
We study a variational problem for hypersurfaces in a wedge in the Euclidean space. Our wedge is bounded by a finitely many hyperplanes passing a common point. The total energy of each hypersurface is the sum of its anisotropic surface energy and the wetting energy of the planar domain bounded by the boundary of the considered hypersurface. An anisotropic surface energy is a generalization of the surface area which was introduced to model the surface tension of a small crystal. We show an existence and uniqueness result of local minimizers of the total energy among hypersurfaces enclosing the same volume. Our result is new even when the special case where the surface energy is the surface area.
Citation: Miyuki Koiso. Stable anisotropic capillary hypersurfaces in a wedge[J]. Mathematics in Engineering, 2023, 5(2): 1-22. doi: 10.3934/mine.2023029
[1] | Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic $ p $-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015 |
[2] | Joan Mateu, Maria Giovanna Mora, Luca Rondi, Lucia Scardia, Joan Verdera . A maximum-principle approach to the minimisation of a nonlocal dislocation energy. Mathematics in Engineering, 2020, 2(2): 253-263. doi: 10.3934/mine.2020012 |
[3] | Bruno Bianchini, Giulio Colombo, Marco Magliaro, Luciano Mari, Patrizia Pucci, Marco Rigoli . Recent rigidity results for graphs with prescribed mean curvature. Mathematics in Engineering, 2021, 3(5): 1-48. doi: 10.3934/mine.2021039 |
[4] | Chiara Gavioli, Pavel Krejčí . Deformable porous media with degenerate hysteresis in gravity field. Mathematics in Engineering, 2025, 7(1): 35-60. doi: 10.3934/mine.2025003 |
[5] | YanYan Li . Symmetry of hypersurfaces and the Hopf Lemma. Mathematics in Engineering, 2023, 5(5): 1-9. doi: 10.3934/mine.2023084 |
[6] | Antonio De Rosa, Luca Lussardi . On the anisotropic Kirchhoff-Plateau problem. Mathematics in Engineering, 2022, 4(2): 1-13. doi: 10.3934/mine.2022011 |
[7] | Stefano Berrone, Stefano Scialò, Gioana Teora . The mixed virtual element discretization for highly-anisotropic problems: the role of the boundary degrees of freedom. Mathematics in Engineering, 2023, 5(6): 1-32. doi: 10.3934/mine.2023099 |
[8] | Giacomo Canevari, Arghir Zarnescu . Polydispersity and surface energy strength in nematic colloids. Mathematics in Engineering, 2020, 2(2): 290-312. doi: 10.3934/mine.2020015 |
[9] | Barbara Brandolini, Florica C. Cîrstea . Anisotropic elliptic equations with gradient-dependent lower order terms and $ L^1 $ data. Mathematics in Engineering, 2023, 5(4): 1-33. doi: 10.3934/mine.2023073 |
[10] | Matteo Novaga, Marco Pozzetta . Connected surfaces with boundary minimizing the Willmore energy. Mathematics in Engineering, 2020, 2(3): 527-556. doi: 10.3934/mine.2020024 |
We study a variational problem for hypersurfaces in a wedge in the Euclidean space. Our wedge is bounded by a finitely many hyperplanes passing a common point. The total energy of each hypersurface is the sum of its anisotropic surface energy and the wetting energy of the planar domain bounded by the boundary of the considered hypersurface. An anisotropic surface energy is a generalization of the surface area which was introduced to model the surface tension of a small crystal. We show an existence and uniqueness result of local minimizers of the total energy among hypersurfaces enclosing the same volume. Our result is new even when the special case where the surface energy is the surface area.
We prove an existence and uniqueness result of stable equilibrium hypersurfaces in wedge-like domains in the (n+1)-dimensional Euclidean space Rn+1 for anisotropic surface energy, which serve as a mathematical model of small crystals and small liquid crystals with anisotropy. Our result is new even when the special case where the surface energy is merely the surface area.
Let γ:Sn→R>0 be a positive continuous function on the unit sphere Sn={ν∈Rn+1;‖ν‖=1} in Rn+1. Let X be a closed piecewise-smooth hypersurface in Rn+1. X will be represented as a mapping X:M→Rn+1 from an n-dimensional oriented connected compact C∞ manifold M into Rn+1. Let ν be the unit normal vector field along X|M∖S[X], where S[X] is the set of singular points of X. The anisotropic energy Fγ(X) of X is defined as Fγ(X):=∫M∖S[X]γ(ν)dA, where dA is the n-dimensional volume form of M induced by X. Such an energy was introduced by Gibbs (1839–1903) in order to model the shape of small crystals, and it is used as a mathematical model of anisotropic surface energy [19,20]. It is known that, for any positive number V>0, among all closed piecewise-smooth hypersurfaces as above enclosing the same (n+1)-dimensional volume V, there exists a unique (up to translation in Rn+1) minimizer Wγ(V) of Fγ [17]. Each Wγ(V) is homothetic to the so-called Wulff shape for γ (the definition of the Wulff shape will be given in §2), which we will denote by Wγ. When γ≡1, Fγ(X) is the usual n-dimensional volume of the hypersurface X and Wγ is the unit sphere Sn.
The Wulff shape Wγ is not smooth in general. However, in this paper we assume that Wγ is a smooth strictly convex hypersurface like the previous works that studied variational problems of anisotropic surface energies in differential geometry (cf. [1,4,5,8,9,10,11,12,14,15]).
Each equilibrium hypersurface X of Fγ for variations that preserve the enclosed (n+1)-dimensional volume (we will call such a variation a volume-preserving variation) has constant anisotropic mean curvature. Here, the anisotropic mean curvature Λ of a piecewise-C2 hypersurface X is defined at each regular point of X as (cf. [8,15]) Λ:=(1/n)(−divMDγ+nHγ), where Dγ is the gradient of γ and H is the mean curvature of X. If γ≡1, Λ=H holds.
