Citation: Matteo Novaga, Marco Pozzetta. Connected surfaces with boundary minimizing the Willmore energy[J]. Mathematics in Engineering, 2020, 2(3): 527-556. doi: 10.3934/mine.2020024
[1] | Miyuki Koiso . Stable anisotropic capillary hypersurfaces in a wedge. Mathematics in Engineering, 2023, 5(2): 1-22. doi: 10.3934/mine.2023029 |
[2] | Daniela De Silva, Ovidiu Savin . Uniform density estimates and Γ-convergence for the Alt-Phillips functional of negative powers. Mathematics in Engineering, 2023, 5(5): 1-27. doi: 10.3934/mine.2023086 |
[3] | Ko-Shin Chen, Cyrill Muratov, Xiaodong Yan . Layered solutions for a nonlocal Ginzburg-Landau model with periodic modulation. Mathematics in Engineering, 2023, 5(5): 1-52. doi: 10.3934/mine.2023090 |
[4] | Giovanni Di Fratta, Alberto Fiorenza, Valeriy Slastikov . On symmetry of energy minimizing harmonic-type maps on cylindrical surfaces. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023056 |
[5] | Fausto Ferrari, Nicolò Forcillo . A new glance to the Alt-Caffarelli-Friedman monotonicity formula. Mathematics in Engineering, 2020, 2(4): 657-679. doi: 10.3934/mine.2020030 |
[6] | Manuel Friedrich . Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials. Mathematics in Engineering, 2020, 2(1): 75-100. doi: 10.3934/mine.2020005 |
[7] | Morteza Fotouhi, Andreas Minne, Henrik Shahgholian, Georg S. Weiss . Remarks on the decay/growth rate of solutions to elliptic free boundary problems of obstacle type. Mathematics in Engineering, 2020, 2(4): 698-708. doi: 10.3934/mine.2020032 |
[8] | Paolo Maria Mariano, Domenico Mucci . Equilibrium of thin shells under large strains without through-the-thickness shear and self-penetration of matter. Mathematics in Engineering, 2023, 5(6): 1-21. doi: 10.3934/mine.2023092 |
[9] | 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 |
[10] | Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040 |
Let φ:Σ→R3 be an immersion of a 2-dimensional manifold Σ with boundary ∂Σ in the Euclidean space R3. We say that an immersion is smooth if it is of class C2. In such a case we define the second fundamental form of φ in local coordinates as
IIij(p)=(∂ijφ(p))⊥, |
for any p∈Σ∖∂Σ, where (⋅)⊥ denotes the orthogonal projection onto (dφ(TpΣ))⊥. Denoting by gij=⟨∂iφ,∂jφ⟩ the induced metric tensor on Σ and by gij the components of its inverse, we define the mean curvature vector by
→H(p)=12gij(p)IIij(p), |
for any p∈Σ∖∂Σ, where sum over repeated indices is understood. The normalization of →H is such that the mean curvature vector of the unit sphere points inside the ball and it has norm equal to one. Denoting by μφ the volume measure on Σ, we define the Willmore energy of φ by
W(φ)=∫Σ|→H|2dμφ. |
For an immersion φ:Σ→R3 we will denote by coφ:∂Σ→R3 the conormal field, i.e., the unit vector field along ∂Σ belonging to dφ(TΣ)∩(dφ|∂Σ(T∂Σ))⊥ and pointing outside of φ(Σ).
The study of variational problems involving the Willmore energy has begun with the works of T. Willmore ([32,33]), in which he proved that round spheres minimize W among every possible immersed compact surface without boundary. The Willmore energy of a sphere is 4π. In [32] the author proposed his celebrated conjecture, claiming that the infimum of W among immersed smooth tori was 2π2. Such conjecture (eventually proved in [19]) motivated the variational study of W in the setting of smooth surfaces without boundary. In such setting many fundamental results have been achieved, and some of them (in particular [15,26,31]) developed a very useful variational approach, that today goes under the name of Simon's ambient approach. Such method relies on the measure theoretic notion of varifold as a generalization of the concept of immersed submanifold. We remark that, more recently, an alternative and very powerful variational method based on a weak notion of immersions has been developed in [23,24,25].
Following Simon's approach, the concept of curvature varifold with boundary ([14,18]), considered as a good generalization of smooth immersed surfaces, will be fundamental in this work. Such notion is recalled in Appendix A. We will always consider integer rectifiable curvature varifolds with boundary, that we will usually call simply varifolds. Roughly speaking a rectifiable varifold is identified by a couple v(M,θV), where M⊂R3 is 2-rectifiable and θV:M→N≥1 is locally H2-integrable on M, and we think at it as a 2-dimensional object in R3 whose points p come with a weight θV(p). We recall here that a 2-dimensional varifold V=v(M,θV) has weight measure μV=θVH2└M, that is a Radon measure on R3; moreover it has (generalized) mean curvature vector →H∈L1loc(μV;R3) and generalized boundary σV if
∫divTMXdμV=−2∫⟨→H,X⟩dμV+∫XdσV∀X∈C1c(R3;R3), |
where σV is a Radon R3-valued measure on R3 of the form σV=νVσ, with |νV|=1 σ-ae and σ is singular with respect to μV; also divTMX(p)=tr(P⊤∘∇X(p)) where P⊤ is the matrix corresponding to the projection onto TpM, that is defined H2-ae on M.
