, \ \ |y_k - x_0| < \frac{1}{k} + \Theta (t_0) \hbox { and } \widehat{u^*}(x_0,t_0) \le u(y_k,s_k) - \varepsilon . \end{aligned}$$
(C.24)

Then, \(\{ s_k \}_{k \in {\mathbb N}}\) converges to \(t_0\). Also, by compactness of \({\overline{B}}_{\Theta (t_0)+1}(x_0)\), there exists a subsequence \(\{ k_i \}_{i \in {\mathbb N}}\) and \(y^* \in {\overline{B}}_{\Theta (t_0)+1}(x_0)\) such that \(\{ y_{k_i} \}_{ i \in {\mathbb N}}\) converges to \(y^*\). Thus, it holds that

$$\begin{aligned} \widehat{u^*}(x_0,t_0) \le \mathop {\lim \,\sup }\limits _{k \rightarrow \infty } u(y_k,s_k) - \varepsilon \le u^*(y^*,t_0) - \varepsilon . \end{aligned}$$
(C.25)

On the other hand, (C.24) implies \(y^* \in {\overline{B}}_{\Theta (t_0)}(x_0)\), which contradicts to (C.25). As a consequence, we get (C.23).

By (C.23) and the continuity of \(\Theta \), there exists \(\delta _1 \in (0, \delta )\) such that

$$\begin{aligned} \widehat{u^*}(x_0,t_0;\Theta ) > u(\cdot , \cdot ) - \varepsilon \hbox { in } \{(x,t) \in Q : |t - t_0|< \delta _1 \hbox { and } |x - x_0 | < \delta _1 + \Theta (t) \}. \end{aligned}$$
(C.26)

As \(\varepsilon \) is arbitrary, we get (C.22) and conclude (C.16).

Similar arguments can be applied if either the left hand side or right hand side in (C.16) is infinity at \((x_0,t_0) \in Q\). In particular, for any \(\varepsilon \in (0, t_0)\), there exists a sequence \(\{ x_k, t_k \}_{ k \in {\mathbb N}}\) in \((t_0 - \varepsilon , t_0 + \varepsilon ) \times B_{\Theta (t_0) + \varepsilon }(x_0)\) such that \(u(x_k, t_k)\) converging to infinity. This implies that the other side is infinity. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, I., Kwon, D. Volume preserving mean curvature flow for star-shaped sets. Calc. Var. 59, 81 (2020). https://doi.org/10.1007/s00526-020-01738-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-020-01738-0

Mathematics Subject Classification

Navigation