Let Ω be a wedge-shaped domain in Rn+1 bounded by k hyperplanes Π1,⋯,Πk (k≤n+1) such that the intersection Π1∩⋯∩Πk includes the origin 0 of Rn+1 (Figure 1). Denote by ˜Nj the unit normal to Πj which points outward from Ω. We assume that ˜N1,⋯,˜Nk are linearly independent. We call each Πi∩Πj (i≠j) an edge of Ω. Let ωj (j=1,⋯,k) be non-negative constants. Let M be an n-dimensional oriented connected compact C∞ manifold with boundary ∂M=σ1∪⋯∪σk, where each σj is topologically Sn−1. Consider any C∞-immersion X:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) of which the restriction X|∂M to ∂M is an embedding. Set Cj=X(σj), and let Dj=Dj(X)⊂Πj be the n-dimensional domain bounded by Cj (j=1,⋯,k). We assume that the unit normal ν to X points outward from the (n+1)-dimensional domain bounded by X(M)∪(∪kj=1Dj) near each Cj. We define the wetting energy W(X) of X as
W(X)=k∑j=1ωjHn(Dj), |
where Hn(Dj) is the n-dimensional Hausdorff measure of Dj. Then, we define the total energy E(X)=Eγ(X) of X by
E(X)=Fγ(X)+W(X). |
Note that X(M)∪D1∪⋯∪Dk is an oriented closed piecewise smooth hypersurface without boundary (possibly with self-intersection). We denote by V(X) the oriented volume enclosed by X(M)∪D1∪⋯∪Dk which is represented as
V(X)=1n+1∫M⟨X,ν⟩dA. |
We call a critical point of E for volume-preserving variations an anisotropic capillary hypersurface (or, simply, a capillary hypersurface). A capillary hypersurface is said to be stable if the second variation of E is nonnegative for all volume-preserving variations of X. Otherwise, it is said to be unstable. In this paper, we prove the following two results on the uniqueness and the existence of stable capillary hypersurfaces.
Theorem 1. Let X:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) be a compact oriented immersed anisotropic capillary hypersurface that is disjoint from any edge of Ω, having embedded boundary and satisfying X(∂M)∩Πj=∂Dj for a nonempty bounded domain Dj in Πj. If X is stable (and all D1,⋯,Dk are convex if n≥3), then X(M) is (up to translation and homothety) a part of the Wulff shape Wγ. Conversely, if X is an embedding onto a part of Wγ (up to translation and homothety), then it is stable.
Theorem 2. There exists an anisotropic capillary hypersurface X that is a part of the Wulff shape (up to translation and homothety) and that intersects Πj with more than two points if and only if ωj<γ(˜Nj) holds.
As for previous works which are closely related to Theorem 1, we have the followings. First, Theorem 1 is a generalization of the main result of [3], where we proved the uniqueness result similar to Theorem 1 for isotropic capillary hypersurfaces in a wedge in Rn+1 with k=2; here, isotropic means that γ≡1. [13] also studies the isotropic case for general k, but it does not prove that parts of the Wulff shape are stable. As for the anisotropic case, we studied the existence and uniqueness of stable anisotropic capillary surfaces between two parallel planes Π1 and Π2 in R3 [9,10,11]. There, stable anisotropic capillary surfaces are not necessarily parts of the Wulff shape (up to translation and homothety). This suggests that the assumption of the linear independence of ˜N1,⋯,˜Nk cannot be removed in our Theorem 1. Finally we mention that Theorem 1 was announced in [7]. Moreover there an outline of the proof of Theorem 1 for n=2 and k=2 was given.
This article is organized as follows. In Section 2 we give preliminary contents. We give the definition and a representation of the Wulff shape. We also give the definitions of various anisotropic curvatures for hypersurfaces. Moreover we recall the definition of anisotropic parallel hypersurfaces and a Steiner-type integral formula for these hypersurfaces. In Section 3, we give the first variation formulas and the Euler–Lagrange equations for our variational problems. Also the proof of Theorem 2 is given. Sections 4, 6 and 7 will be devoted to proving the uniqueness part of Theorem 1. The existence part that is the last statement of Theorem 1 will be proved in Section 5.
In this paper, we call the boundary Wγ of the convex set ˜W[γ]:=∩ν∈Sn{X∈Rn+1;⟨X,ν⟩≤γ(ν)} the Wulff shape for γ, where ⟨ ⟩ means the standard inner product in Rn+1. In other literatures, ˜W[γ] is often called the Wulff shape.
From now on, any parallel translation of the Wulff shape Wγ will be also called the Wulff shape, and it will be denoted also by Wγ, if it does not cause any confusion.
From now on, we assume that, for simplicity, γ is of class C∞. We also assume that the homogeneous extension ¯γ:Rn+1→R≥0 of γ that is defined as ¯γ(rX):=rγ(X) (∀X∈Sn, ∀r≥0) is a strictly convex function. In this case, we say that γ is strictly convex, which is equivalent to the n×n matrix D2γ+γ⋅In is positive definite at any point in Sn. Here, D2γ is the Hessian of γ on Sn and In is the identity matrix of size n. The Wulff shape Wγ is smooth and strictly convex (that is, each principal curvature of Wγ with respect to the inward-pointing normal is positive at each point of Wγ) if and only if γ is of class C2 and strictly convex.
The Cahn–Hoffman map ξγ:Sn→Rn+1 for γ is defined as ξγ(ν)=Dγ|ν+γ(ν)ν, (ν∈Sn). Here, the tangent space Tν(Sn) of Sn at ν is naturally identified with the hyperplane in Rn+1 which is tangent to Sn at ν. Because γ is strictly convex, ξγ gives an embedding onto Wγ. Moreover, the outward-pointing unit normal at a point ξγ(ν) to Wγ coincides with ν (cf. [8]).
The Cahn-Hoffman field ˜ξ along X for γ is defined as ˜ξ:=ξγ∘ν:M→Rn+1. Since the unit normal ν(p) of X at p∈M coincides with the unit normal of ξγ at the point ν(p), we can identify TpM with T˜ξ(p)ξγ(Sn).
The linear map Sγp:TpM→TpM (p∈M) given by the n×n matrix Sγ:=−d˜ξ is called the anisotropic shape operator of X. Various anisotropic curvatures of X are defined as follows.
Definition 1 (anisotropic curvatures; cf. [5,15]). (i) The eigenvalues of Sγ are called the anisotropic principal curvatures of X. We denote them by kγ1,⋯,kγn.
(ii) Let σγr be the elementary symmetric functions of kγ1,⋯,kγn:
σγr:=∑1≤l1<⋯<lr≤nkγl1⋯kγlr,r=1,⋯,n. | (2.1) |
Set σγ0:=1. Hγr:=σγr/nCr is called the rth anisotropic mean curvature of X, where nCr=n!k!(n−k)!.
(iii) Hγ1 is called also the anisotropic mean curvature of X, and we often denote it by Λ; that is, Λ=1nn∑i=1kγi=−1ntraceM(d˜ξ). $
As we mentioned above, for the Cahn-Hoffman map ξγ:Sn→Rn+1, it is shown that the unit normal vector field νξγ is given by ξ−1γ. Hence, the anisotropic shape operator of ξγ is Sγ=−d(ξγ∘νξγ)=−d(idSn)=−In. Therefore, the anisotropic principal curvatures of ξγ are −1, and hence, each rth anisotropic mean curvature of ξγ is (−1)r. Particularly, the anisotropic mean curvature of ξγ for the normal ν and that of Wγ for the outward-pointing unit normal is −1 at any point. More generally, the anisotropic mean curvature of an immersion X:M→Rn+1 is the mean value of the ratios of the principal curvatures of the Wulff shape and the corresponding curvatures of X which is explained as follows.