By analogy with the case of sooth surfaces, we define the Willmore energy of a varifold V=v(M,θV) by setting
W(V)=∫|→H|2dμV∈[0,+∞], |
if V has generalized mean curvature →H, and W(V)=+∞ otherwise.
A rectifiable varifold V=v(M,θV) defines a Radon measure on G2(R3):=R3×G2,3, where G2,3 is the Grassmannian of 2-subspaces of R3, identified with the metric space of matrices corresponding to the orthogonal projection on such subspaces. More precisely for any f∈C0c(G2(R3)) we define
V(f):=∫G2(R3)f(p,P)dV(p,P)=∫R3f(p,TpM)dμV(p). |
In this way a good notion of convergence in the sense of varifolds is defined, i.e., we say that a sequence Vn=v(Mn,θVn) converges to V=v(M,θV) as varifolds if
Vn(f)→V(f), |
for any f∈C0c(G2(R3)).
More recently, varifolds with boundary and Simon's method have been used also in the study of variational problems in the presence of boundary conditions. A seminal work is [26], in which the author constructs branched surfaces with boundary that are critical points of the Willmore energy with imposed clamped boundary conditions, i.e., with fixed boundary curve and conormal field. Another remarkable work is [11], in which an analogous result is achieved in the minimization of the Helfich energy. We also mention [22], in which the minimization problem of the Willmore energy of surfaces with boundary with fixed topology is considered, and the only constraint is the boundary curve, while the conormal is free, yielding the so-called natural Navier boundary condition. A couple of previous works in which Simon's method is applied in the study of closed surfaces are [21] and [28].
If γ=γ1∪...∪γα is a finite disjoint union of smooth closed compact embedded curves, a classical formulation of the Plateau's problem with datum γ may be to solve the minimization problem
min{μφ(Σ)|φ:Σ→R3,φ|∂Σ:∂Σ→γ embedding}, | (1.1) |
that is one wants to look for the surface of least area having the given boundary. From a physical point of view, solutions of the Plateau's problem are good models of soap elastic films having the given boundary [20]. Critical points of the Plateau's problem are called minimal surfaces and they are characterized by having zero mean curvature (this is true also in the non-smooth context of varifolds in the appropriate sense, see [30]). In particular, minimal surfaces or varifolds with vanishing mean curvature have zero Willmore energy. However, as we are going to discuss, the Plateau's problem, and more generally the minimization of the Area functional, may be incompatible with some constraints, such as a connectedness constraint.
In this paper we want to study the minimization of the Willmore energy of varifolds V with given boundary conditions, i.e., both conditions of clamped or natural type on the generalized boundary σV, adding the constraint that the support of the varifold must connect the assigned curves γ1,...,γα. Hence the minimization problems we will study have the form
P:=min{W(V)|V=v(M,θV):σV=σ0,suppV∪γ compact, connected }, | (1.2) |
for some assigned vector valued Radon measure σ0, or
Q:=min{W(V)|V=v(M,θV):|σV|≤μ,suppV∪γ compact, connected }, | (1.3) |
for some assigned positive Radon measure μ with suppμ=γ.
Let us introduce a remarkable particular case that motivates our study. Let C=[0,1]2/∼ be a cylinder. Let R≥1 and h>0. We define
ΓR,h:={x2+y2=1,z=h}∪{x2+y2=R2,z=−h},R≥1,h>0, |
that is a disjoint union of two parallel circles of possibly different radii. We consider the class of immersions
FR,h:={φ:C→R3|φ smooth immersion,φ|∂C:∂C→ΓR,h smooth embedding}. |
By Corollary 3 in [27], if a minimal surface has ΓR,h as boundary, then it necessarily is a catenoid or a pair of planar disks. Moreover there exists a threshold value h0>0 such that ΓR,h is the boundary of a catenoid if and only if h≤h0. For example, in the case of R=1 one has h0=(mint>0cosh(t)t)−1. In particular for any h>h0 there are no minimal surfaces (and thus no solutions of the Plateau's problem) connecting the two components of ΓR,h, even in a perturbative setting h≃h0+ε. This rigidity in the behavior of minimal surfaces suggests that in some cases an energy different from the Area functional may be a good model for connected soap films, like for describing the optimal elastic surface connecting ΓR,h in the perturbative case h≃h0+ε. Since surfaces with zero Willmore energy recover critical points of the Plateau's problem, we expect the minimization of W to be a good process for describing optimal elastic surfaces under constraints, like connectedness ones, that do not match with the Area functional.
Also, from the modeling point of view, we remark the importance of Willmore-type energies, like the Helfrich energy, in the physical study of biological membranes ([12,29]), and in the theory of elasticity in engineering (see [13] and references therein).
We have to mention some remarkable results about critical points of the Willmore energy (called Willmore surfaces) with boundary. Apart from the above cited [26], Willmore surfaces with a boundary also of the form ΓR,h have been studied together with the rotational symmetry of the surface in [3,4,6,7,8,9,10]; a new result about symmetry breaking is [17]. Also, interesting results about Willmore surfaces in a free boundary setting is contained in [1]. A relation between Willmore surfaces and minimal surfaces is investigated in [5].
Let us collect here the main results of the paper. If γ=γ1∪...∪γα is a disjoint union of smooth embedded compact 1-dimensional manifolds, we give a sufficient condition guaranteeing existence in minimization problems of the form (1.2) or (1.3). We obtain the following two Existence Theorems.
Theorem 4.1. Let γ=γ1∪...∪γα be a disjoint union of smooth embedded compact 1-dimensional manifolds with α∈N≥2.
Let
σ0=ν0mH1└γ |
be a vector valued Radon measure, where m:γ→N≥1 and ν0:γ→(Tγ)⊥ are H1-measurable functions with m∈L∞(H1└γ) and |ν0|=1 H1-ae.
Let P be the minimization problem
P:=min{W(V)|V=v(M,θV):σV=σ0,suppV∪γ compact, connected }. | (1.4) |
If infP<4π, then P has minimizers.
Theorem 4.2. Let γ=γ1∪...∪γα be a disjoint union of smooth embedded compact 1-dimensional manifolds with α∈N≥2.
Let m:γ→N≥1 by H1-measurable with m∈L∞(H1└γ).
Let Q be the minimization problem
Q:=min{W(V)|V=v(M,θV):|σV|≤mH1└γ,suppV∪γ compact, connected }. | (1.5) |
If infQ<4π, then Q has minimizers.
Both Existence Theorems are obtained by applying a direct method in the context of varifolds. The technical assumption on the fact that the infimum of the considered problem is strictly less than 4π guarantees compactness of minimizing sequences; we mention that it is an open problem to understand whether a uniform bound possibly greater than 4π on the Willmore energy of a sequence of varifolds with boundary implies precompactness of the sequence, even in presence of boundary conditions. In both cases the connectedness constraint passes to the limit by means of the following theorem, that relates varifolds convergence with convergence in Hausdorff distance of the supports of the varifolds.
Theorem 3.4. Let Vn=v(Mn,θVn)≠0 be a sequence of curvature varifolds with boundary with uniformly bounded Willmore energy converging to V=v(M,θV)≠0. Suppose that the Mn's are connected and uniformly bounded.
Suppose that suppσVn=γ1n∪...∪γαn where the γin's are disjoint compact embedded 1-dimensional manifolds, ˉγ1,...,ˉγβ with β≤α are disjoint compact embedded 1-dimensional manifolds, and assume that γin→ˉγi in dH for i=1,...,β and that H1(γin)→0 for i=β+1,...,α.
Then Mn→M∪ˉγ1∪...∪ˉγβ in Hausdorff distance dH (up to subsequence) and M∪ˉγ1∪...∪ˉγβ is connected. Moreover γin→{pi} in dH for any i=β+1,...,α for some points {pi}, each pi∈M, and suppσV⊂ˉγ1∪...∪ˉγβ∪{pβ+1,...,βα}.
The paper is organized as follows. In Section 2 we recall the monotonicity formula for curvature varifolds with boundary and its consequences on the structure of varifolds with bounded Willmore energy. Such properties are proved in Appendix B. In Section 3 we prove some properties of the Hausdorff distance and we prove Theorem 3.4. Section 4 is devoted to the proof of the Existence Theorems 4.1 and 4.2; we also describe remarkable cases in which such theorems apply, such as in the above discussed perturbative setting. Theorem 3.4 and the monotonicity formula give us results also about the asymptotic behavior of connected varifolds with suitable boundedness assumptions; more precisely we prove that rescalings of a sequence of varifolds Vn with diam(suppVn)→∞ converge to a sphere both as varifolds and in Hausdorff distance (Corollary 5.2). Finally in Section 6 we apply all the previous results to the motivating case of varifolds with boundary conditions on curves of the type of ΓR,h. We prove that for any R and h the minimization problem of type Q has minimizers and their rescalings asymptotically approach a sphere (Corollary 6.2). Appendix A recalls the definitions about curvature varifolds with boundary and a useful compactness theorem.
We adopt the following notation.
● The symbol Br(p) denotes the open ball of radius r and center p in R3.
● The symbol ⟨⋅,⋅⟩ denotes the Euclidean inner product.
● The symbol Hk denotes the k-dimensional Hausdorff measure in R3.
● The symbol dH denotes the Hausdorff distance.
● If φ:Σ→R3 is a smooth immersion of a 2-dimensional manifold with boundary, then in local coordinates we denote by IIij the second fundamental form, by →H the mean curvature vector, by gij the metric tensor, by gij its inverse, by μφ the volume measure on Σ induced by φ, and by coφ the conormal field.