Remark 1 (cf. [8]). Let X:M→Rn+1 be an immersion. Take any point p∈M. We compute the anisotropic mean curvature Λ(p) of X at p. Let {e1,⋯,en} be a locally defined frame on Sn such that (D2γ+γ⋅In)ei=(1/μi)ei, where μi are the principal curvatures of ξγ with respect to ν. Note that the basis {e1,⋯,en} at ν(p) also serves as an orthogonal basis for the tangent hyperplane of X at p. Let (−wij) be the matrix representing dν with respect to this basis. Then
−Sγ=dξγ∘dν=(D2γ+γ⋅In)dν=(−w11/μ1⋯−w1n/μ1 ⋅ ⋯ ⋅ ⋅ ⋯ ⋅ ⋅ ⋯ ⋅−wn1/μn⋯−wnn/μn). |
This with the definition of Λ gives
Λ=−1ntraceM(d˜ξ)=(1/n)(w11/μ1+⋯+wnn/μn). | (2.2) |
Sγ is not symmetric in general. However, we have the following good properties of the anisotropic curvatures.
Remark 2. (i) If dξγ=D2γ+γ⋅In is positive definite at a point ν(p) (p∈M), then all of the anisotropic principal curvatures of X at p are real [4].
(ii) kγi is not a real value in general. However, each Hγr is always a real valued function on M [6].
At the end of this section, we recall a useful concept that is "anisotropic parallel hypersurface" and an important integral formula that is a generalization of the classical Steiner's formula. Anisotropic parallel hypersurface is a generalization of parallel hypersurface and is defined as follows.
Definition 2 (Anisotropic parallel hypersurface, cf. [15]). Let X:M→Rn+1 be an immersion. For any real number t, we call the map Xt:=X+t˜ξ:M→Rn+1 the anisotropic parallel deformation of X of height t. If Xt is an immersion, then we call it the anisotropic parallel hypersurface of X of height t.
The anisotropic energy Fγ(Xt) of the anisotropic parallel hypersurface Xt:=X+t˜ξ is a polynomial of t with degree at the most n as follows.
Fact 1 (Steiner-type formula [4]). Assume that γ:Sn→R>0 is of class C∞. Let X:M→Rn+1 be an immersion. Consider anisotropic parallel hypersurfaces Xt=X+t˜ξ:M→Rn+1. Then, the n-dimensional volume form dAXt and the anisotropic energy Fγ(Xt) of Xt have the following representations.
dAXt=(1−tkγ1)⋯(1−tkγn)dA=n∑r=0(−1)rtr(nCr)HγrdA, | (2.3) |
Fγ(Xt)=∫Mγ(ν)n∑r=0(−1)rtr(nCr)HγrdA. | (2.4) |
The isotropic version of Fact 1 is known as the Weyl's tube formula [18]. The isotropic 2-dimensional version is the well-known Steiner's formula.
In order to obtain the Euler–Lagrange equations for our capillary problem, first we recall the first variation formula for the anisotropic surface energy Fγ.
Proposition 1 ([6]). Let Xϵ:M→Rn+1 (ϵ∈J:=(−ϵ0,ϵ0)), be a smooth variation of X; that is, ϵ0>0 and X0=X. Set
δX:=∂Xϵ∂ϵ|ϵ=0,ψ:=⟨δX,ν⟩. |
Then, the first variation of the anisotropic energy Fγ is given as follows.
δFγ:=dFγ(Xϵ)dϵ|ϵ=0=−∫MnΛψdA−∮∂M⟨δX,R(p(˜ξ))⟩ds, | (3.1) |
where ds is the (n−1)-dimensional volume form of ∂M induced by X, N is the outward-pointing unit conormal along ∂M, R is the π/2-rotation on the (N,ν)-plane, and p is the projection from Rn+1 to the (N,ν)-plane.
On the other hand, the first variation of the (n+1)-dimensional volume enclosed by the region between X and Xϵ is given by ([2]) as
δV=∫M⟨δX,ν⟩dA. | (3.2) |
Similarly, the first variation of Hn(Dj) is given as follows.
δHn(Dj)=∫σj⟨δX,ρ⟩ds, | (3.3) |
where ρ is the outward-pointing unit normal along X|σj:σj→Πj, and ds is the (n−1)-dimensional volume form of X|σj.
The Eq (3.1) with (3.2), (3.3) gives the following Euler–Lagrange equations.
Proposition 2. An immersion X:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) is a capillary hypersurface if and only if both of the following conditions (i) and (ii) hold.
(i) The anisotropic mean curvature Λ of X is constant on M.
(ii) ⟨˜ξ,˜Nj⟩=ωj on σj (j=1,⋯,k), where ˜ξ is the Cahn–Hoffman field along X.
Proof. Note that X is a capillary hypersurface if and only if, for all volume-preserving variations Xϵ:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk), (−ϵ0<ϵ<ϵ0), δE=0 holds. This is equivalent to the fact that, there exists a constant Λ0 such that
δ(E+nΛ0V)=δFγ+δW+nΛ0δV=−n∫M(Λ−Λ0)⟨δX,ν⟩dA−k∑j=1∮σj⟨δX,(R(p(˜ξ))−ωjρ)⟩ds | (3.4) |
=0 | (3.5) |
holds for all variations Xϵ:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk), (−ϵ0<ϵ<ϵ0, ϵ0>0) of X.
First we assume that X is a capillary hypersurface. By taking variations that preserve the boundary values X|∂M, we have from (3.4), (3.5) that
∫M(Λ−Λ0)ψdA=0,∀ψ∈C∞0(Σ), | (3.6) |
which implies
Λ−Λ0=0,on M. | (3.7) |
This proves (i). Next, take variations that preserve the boundary values X|σ2∪⋯∪σk. Then we have from (3.4), (3.5), (3.6) that
∮σ1⟨δX,(R(p(˜ξ))−ω1ρ)⟩ds=0 | (3.8) |
holds for all variations Xϵ satisfying Xϵ(σ1)⊂Π1. This means that
(R(p(˜ξ))−ω1ρ)∥˜N1,on σ1 | (3.9) |
holds. Note that the (N,ν)-plane is the same as the (˜N1,ρ)-plane because both of them are the orthogonal compliment of the (n−1)-dimensional tangent space of X(σ1) in Rn+1 at each point in X(σ1). Therefore, (3.9) is equivalent to
⟨p(˜ξ)−ω1˜N1,˜N1⟩=0,on σ1. | (3.10) |
And (3.10) is equivalent to
⟨˜ξ−ω1˜N1,˜N1⟩=0,on σ1, | (3.11) |
which means that
⟨˜ξ,˜N1⟩=ω1,on σ1. | (3.12) |
Similarly, we have
⟨˜ξ,˜Nj⟩=ωj,on σj,j=1,⋯,k, | (3.13) |
which proves (ii).
Next we assume that (i) and (ii) are satisfied. Then, by using the computations above, we know that X is a capillary hypersurface.
Theorem 2 is given by Proposition 2.
Here we pose a new variational problem that will be used in the proof of a balancing formula (Lemma 2) in §4. Consider a more general class of surfaces than above:
S:={X:(M,σ1,⋯,σk)→(Rn+1,˜Π1,⋯,˜Πk);Xis an immersion,each˜Πjis any hyperplane that is parallel toΠj,andthe restrictionX|∂Mis an embedding onto the disjoint union ofktopologicalSn−1.}. |
For X∈S, set Cj:=X(σj) and denote by Dj the n-dimensional domain bounded by Cj in ˜Πj. Moreover, set
SX:=X(M)∪D1∪⋯∪Dk. |
Denote by dA the n-dimensional standard volume form on SX. Note that each ˜Nj is a unit normal to ˜Πj. Define the algebraic volume ¯V(X) enclosed by SX as
¯V(X):=1n+1∫M⟨X,ν⟩dA+1n+1k∑j=1∫Dj⟨x,˜Nj⟩dA, |
where we denoted the variable point in Dj by x. And define the energy ¯Fγ(X) of X as
¯Fγ(X)=Fγ(X)+k∑j=1ωjHn(Dj). |
Then, similarly to our capillary problem, we obtain the following first variation formulas.
Lemma 1.
δ¯V=∫M⟨δX,ν⟩dA+k∑j=1∫Dj⟨δX,˜Nj⟩dA, | (3.14) |
δFγ=−∫MnΛ⟨δX,ν⟩−∮∂M⟨δX,R(p(˜ξ))⟩ds, | (3.15) |
δHn(Dj)=∫σj⟨δX,ρ⟩ds. | (3.16) |
Proof. (3.14) is a standard formula (cf. [2]). (3.15) is given in proposition 1. We will prove (3.16). Consider the orthogonal projection pj:˜Πj→Πj. Then,
pj(δX)=δX−⟨δX,˜Nj⟩˜Nj. |
Since
H((Dj)ϵ)=1n∫σj⟨Xϵ,ρϵ⟩dsϵ=1n∫σj⟨pj(Xϵ),ρϵ⟩dsϵ, |
we obtain
δHn(Dj)=∫σj⟨δ(pj(X)),ρ⟩ds=∫σj⟨pj(δX),ρ⟩ds=∫σj⟨δX−⟨δX,˜Nj⟩˜Nj,ρ⟩ds=∫σj⟨δX,ρ⟩ds, |
which proves (3.16).
In order to prove the uniqueness part of Theorem 1, we will show that any capillary hypersurface X:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) is unstable unless it is a part of a rescaling of the Wulff shape.
In view of the condition (ii) in Proposition 2, it is useful to consider the anisotropic energy for (n−1)-dimensional closed hypersurfaces in Πj. First, define hyperplanes Pj in Rn+1 (j=1,⋯,k) by
Pj:={x∈Rn+1;⟨x,˜Nj⟩=ωj}. |
Then, set the followings:
ˆWj:=Wγ∩Pj,ˆOj:=ωj˜Nj. |
Assume that ωj≥0 is sufficiently small so that ˆWj includes at least two distinct points. Then, ˆWj is a strictly convex closed (n−1)-dimensional C∞ hypersurface in the n-dimensional linear space Pj. We regard the point ˆOj as the origin of Pj. Denote by ˆγj:Sn−1→R>0 the support function of ˆWj, that is, for any ρ∈Sn−1, ˆγj(ρ) is the distance between the origin ˆOj and the tangent space of ˆWj at the uniquely-determined point w∈ˆWj such that the outward-pointing unit normal to ˆWj at p coincides with ρ. Then, ˆWj is the Wulff shape for ˆγj. For later use, we denote by ˆξj the Cahn–Hoffman map for ˆγj.
Now, let χ:Sn−1→Πj be a C∞ embedding with outward unit normal ρ=ρj. Define the anisotropic energy of χ by
ˆFj(χ):=∫Sn−1ˆγj(ρ)ds, | (4.1) |
where ds is the (n−1)-dimensional volume form of χ.
From now on, we assume that
X:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) |
is a capillary hypersurface. Set
χj:=X|σj. |
Denote by ρ the outward-pointing unit normal to χj in the hyperplane Πj. X has the following property, which we call the balancing formula that is a generalization of the balancing formula for the isotropic case [3].
Lemma 2.
ˆFj(χj)=−nΛHn(Dj),j=1,⋯,k. | (4.2) |
Proof. Let u be a constant vector in Rn+1. Under parallel translations:
Xt=X+tu, |
¯V, Fγ, Hn(Dj) are invariant. Hence, using the first variation formulas we gave above, we compute
0=ddt|t=0¯Fγ(Xt)=ddt|t=0(¯Fγ(Xt)+nΛ¯V(Xt))=−n∫MΛ⟨u,ν⟩dA−∮∂M⟨u,R(p(˜ξ))⟩ds+k∑j=1ωj∮σj⟨u,ρ⟩ds+nΛ∫M⟨u,ν⟩dA+nΛk∑j=1∫Dj⟨u,˜Nj⟩dA=−∮∂M⟨u,R(p(˜ξ))⟩ds+nΛk∑j=1∫Dj⟨u,˜Nj⟩dA. |
By setting u=(1,0,⋯,0), (0,1,⋯,0), ⋯, (0,⋯,0,1), we have
−nΛk∑j=1Hn(Dj)˜Nj=−k∑j=1∮σjR(p(˜ξ))ds. | (4.3) |
On σj, since ⟨˜ξ,˜Nj⟩=ωj and ⟨˜ξ,ρj⟩=⟨ˆξj,ρj⟩=ˆγj hold, we can write
˜ξ=ωj˜Nj+ˆγjρj+τ, |
where τ is tangent to Cj. Then, we have
R(p(˜ξ))=R(ωj˜Nj+ˆγjρj)=ωjρj−ˆγj˜Nj. | (4.4) |
Note that, by the divergence theorem, it holds that
∮σjρjds=0. |
Hence, substituting Eq (4.4) into Eq (4.3), we obtain
−nΛk∑j=1Hn(Dj)˜Nj=k∑j=1∮σjˆγj˜Njds=k∑j=1ˆFj(χj)˜Nj. | (4.5) |
Because ˜N1, ⋯, ˜Nk are linearly independent, Eq (4.5) implies Eq (4.2).