● If v is a vector and M is 2-rectifiable in R3, the symbol (v)⊥ denotes the projection of v onto TpM⊥; hence v⊥ is defined H2-ae on M and it implicitly depends on the point p∈M.
● The symbol V=v(M,θV) denotes an integer rectifiable varifold. Also μV=θVH2└M is the weight measure. If they exist, the generalized mean curvature and boundary are usually denoted by →H (or →HV) and σV.
● The symbol C denotes a fixed cylinder, i.e., C=[0,1]2/∼.
● For given R≥1 and h>0, the symbol ΓR,h denotes an embedded 1-dimensional manifold of the form
ΓR,h:={x2+y2=1,z=h}∪{x2+y2=R2,z=−h},R≥1,h>0, |
that is a disjoint union of two parallel circles of possibly different radii. Observe that the distance between the two circles is equal to 2h.
● For a given boundary datum ΓR,h as above, we define the class
FR,h:={φ:C→R3|φ smooth immersion,φ|∂C:∂C→ΓR,h smooth embedding}. |
Here we recall the fundamental monotonicity formula for curvature varifolds with boundary, together with some immediate consequences on surfaces and on the structure of varifolds with finite Willmore energy.
This classical formula is completely analogous to its version without boundary ([15,31]), hence we refer to Appendix B for the technicalities we need.
Let 0<σ<ρ and p0∈R3. If V is an integer rectifiable curvature varifold with boundary with bounded Willmore energy (here the support of V is not necessarily bounded), with μV the induced measure in R3, and generalized boundary σV, it holds that
A(σ)+∫Bρ(p0)∖Bσ(p0)|→H2+(p−p0)⊥|p−p0|2|2dμV(p)=A(ρ), | (2.1) |
where
A(ρ):=μV(Bρ(p0))ρ2+14∫Bρ(p0)|H|2dμV(p)+Rp0,ρ, | (2.2) |
and
Rp0,ρ:=∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)+12∫Bρ(p0)(1|p−p0|2−1ρ2)(p−p0)dσV(p)=:∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)+Tp0,ρ. | (2.3) |
In particular the function ρ↦A(ρ) is non-decreasing.
When more than a varifold is involved, we will usually denote by AV(⋅) the monotone quantity associated to V for chosen p0∈R3.
It is useful to remember that Tp0,ρ=0 if Bρ(p0)∩suppσV=∅, and that
|∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)|→ρ→00 | (2.4) |
whenever W(V)<+∞ and p0∉suppσV (see (B.8) in Appendix B).
Let us list some immediate consequences on surfaces with boundary.
Lemma 2.1. Let Σ⊂R3 be a compact connected immersed surface with boundary. Then
∀p0∈R3:4limσ↘0|Σ∩Bσ(p0)|σ2+4∫Σ|→H2+(p−p0)⊥|p−p0|2|2=W(Σ)+2∫∂Σ⟨p−p0|p−p0|2,co⟩. | (2.5) |
In particular
∀p0∈R3∖∂Σ:4limσ↘0|Σ∩Bσ(p0)|σ2+4∫Σ|→H2+(p−p0)⊥|p−p0|2|2≤W(Σ)+2H1(∂Σ)d(p0,∂Σ). | (2.6) |
Moreover calling dH the Hausdorff distance (see Section 3) and writing dH(Σ,∂Σ)=d(¯p0,∂Σ) for some ¯p0∈Σ∖∂Σ, it holds that
4limσ↘0|Σ∩Bσ(¯p0)|σ2+4∫Σ|→H2+(p−¯p0)⊥|p−¯p0|2|2≤W(Σ)+2H1(∂Σ)dH(Σ,∂Σ). | (2.7) |
Proof. It suffices to prove (2.5). Since Σ is smooth we have that
|∫Bρ(p0)(1|p−p0|2−1ρ2)⟨p−p0,co⟩dH1(p)|≤∫Bρ(p0)|1|p−p0|2−1ρ2|Op0(|p−p0|2)dH1(p)→ρ→00. |
Since Σ is smooth, by (2.1) we have that
A(σ)→σ→0limσ↘0|Σ∩Bσ(p0)|σ2, |
while by compactness it holds that
A(ρ)→ρ→∞14W(Σ)+12∫∂Σ⟨p−p0|p−p0|2,co⟩, |
and we get (2.5).
Let us mention that (2.6) already appears in [24].
More importantly, the monotonicity formula implies fundamental structural properties on varifolds with bounded Willmore energy. First we remark such results in the case of varifolds without boundary, as proved in [15].
Remark 2.2. Let V=v(M,θV) be an integer rectifiable varifold with σV=0 and finite Willmore energy. Then at any point p0∈R3 there exists the limit
limr→0μV(Br(p0))πr2=θV(p0), | (2.8) |
and θV is upper semicontinuous on R3 (see (A.7) and (A.9) in [15]). In particular M={p∈R3:θV(p)≥12} is closed.