Now, consider the anisotropic parallel hypersurfaces Xt:=X+t˜ξ (t∈R, |t|<<1) of X (Figure 2, upper left). If ωj>0 for some j∈{1,⋯,k}, then Xt does not satisfy the boundary condition, that is, the boundary Xt(∂M) of the hypersurface may not be included in the boundary ∂Ω⊂Π1∪⋯∪Πk of the wedge-shaped domain Ω. We will prove that, by taking a suitable parallel translation Zt=Xt+t→v of Xt, Zt satisfies the boundary condition (Figure 2, upper right).
For any a∈R, define the hyperplane Πaj in Rn+1 that is parallel to the hyperplane Πj as follows.
Πaj:=Πj+a˜Nj={P+a˜Nj|P∈Πj}. | (4.6) |
Then, we show:
Lemma 3. Xt(σj)⊂Πtωjj holds.
Proof. From Proposition 2 (ii), we have
⟨˜ξ,˜Nj⟩=ωj,on σj, (j=1,⋯,k). | (4.7) |
Since X(σj)⊂Πj, (4.7) gives the desired result.
By using Lemma 3, we will show the following.
Lemma 4. There exists some →v∈Rn+1 such that
⟨→v,˜Nj⟩=−ωj,j=1,⋯,k | (4.8) |
is satisfied, and the parallel translation Zt=Xt+t→v of Xt satisfies the boundary condition, that is,
Zt(σj)⊂Πj,j=1,⋯,k | (4.9) |
holds.
Proof. From Lemma 3,
Xt(σj)⊂Πj+tωj˜Nj |
holds. Hence, for any →v∈Rn+1,
Xt(σj)+t→v⊂Πj+t(ωj˜Nj+→v) |
holds. Therefore, Xt(σj)+t→v⊂Πj if and only if ωj˜Nj+→v∈Πj, which is equivalent to
⟨ωj˜Nj+→v,˜Nj⟩=0, | (4.10) |
that is,
⟨→v,˜Nj⟩=−ωj. | (4.11) |
Now set →v=(v1,⋯,vn+1). Then,
⟨→v,˜Nj⟩=−ωj,j=1,⋯,k | (4.12) |
is a system of linear equations in the (n+1) variables v1,⋯,vn+1 with k equations satisfying k≤n+1. Hence, (4.12) has at least one solution →v, which proves the desired result.
We have proved that Zt=Xt+t→v satisfies the boundary condition. However, it is not a volume-preserving variation in general. We can take suitable homotheties
Yt:=μ(t)Zt=μ(t)(X+t(˜ξ+→v)), μ(t)≥0, μ(0)=1, |
of Zt if necessary so that Yt:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) is a volume-preserving variation of X (Figure 2).
Denote by e(t) the total energy E(Yt) of Yt. Then we obtain
Claim 1.
e"(0)=−1n∫Mγ(ν)∑1≤i<j≤n(kγi−kγj)2dA | (4.13) |
−n−1nk∑j=1ωj(n∫σjˆγj(ρ)ˆΛds+(∫σjˆγj(ρ)ds)2Hn(Dj)), | (4.14) |
where ˆΛ is the anisotropic curvature of χj for ˆγj.
Claim 1 will be proved in §6. Note that, from Remark 2(i), each kγj is real. Since X has constant anisotropic mean curvature Λ, the first term of the right hand side of Eq (4.13) is nonnegative if and only if kγ1=⋯=kγn=Λ/n≠0. Hence, by Corollary 1 in [15], X(M)⊂(1/|Λ|)Wγ holds. Here we used Λ≠0 which is true because of Lemma 2.
Let us study the second term of the right hand side of the equation of e"(0) in Claim 1. Set
Bj:=∫σjˆγj(ρ)ˆΛds+(∫σjˆγ(ρ)ds)2nHn(Dj). | (4.15) |
Then we can prove the following statement (see §7 for its proof).
Claim 2. Bj≥0 holds and that the equality holds if and only if χj(σj)=rˆWj for some r>0.
Now we are in the final position to complete the proof of the first half of Theorem 1. If the capillary hypersurface X is stable, then e"(0)≥0 must hold. Hence, by the above observations, X(M)⊂(1/|Λ|)Wγ holds.
Let X:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) be an anisotropic capillary hypersurface, and we assume that X is an embedding onto a part of Wγ (up to translation and homothety). We will prove that X is stable. It is sufficient to prove the stability for the case where X is an embedding onto a part of Wγ. Then, there exists a point Q∈Rn+1 such that
X(M)=(Wγ+Q)∩¯Ω |
holds. Set
Σ:=X(M)∪D1∪⋯∪Dk=((Wγ+Q)∩¯Ω)∪D1∪⋯∪Dk. | (5.1) |
Then, Σ is a closed convex piecewise-smooth hypersurface in Rn+1. We will derive the support function of Σ ([16]). For any x∈Sn, we define a hyperplane P(x) that is orthogonal to x as follows.
P(x):={y∈Rn+1|⟨x,y⟩=0}. | (5.2) |
Define a continuous function φ:Sn→R>0 as follows.
φ(x):=max{t∈R|(Q+tx+P(x))∩Σ≠∅}. | (5.3) |
Then, the homogeneous extension ¯φ:Rn+1→R of φ is the support function of Σ. Then, Σ is the Wulff shape for φ ([17]), that is, Σ=Wφ holds. Note that
γ(x):=max{t∈R|(tx+P(x))∩Wγ≠∅}=max{t∈R|(Q+tx+P(x))∩(Wγ+Q)≠∅} | (5.4) |
holds. Since γ(x)≥φ(x) holds, we have ˜W[γ]⊃˜W[φ]. Hence we have the followings:
(i) If x∈ν(M), then φ(x)=γ(x).
(ii) If x=˜Nj, then φ(x)=ωj<γ(x), (j=1,⋯,k).
(iii) If x∈Sn∖(ν(M)∪{˜N1,⋯,˜Nk}), then φ(x)<γ(x).
Now let Xt:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) be a volume-preserving variation of X. Denote by Dj(Xt) the domain bounded by Xt(σj) in Πj. Set
Σt:=Xt(M)∪D1(Xt)∪⋯,∪Dk(Xt). |
Then, because γ≥φ and Σ is the minimizer of Fφ among all closed piecewise-smooth hypersurfaces enclosing the same (n+1)-dimensional volume, we obtain
Eγ(X)=Fφ(Σ)≤Fφ(Σt)=Fφ(Xt)+W(Xt)≤Fγ(Xt)+W(Xt)=Eγ(Xt) |
holds. Therefore, X is stable.
Recall
Yt=μ(t)Zt, μ(t)≥0, μ(0)=1, | (6.1) |
Zt=Xt+t→v, | (6.2) |
Xt=X+t˜ξ, | (6.3) |
Yt:(M,σ1,⋯,σk)→(¯Ω,Π1,⋯,Πk) is a volume-preserving variation of X that satisfies the boundary condition. Note that the unit normal vector field along Xt, Zt, and Yt coincide with ν:M→Sn that is the unit normal vector field along X. Hence,
e(t):=E(Yt)=Fγ(Yt)+W(Yt)=∫Mγ(ν)dAYt+k∑j=1ωjHn(Dj(Yt)) |
holds, where dAYt is the n-dimensional volume form of Yt.
Set
V0:=V(X),E0:=E(X),F0:=Fγ(X),W0:=Wγ(X), | (6.4) |
and
v(t):=V(Yt),f(t):=Fγ(Xt),w(t):=Wγ(Xt). | (6.5) |
Then
e(t)=(μ(t))n(f(t)+w(t)), | (6.6) |
v(t)=(μ(t))n+1V(Zt), | (6.7) |
e′(t)=n(μ(t))n−1μ′(t)(f(t)+w(t))+(μ(t))n(f′(t)+w′(t)), | (6.8) |
e"(t)=n(n−1)(μ(t))n−2(μ′(t))2(f(t)+w(t))+2n(μ(t))n−1μ′(t)(f′(t)+w′(t))+n(μ(t))n−1μ"(t)(f(t)+w(t))+(μ(t))n(f"(t)+w"(t)), | (6.9) |
v′(t)=(n+1)(μ(t))nμ′(t)V(Zt)+(μ(t))n+1ddtV(Zt), | (6.10) |
v"(t)=(n+1)n(μ(t))n−1(μ′(t))2V(Zt)+(n+1)(μ(t))nμ"(t)V(Zt)+2(n+1)(μ(t))nμ′(t)ddtV(Zt)+(μ(t))n+1d2dt2V(Zt). | (6.11) |
In order to compute e"(0), we need to compute μ′(0),μ"(0),f′(0)+w′(0),f"(0), and w"(0). First, using the Steiner-type formula (2.4), we obtain the followings.
f′(0)=−n∫MΛγ(ν)dA=−nΛF0, | (6.12) |
f"(0)=∫M2(nC2)γ(ν)Hγ2dA=∫M2γ(ν)σγ2dA=2∫Mγ(ν)∑1≤i<j≤nkγikγjdA. | (6.13) |
Before computing the other derivatives, we prepare two useful lemmas.
Lemma 5. We have the following equalities.
(i)
ddtV(Zt)=E(Zt). | (6.14) |
(ii) In the special case where k=0, that is ∂M=∅, we have
ddtV(Xt)=Fγ(Xt). | (6.15) |
Proof. Since
dAZt=dAXt,˜ξ=Dγ|ν+γ(ν)ν, |
by using the first variation formula (3.2) for the volume, we have
ddtV(Zt)=∫M⟨∂Zt∂t,ν⟩dAXt=∫M(γ(ν)+⟨→v,ν⟩)dAXt=Fγ(Xt)+∫M⟨→v,ν⟩dAXt. | (6.16) |
We compute the last term of (6.16). Denote by dA(Πtωjj) the standard volume form on the n-plane Πtωjj. Using the divergence formula and the equality ⟨→v,˜Nj⟩=−ωj (see Lemma 4), we have
∫M⟨→v,ν⟩dAXt=−k∑j=1∫Dj(Xt)⟨→v,˜Nj⟩dA(Πtωjj)=k∑j=1ωjHn(Dj(Xt))=W(Zt). | (6.17) |
(6.16) with (6.17) gives the desired equality (6.14). The proof of (6.15) is similar.
Lemma 6. We have the following equality.
E0=−(n+1)ΛV0. | (6.18) |
Proof. Consider the variation ˆXϵ:=(1+ϵ)X of X, and set F(ϵ):=Fγ(ˆXϵ). Then,
F(ϵ)=(1+ϵ)nF0. |
Hence,
F′(0)=nF0. | (6.19) |
On the other hand, the first variation formula (3.1) of Fγ gives
F′(0)=−n∫MΛ⟨X,ν⟩dA−∫∂M⟨X,R(p(˜ξ))⟩ds=−n(n+1)ΛV0−k∑j=1∫σj⟨X,R(p(˜ξ))⟩ds. | (6.20) |
We compute the integrand of the second term of (6.20) on σj. Proposition 2 (ii) gives ⟨˜ξ,˜Nj⟩=ωj on σj. Hence,
p(˜ξ)=ωj˜Nj+⟨p(˜ξ),ρ⟩ρ. | (6.21) |
Using (6.21) and the equality ⟨X,˜Nj⟩=0, we have
⟨X,R(p(˜ξ))⟩=⟨X,ωjρ−⟨p(˜ξ),ρ⟩˜Nj⟩=ωj⟨X,ρ⟩. | (6.22) |
Inserting (6.22) to (6.20), we obtain
F′(0)=−n(n+1)ΛV0−k∑j=1ωj∫σj⟨X,ρ⟩ds=−n(n+1)ΛV0−nk∑j=1ωjHn(Dj)=−n(n+1)ΛV0−nW0. | (6.23) |
(6.19) with (6.23) gives the desired equality (6.18).
Let us continue the proof of Claim 1. Using the equalities (6.1), (6.10), and (6.14), we have
0=v′(0)=(n+1)μ′(0)V0+E0. | (6.24) |
Hence we have
μ′(0)=−E0(n+1)V0. | (6.25) |
Next we compute f′(0)+w′(0). Note that e′(0)=0 because X is a capillary hypersurface. Using (6.8) and (6.25), we obtain
0=e′(0)=nμ′(0)E0+f′(0)+w′(0)=−nE20(n+1)V0+f′(0)+w′(0), | (6.26) |
which implies that
f′(0)+w′(0)=nE20(n+1)V0 | (6.27) |
holds.