Recall that if suppV is also compact and non-empty, then W(V)≥4π ((A.19) in [15]) and θV is uniformly bounded on R3 by a constant depending only on W(V) ((A.16) in [15]).
In complete analogy with Remark 2.2 we prove in Appendix B (see Proposition) that if V is a 2-dimensional integer rectifiable curvature varifold with boundary, denoting by S a compact 1-dimensional embedded manifold containing the support suppσV with |σV|(S)<+∞ and assuming that
W(V)<+∞,lim supR→∞μV(BR(0))R2≤K<+∞, |
then the limit
limρ↘0μV(Bρ(p))ρ2 |
exists at any point p∈R3∖S, the multiplicity function θV(p)=limρ↘0μV(Bρ(p))ρ2 is upper semicontinuous on R3∖S and bounded by a constant C(d(p,S),|σV|(S),K,W(V)) depending only on the distance d(p,S), |σV|(S), K, and W(V). Moreover V=v(M,θV) where M={p∈R3∖S|θv(p)≥12}∪S is closed.
Whenever a varifold v(M,θV) satisfies the above assumptions, we will always assume that M={p∈R3∖S|θv(p)≥12}∪S.
These structural properties on curvature varifolds with finite Willmore energy, together with the analogous properties recalled in Remark 2.2, should be always kept in mind in what follows.
The convergence of sets with respect to the Hausdorff distance will play an important role in our study. For every sets X,Y⊂R3 we define the Hausdorff distance dH between X and Y by
dH(X,Y):=inf{ε>0|X⊂Nε(Y),Y⊂Nε(X)}=max{supx∈Xinfy∈Y|x−y|,supy∈Yinfx∈X|x−y|}. | (3.1) |
We say that a sequence of sets Xn converges to a set X in dH if limndH(Xn,X)=0.
Now we prove some useful properties of the Hausdorff distance.
Lemma 3.1. Suppose that Xn→X in dH. Then:
i) Xn→¯X in dH.
ii) If Xn is connected for any sufficiently large n and X is bounded, then ¯X is connected as well.
Proof. ⅰ) Just note that if X⊂Nε2(Xn), then ¯X⊂Nε(Xn).
ⅱ) By i) we can assume without loss of generality that X is closed, and thus compact. Suppose by contradiction that there exist two closed sets A,B⊂X such that A∩B=∅, A≠∅, B≠∅, and A∪B=X. Since X is compact, A and B are compact as well, and thus d(A,B):=infx∈A,y∈B|x−y|=ε>0. By assumption, for any n≥n(ε4) we have that Xn⊂Nε4(X)=Nε4(A)∪Nε4(B) and Nε4(A)∩Nε4(B)=∅. The sets Nε4(A)∩Xn and Nε4(B)∩Xn are disjoint and definitively non-empty, and open in Xn. This implies that Xn is not connected for n large enough, that gives a contradiction.
Lemma 3.2. Suppose Xn is a sequence of uniformly bounded closed sets in R3 and let X⊂R3 be closed. Then Xn→X in dH if and only if the following two properties hold:
a) for any subsequence of points ynk∈Xnk such that ynk→ky, we have that y∈X,
b) for any x∈X there exists a sequence yn∈Xn converging to x.
Proof. Suppose first that dH(Xn,X)→0. If there exists a converging subsequence ynk∈Xnk with limit y∉X, then d(ynk,X)≥ε0>0, and thus Xnk⊄Nε02(X) for k large, that is impossible; so we have proved a). Now let x∈X be fixed. Consider a strictly decreasing sequence εm↘0. For any εm>0 let nεm be such that X⊂Nεm(Xn) for any n≥nεm. This means that Bεm(x)∩Xn≠∅ for any n≥nεm and any m∈N. We can define the sequence
n↦xn∈Xn∩Bεmn(x), |
where
mn=sup{m∈N|Xn∩Bεm(x)≠∅}, |
understanding that xn=x if mn=∞, in fact since Xn is closed we have that x∈Xm if mn=∞. The sequence εmn converges to 0 as n→∞, otherwise there exists η>0 such that Xn∩Bη(x)=∅ for any n large, but this contradicts the convergence in dH. Hence xn→x and we have proved b).
Suppose now that a) and b) hold. If there is ε0>0 such that Xn⊄Nε0(X) for n large, then a subsequence xnk converges to a point y such that d(y,X)≥ε0>0, that is impossible. If there is ε0>0 such that X⊄Nε0(Xn) for n large, then there is a sequence zn∈X such that d(zn,Xn)≥ε0>0. By b) we have that X is bounded, then a subsequence znk converges to z∈X, and d(z,Xnk)≥ε02 definitely in k. But then z is not the limit of any sequence xnk∈Xnk. However z is the limit of a sequence ˉxn∈Xn by b), and thus it is the limit of the subsequence ˉxnk, and this gives a contradiction.
Corollary 3.3. Let Xn be a sequence of uniformly bounded closed sets. Suppose that Xn→X in dH and Xn→Y in dH. If both X and Y are closed, then X=Y.
Proof. Both X and Y are bounded. We can apply Lemma 3.2, that immediately implies that X⊂Y and Y⊂X using the characterization of convergence in dH given by points a) and b).
The above properties allow us to relate the convergence in the sense of varifolds to the convergence of their supports in Hausdorff distance.
Theorem 3.4. Let Vn=v(Mn,θVn)≠0 be a sequence of curvature varifolds with boundary with uniformly bounded Willmore energy converging to V=v(M,θV)≠0. Suppose that the Mn's are connected and uniformly bounded.
Suppose that suppσVn=γ1n∪...∪γαn where the γin's are disjoint compact embedded 1-dimensional manifolds, ˉγ1,...,ˉγβ with β≤α are disjoint compact embedded 1-dimensional manifolds, and assume that γin→ˉγi in dH for i=1,...,β and that H1(γin)→0 for i=β+1,...,α.
Then Mn→M∪ˉγ1∪...∪ˉγβ in Hausdorff distance dH (up to subsequence) and M∪ˉγ1∪...∪ˉγβ is connected. Moreover γin→{pi} in dH for any i=β+1,...,α for some points {pi}, each pi∈M, and suppσV⊂ˉγ1∪...∪ˉγβ∪{pβ+1,...,βα}.
Proof. Let us first observe that by the uniform boundedness of Mn, we get that γin converges to some compact set Xi in dH up to subsequence for any i=β+1,...,α. Each Xi is connected by Lemma 3.1, then by Golab Theorem we know that H1(Xi)≤lim infnH1(γin)=0, hence Xi={pi} for any i=β+1,...,α for some points pβ+1,...,pα. Call X={pβ+1,...,pα}.
By assumption we know that μVn⋆⇀μV as measures on R3, also Mn and M can be taken to be closed. Moreover suppσV⊂X∪ˉγ1∪...∪ˉγβ. In fact Vn are definitely varifolds without generalized boundary on any open set of the form Nε(X∪ˉγ1∪...∪ˉγβ) and they converge as varifolds to V on such an open set with equibounded Willmore energy.
We want to prove that the sets Mn and M∪X∪ˉγ1∪...∪ˉγβ satisfy points a) and b) of Lemma 3.2 and that X⊂M.
Let x∈M∪ˉγ1∪...∪ˉγβ∪X. If x∈ˉγ1∪...∪ˉγβ∪X, then by assumption and Lemma 3.2 there is a sequence of points in suppσVn converging to x. So let x∈M∖(ˉγ1∪...∪ˉγβ∪X). We know that there exists the limit limρ↘0μV(Bρ(x))πρ2≥1, hence we can write that for any ρ∈(0,ρ0) with ρ0<d(x,suppσV) we have that μV(Bρ(x))≥π2ρ2. There exists a sequence ρm↘0 such that limnμVn(Bρm(x))=μV(Bρm(x)) for any m. Hence Mn∩Bρm(x)≠∅ for any m definitely in n. Arguing as in Lemma 3.2 we find a sequence xn∈Mn converging to x, and thus the property b) of Lemma 3.2 is achieved.
For any ε>0 let Aε:=Nε(X∪ˉγ1∪...∪ˉγβ). Let us show that for any ε>0 it occurs that Mn∖Aε converges to (M∪X∪ˉγ1∪...∪ˉγβ)∖Aε=M∖Aε in dH, i.e. we want to check property a) of Lemma 3.2 for such sets.
Once this convergence is established, we get that Mn→M∪X∪ˉγ1∪...∪ˉγβ in dH and we can show that the whole thesis follows. In fact we have that for any ε>0 for any η>0 it holds that
Mn∖Aε⊂Nη(M∪X∪ˉγ1∪...∪ˉγβ∖Aε),(M∪X∪ˉγ1∪...∪ˉγβ)∖Aε⊂Nη(Mn∖Aε), |
for any n≥nε,η. In particular
Mn=Mn∖Aε∪Aε⊂Nη(M∖Aε)∪Aε⊂Nη+2ε(M∪X∪ˉγ1∪...∪ˉγβ), |
M∪X∪ˉγ1∪...∪ˉγβ=(M∪X∪ˉγ1∪...∪ˉγβ)∖Aε∪Aε⊂Nη(Mn∖Aε)∪Aε⊂Nη+2ε(Mn), |
for any n≥nε,η. Setting ε=η we see that for any η>0 it holds that
Mn⊂N3η(M∪X∪ˉγ1∪...∪ˉγβ),(M∪X∪ˉγ1∪...∪ˉγβ)⊂N3η(Mn), |
for any n≥n2η,η. Hence Mn→M∪X∪ˉγ1∪...∪ˉγβ in dH. Therefore M∪X∪ˉγ1∪...∪ˉγβ is closed and connected. Moreover we get that X⊂M, in fact for any pi∈X for any K∈N≥1 by connectedness of Mn we find some subsequence ynk∈Mn∩∂B1K(pi) converging to a point yK∈M∩∂B1K(pi). Since M is closed, passing to the limit K→∞ we see that pi∈M. In particular Mn→M∪ˉγ1∪...∪ˉγβ in dH and the proof is completed.