Now we compute w"(0). Using the first variation formula (3.2) for volume, we obtain
w′(t)=k∑j=1ωj∫σj⟨˜ξ,ρt⟩dst=k∑j=1ωjˆFj(Xt|σj), | (6.28) |
where ρt is the outward-pointing unit normal vector field along Xt|σj:σj→Πtωjj, and dst is the (n−1)-dimensional volume form of Xt|σj. Using (6.28) and the first variation formula (3.1) of the anisotropic energy, we obtain
w"(0)=−(n−1)k∑j=1ωj∫σjˆΛˆγj(ρ)ds. | (6.29) |
Finally we compute μ"(0) by using v"(0)=0 and (6.11). From (6.14), we have
ddtV(Zt)|t=0=E0, | (6.30) |
d2dt2V(Zt)|t=0=ddtE(Zt)|t=0=nE20(n+1)V0, | (6.31) |
here in the last equality we used (6.8). Inserting (6.25), (6.30), (6.31) to (6.11), we compute to obtain
μ"(0)=2E20(n+1)2V20. | (6.32) |
Inserting (6.13), (6.25), (6.8), (6.29), and (6.32) to (6.9) at t=0, we obtain
e"(0)=−n(n−1)E30(n+1)2V20+2∫Mγ(ν)∑1≤i<j≤nkγikγjdA−(n−1)k∑j=1ωj∫σjˆΛˆγj(ρ)ds. | (6.33) |
From Lemma 6, we have
E0V0=−(n+1)Λ. |
Inserting this to (6.33), we obtain
e"(0)=−n(n−1)Λ2E0+2∫Mγ(ν)∑1≤i<j≤nkγikγjdA−(n−1)k∑j=1ωj∫σjˆΛˆγj(ρ)ds=−1n∫Mγ(ν)∑1≤i<j≤n(kγi−kγj)2dA−n(n−1)Λ2W0−(n−1)k∑j=1ωj∫σjˆΛˆγj(ρ)ds. | (6.34) |
Now we are in the final stage to prove Claim 1. The balancing formula (4.2) implies
Λ=−∫σjˆγj(ρ)dsnHn(Dj),j=1,⋯,k. |
Hence
−n(n−1)Λ2W0=−n(n−1)k∑j=1Λ2ωjHn(Dj)=−n−1nk∑j=1ωj(∫σjˆγj(ρ)ds)2Hn(Dj) |
holds. This with (6.34) proves Claim 1.
In this section, we examine the sign of
Bj:=∫σjˆγj(ρ)ˆΛds+(ˆFj(χj))2nHn(Dj), | (7.1) |
where
ˆFj(χj)=∫σjˆγ(ρ)ds. |
First, we recall two known useful propositions, which we prove for completeness.
Proposition 3. Let γ:Sm→R>0 be a positive strictly convex function of class at least C2, and W⊂Rm+1 be its Wulff shape. Denote by Fγ(W) the anisotropic energy of W for γ, and by Hm+1(W) the (m+1)-dimensional Hausdorff measure of the domain bounded by W in Rm+1. Then,
Fγ(W)=(m+1)Hm+1(W) | (7.2) |
holds.
Proof. Let ξ:Sm→Rm+1 be the Cahn-Hoffman map of γ defined by ξ(ν)=Dγ|ν+γ(ν)ν. Then ξ is an embedding onto W. Hence, denoting by dAξ the volume form of ξ, we have
Fγ(W)=∫Smγ(ν)dAξ=∫Sm⟨ξ(ν),ν⟩dAξ=(m+1)Hm+1(W), |
which proves (7.2).
Proposition 4 (Anisotropic isoperimetric inequality). Let γ, W be the same as in Proposition 3. We also use the same notation as in Proposition 3. Then, for any closed embedded piecewise-smooth hypersurface M in Rm+1, it holds that
(Fγ(M))m+1≥(m+1)m+1Hm+1(W)(Hm+1(M))m, | (7.3) |
where the equality holds if and only if M coincides with W up to homothety and translation in Rm+1.
Proof. Recall that the Wulff shape W is the minimizer of Fγ among closed hypersurfaces enclosing the same (m+1)-dimensional Hausdorff measure. Set
c=(Hm+1(W)Hm+1(M))1m+1. |
Then, the hypersurface
Mc:=cM |
similar to M encloses the same (m+1)-dimensional Hausdorff measure as W. Hence,
Fγ(Mc)≥Fγ(W) | (7.4) |
holds, where the equality holds if and only if M coincides with W up to homothety and translation. On the other hand,
Fγ(Mc)=cmFγ(M)=(Hm+1(W)Hm+1(M))mm+1Fγ(M). | (7.5) |
The inequality (7.4) combined with the equalities (7.2) and (7.5) gives (7.3).
Now we examine Bj. From now on, for simplicity, we identify an embedded closed hypersurface in an euclidean space with the domain bounded by this hypersurface. We also identify an embedded closed hypersurface with its representation mapping.
Using Proposition 4, we have
Bj=∫σjˆγj(ρ)ˆΛds+(ˆFj(χj))2nHn(Dj)≥∫σjˆγj(ρ)ˆΛds+n(Hn(ˆWγ))2n(Hn(Dj))n−2n. | (7.6) |
First we study the special case where n=2. In this case, using (7.6) with Proposition 3, we have
Bj≥∫σjˆγj(ρ)ˆΛds+2H2(ˆWj)=∫σjˆγj(ρ)ˆΛds+ˆFj(ˆWj). | (7.7) |
We use the representation (2.2) of the anisotropic mean curvature in Remark 1. Note that σj is topologically S1. Denote by G:ˆWj→S1 the outward-pointing unit normal vector field on ˆWj and by ˆs the arc-length parameter of ˆWj. Then G=ˆξ−1j, where ˆξj:S1→ˆWj is the Cahn-Hoffman map of ˆWj. Hence, if we take ρ∈S1 as the parameter of ˆWj through ˆξj, G is the identity mapping on S1. We also denote by κj, ˆκj the curvatures of χj, ˆWj with respect to the outward-pointing unit normal, respectively. Then, using (2.2), we have
∫σjˆγj(ρ)ˆΛds=−∫σjˆγj(ρ)κjˆκjds=−∫S1ˆγj(ρ)dρdsdGdˆsds=−∫S1ˆγj(ρ)dρdGdˆs=−∫S1ˆγj(ρ)dˆs=−ˆFj(ˆWj). | (7.8) |
Inserting (7.8) into (7.7), we have Bj≥0.
Next we assume n≥3. We assume that Dj are convex. Below, for simplicity, we omit the subscript j, that is, we write D instead of Dj, for instance. On the Minkowski sum D+tˆW, there holds
Hn(D+tˆW)=n∑i=0(nCi)tiv((n−i)times⏞D,…,D,itimes⏞ˆW,…,ˆW), | (7.9) |
where v(K1,⋯,Kn) is the so-called the mixed volume of convex bodies K1,⋯,Kn in Rn ([16, Theorem 5.1.7]). On the other hand, from Lemma 5 and the Steiner-type formula (2.4), we have
Hn(D+tˆW)=Hn(χ+tˆξ)=Hn(χ)+∫t0ˆF(χ+tˆξ)dt=Hn(D)+∫t0(∫σˆγ(ρ)n−1∑r=0(−1)r(n−1Cr)trˆHˆγrds)dt=Hn(D)+n−1∑r=0(−1)rr+1(n−1Cr)tr+1∫σˆγ(ρ)ˆHˆγrds. | (7.10) |
Comparing (7.9) with (7.10), we obtain
v(ntimes⏞D,…,D)=Hn(D), | (7.11) |
v((n−1)times⏞D,…,D,ˆW)=1n∫σˆγ(ρ)ds, | (7.12) |
v((n−2)times⏞D,…,D,ˆW,ˆW)=−1n∫σˆγ(ρ)ˆΛds. | (7.13) |
The Minkowski's second inequality ([16,Theorem 7.2.1]) gives
(v((n−1)times⏞D,…,D,ˆW))2≥v(ntimes⏞D,…,D)⋅v((n−2)times⏞D,…,D,ˆW,ˆW), | (7.14) |
where the equality holds if and only if D is homothetic to ˆW. Combining (7.11)–(7.14), we obtain
(∫σˆγ(ρ)ds)2≥−nHn(D)∫σˆγ(ρ)ˆΛds, | (7.15) |
here the equality holds if and only if D is homothetic to ˆW. This completes the proof of Claim 2.