So we are left to prove that Mn∖Aε converges to (M∪X∪ˉγ1∪...∪ˉγβ)∖Aε=M∖Aε in dH for any fixed ε>0. Consider any converging sequence ynk∈Mnk∖Aε. For simplicity, let us denote yn such sequence. Suppose by contradiction that yn→y but y∉M∪Aε. Since M is closed, there exist ζ>0 such that Bζ(y)∩M=∅ for n large. Since Mn is connected and M≠∅ we can write that ∂Bζ(y)∩Mn≠∅ for any σ∈(ζ4,ζ2) for n large enough. Since yn∉Aε, up to choosing a smaller ζ we can assume that Bζ(y) does not intersect suppσVn for n large. Fix N∈N with N≥2 and consider points
zn,k∈∂B(1+kN)ζ4(y)∩Mn≠∅, |
for any k=1,...,N−1.
The open balls
{B12Nζ4(zn,k)}N−1k=1 |
are pairwise disjoint. Passing to the limit σ↘0, setting ρ=ζ8N, and using Young's inequality in Equation (2.1) evaluated on the varifold Vn at the point p0=zn,k we get that
π≤μVn(Bζ8N(zn,k))(ζ8N)2+14∫Bζ8N(zn,k)|→HVn|2dμVn+1(ζ8N)2∫Bζ8N(zn,k)⟨→HVn,p−zn,k⟩dμVn(p)≤32μVn(Bζ8N(zn,k))(ζ8N)2+34∫Bζ8N(zn,k)|→HVn|2dμVn, | (3.2) |
for any n large and any k=1,...,N−1. Since
lim supnμVn(Bζ8N(zn,k))≤lim supnμVn(¯Bζ2(y))≤μV(B34ζ(y))=0, |
summing over k=1,...,N−1 in (3.2) and passing to the limit n→∞ we get that
π(N−1)≤lim supn34N−1∑k=1∫Bζ8N(znk)|→HVn|2dμVn≤34lim supnW(Vn). |
Since N can be chosen arbitrarily big from the beginning, we get a contradiction with the uniform bound on the Willmore energy of the Vn's.
Hence we have proved that Mn→M∪ˉγ1∪...∪ˉγβ in dH. By Lemma 3.1 we get that M∪ˉγ1∪...∪ˉγβ is connected.
Remark 3.5. Arguing as in the second part of the proof of Theorem 3.4, we get the following useful statement.
Assuming Vn=v(Mn,θVn)≠0 is a sequence of curvature varifolds with boundary with uniformly bounded Willmore energy converging to V=v(M,θV)≠0. Suppose that the Mn's are connected and closed and that M is closed. Suppose that suppσVn is as in Theorem 3.4. If a subsequence ynk∈Mnk converges to y, then y∈M∪ˉγ1∪...∪ˉγβ.
Observe that the supports Mn,M are not necessarily bounded here.
Remark 3.6. The connectedness assumption in Theorem 3.4 is essential. Consider in fact the following example: let Mn=∂B1(0)∪∂B1n(0) and θVn(p)=1 for any p∈Mn. Hence the varifolds v(Mn,θVn) converge to v(∂B1(0),1) as varifolds and they have uniformly bounded energy equal to 8π, but clearly Mn does not converge to ∂B1(0) in dH.
Remark 3.7. The statement of Theorem 3.4 also holds if we assume suppσVn⊂γ1n∪...∪γαn and Mn∪γ1n∪...∪γαn connected. In this case, using the notation of the proof of Theorem 3.4, we have that Mn∪γ1n∪...∪γαn converges to M∪X∪ˉγ1∪...∪ˉγβ in dH and M∪X∪ˉγ1∪...∪ˉγβ is connected.
Now we want to prove the two main Existence Theorems about boundary valued minimization problems on connected varifolds.
Theorem 4.1. Let γ=γ1∪...∪γα be a disjoint union of smooth embedded compact 1-dimensional manifolds with α∈N≥2.
Let
σ0=ν0mH1└γ |
be a vector valued Radon measure, where m:γ→N≥1 and ν0:γ→(Tγ)⊥ are H1-measurable functions with m∈L∞(H1└γ) and |ν0|=1 H1-ae.
Let P be the minimization problem
P:=min{W(V)|V=v(M,θV):σV=σ0,suppV∪γcompact, connected}. | (4.1) |
If infP<4π, then P has minimizers.