This work was partially supported by JSPS KAKENHI Grant Number JP20H01801, JSPS Grant-in-Aid for Scientific Research on Innovative Areas "Discrete Geometric Analysis for Materials Design": Grant Numbers JP18H04487 and JP20H04642, JST CREST GrantNumber JPMJCR1911, and JSPS(Japan)-FWF(Austria) Joint Research Project (2018-2019, Project leader (Japan): Miyuki Koiso).
The authors declare no conflict of interest.
[1] |
N. Ando, Hartman-Wintner's theorem and its applications, Calc. Var., 43 (2012), 389–402. http://doi.org/10.1007/s00526-011-0415-x doi: 10.1007/s00526-011-0415-x
![]() |
[2] |
J. L. Barbosa, M. Do Carmo, J. Eschenburg, Stability of hypersurfaces of constant mean curvature in Riemannian manifolds, Math. Z., 197 (1988), 123–138. https://doi.org/10.1007/BF01161634 doi: 10.1007/BF01161634
![]() |
[3] |
J. Choe, M. Koiso, Stable capillary hypersurfaces in a wedge, Pac. J. Math., 280 (2016), 1–15. http://doi.org/10.2140/pjm.2016.280.1 doi: 10.2140/pjm.2016.280.1
![]() |
[4] |
Y. He, H. Li, Integral formula of Minkowski type and new characterization of the Wulff shape, Acta. Math. Sin.-English Ser., 24 (2008), 697–704. https://doi.org/10.1007/s10114-007-7116-6 doi: 10.1007/s10114-007-7116-6
![]() |
[5] |
Y. He, H. Li, H. Ma, J. Ge, Compact embedded hypersurfaces with constant higher order anisotropic mean curvatures, Indiana Univ. Math. J., 58 (2009), 853–868. http://doi.org/10.1512/iumj.2009.58.3515 doi: 10.1512/iumj.2009.58.3515
![]() |
[6] | M. Koiso, Uniqueness of stable closed non-smooth hypersurfaces with constant anisotropic mean curvature, arXiv.1903.03951. |
[7] |
M. Koiso, Uniqueness of closed equilibrium hypersurfaces for anisotropic surface energy and application to a capillary problem, Math. Comput. Appl., 24 (2019), 88. https://doi.org/10.3390/mca24040088 doi: 10.3390/mca24040088
![]() |
[8] |
M. Koiso, B. Palmer, Geometry and stability of surfaces with constant anisotropic mean curvature, Indiana Univ. Math. J., 54 (2005), 1817–1852. http://doi.org/10.1512/iumj.2005.54.2613 doi: 10.1512/iumj.2005.54.2613
![]() |
[9] |
M. Koiso, B. Palmer, Stability of anisotropic capillary surfaces between two parallel planes, Calc. Var., 25 (2006), 275–298. http://doi.org/10.1007/s00526-005-0336-7 doi: 10.1007/s00526-005-0336-7
![]() |
[10] |
M. Koiso, B. Palmer, Anisotropic capillary surfaces with wetting energy, Calc. Var., 29 (2007), 295–345. http://doi.org/10.1007/s00526-006-0066-5 doi: 10.1007/s00526-006-0066-5
![]() |
[11] |
M. Koiso, B. Palmer, Uniqueness theorems for stable anisotropic capillary surfaces, SIAM J. Math. Anal., 39 (2007), 721–741. https://doi.org/10.1137/060657297 doi: 10.1137/060657297
![]() |
[12] |
M. Koiso, B. Palmer, Anisotropic umbilic points and Hopf's theorem for surfaces with constant anisotropic mean curvature, Indiana Univ. Math. J., 59 (2010), 79–90. http://doi.org/10.1512/iumj.2010.59.4164 doi: 10.1512/iumj.2010.59.4164
![]() |
[13] |
H. Li, C. Xiong, Stability of capillary hypersurfaces with planar boundaries, J. Geom. Anal., 27 (2017), 79–94. https://doi.org/10.1007/s12220-015-9674-7 doi: 10.1007/s12220-015-9674-7
![]() |
[14] |
B. Palmer, Stability of the Wulff shape, Proc. Am. Math. Soc., 126 (1998), 3661–3667. http://doi.org/10.1090/S0002-9939-98-04641-3 doi: 10.1090/S0002-9939-98-04641-3
![]() |
[15] |
R. C. Reilly, The relative differential geometry of nonparametric hypersurfaces, Duke Math. J., 43 (1976), 705–721. http://doi.org/10.1215/S0012-7094-76-04355-6 doi: 10.1215/S0012-7094-76-04355-6
![]() |
[16] | R. Schneider, Convex bodies: the Brunn-Minkowski theory, 2 Eds., New York: Cambridge University Press, 2014. https://doi.org/10.1017/CBO9781139003858 |
[17] |
J. E. Taylor, Crystalline variational problems, Bull. Amer. Math. Soc., 84 (1978), 568–588. https://doi.org/10.1090/S0002-9904-1978-14499-1 doi: 10.1090/S0002-9904-1978-14499-1
![]() |
[18] |
H. Weyl, On the volume of tubes, Amer. J. Math., 61 (1939), 461–472. https://doi.org/10.2307/2371513 doi: 10.2307/2371513
![]() |
[19] |
W. L. Winterbottom, Equilibrium shape of a small particle in contact with a foreign substrate, Acta Metal., 15 (1967), 303–310. https://doi.org/10.1016/0001-6160(67)90206-4 doi: 10.1016/0001-6160(67)90206-4
![]() |
[20] |
G. Wulff, Zur Frage der Geschwindigkeit des Wachsthums und der Auflösung der Krystallflächen, Zeitschrift für Krystallographie und Mineralogie, 34 (1901), 449–530. https://doi.org/10.1524/zkri.1901.34.1.449 doi: 10.1524/zkri.1901.34.1.449
![]() |
1. | Xiaohan Jia, Guofang Wang, Chao Xia, Xuwen Zhang, Alexandrov’s Theorem for Anisotropic Capillary Hypersurfaces in the Half-Space, 2023, 247, 0003-9527, 10.1007/s00205-023-01861-0 | |
2. | César Rosales, Compact anisotropic stable hypersurfaces with free boundary in convex solid cones, 2023, 62, 0944-2669, 10.1007/s00526-023-02528-0 |