Proof. Let Vn=v(Mn,θVn) be a minimizing sequence for the problem P. Call I=infP<4π, and suppose without loss of generality that W(Vn)<4π for any n. For any p0∈Mn∖γ passing to the limits σ→0 and ρ→∞ in the monotonicity formula (2.1) we get
4π≤W(Vn)+2|σ0|(γ)d(p0,γ), |
then
supp0∈Mn∖γd(p0,γ)≤2|σ0|(γ)4π−W(Vn)≤C(σ0,I). |
Hence the sequence Mn is uniformly bounded in R3. Integrating the tangential divergence of the field X(p)=χ(p)(p) where χ(p)=1 for any p∈BR0(0)⊃Mn for any n we get that
2μVn(R3)=∫divTMnXdμVn=−2∫⟨HVn,X⟩dμVn+∫⟨X,ν0⟩d|σ0|≤C(σ0,I)μVn(R3)12+C(σ0,I), |
for any n, and then μVn is uniformly bounded. By the classical compactness theorem for rectifiable varifolds (see Section 5 of Chapter 8 in [30]) we have that Vn→V=v(M,θV) in the sense of varifolds (up to subsequence), and M is compact.
By an argument analogous to the proof of Theorem 3.4 we can show that V≠0. Suppose in fact that V=0. Since α≥2 and the curves γ1,...,γα are disjoint and embedded, there exist a embedded torus ϕ:S1×S1→R3∖γ dividing R3 into two connected components A1,A2 such that A1⊃γ1 and A2⊃γ2∪...∪γα. Since Mn is connected and uniformly bounded, there is a sequence of points yn∈Mn∩ϕ(S1×S1) with a converging subsequence ynk→y. Observe that there is Δ>0 such that d(yn,γ)≥Δ. Since V=0 we have that y∉suppV. Let N≥4 be a natural number and consider the balls {BjNΔ2(y)}Nj=1. Up to subsequence, for n sufficiently large there is zn,j∈∂BjNΔ2(y)∩Mn. Also the balls
{BΔ4N(zn,j)}Nj=1 |
are pairwise disjoint. As in (3.2) we get that
π≤32μVn(BΔ4N(zn,j))(Δ4N)2+34∫BΔ4N(zn,j)|HVn|2dμVn |
for any j=1,...,N. Since lim supnμVn(BΔ4N(zn,j))≤μV(B34Δ(y))=0, summing over j=1,...,N and passing to the limit in n we get
4π≤Nπ≤34limnW(Vn)≤3π, |
that gives a contradiction. Hence Theorem 3.4 implies that suppV∪γ=M∪γ is connected. Since W(V)≤I by lower semicontinuity, we are left to show that σV=σ0.
Since γ is smooth we can write that
|π(Tγ)⊥(p−q0)|≤Cγ|p−q0|2 | (4.2) |
as p→q0 with p∈γ for some constant Cγ depending on the curvature of γ. Let 0<σ<s with s=s(γ) such that (4.2) holds for p∈γ∩Bs(q) for any q∈γ. For any q0∈γ the monotonicity formula (2.1) at q0 on Vn gives
μVn(Bσ(q0))σ2≤−1σ2∫Bσ(q0)⟨HVn,p−q0⟩dμVn(p)−12∫Bσ(q0)(1|p−q0|2−1σ2)⟨p−q0,ν0⟩d|σ0|(p)+limρ→∞AVn(ρ)≤W(Vn)12(μVn(Bσ(q0))σ2)12+12∫Bσ(q0)Cγ|p−q0|2|p−q0|2+1σd|σ0|(p)+π+12∫⟨p−q0,ν0⟩|p−p0|2d|σ0|(p)≤W(Vn)12(μVn(Bσ(q0))σ2)12+Cγ|σ0|(Bσ(q0))+1σ|σ0|(Bσ(q0))+π+121s|σ0|(γ∖Bσ(q))≤C(I)(μVn(Bσ(q0))σ2)12+C(γ,σ0). |
In particular
μVn(Bσ(q))≤C(I,γ,σ0)σ2 | (4.3) |
for any q0∈γ, any σ∈(0,s), and any n.
Consider now any X∈C0c(Br(q0)) for fixed q0∈γ and r∈(0,s). By varifold convergence we have that
limn−2∫⟨HVn,X⟩dμVn+∫⟨X,ν0⟩d|σ0|=−2∫⟨HV,X⟩dμV+∫⟨X,νV⟩d|σV|, | (4.4) |
where we wrote σV=νV|σV|. Now let m∈N be large and consider the cut off function
Λm(p)={1−md(p,γ)d(p,γ)≤1m,0d(p,γ)>1m. | (4.5) |
Take now X=ΛmY for some Y∈C0c(Br(q0)). We have that
lim supm→∞limn|∫⟨HVn,X⟩dμVn|=lim supm→∞limn|∫Br(q0)∩N1m(γ)Λm⟨HVn,Y⟩dμVn|≤‖Y‖∞lim supmlimnW(Vn)12μVn(Br(q0)∩N1m(γ))12. |
Moreover, there exists a constant C(γ) such that Br(q0)∩N1m(γ)⊂∪C(γ)mi=1B2m(qi) for some points qi∈γ and at most C(γ)m balls {B2m(qi)}i. Hence for 2m<s we can estimate
μVn(Br(q0)∩N1m(γ))≤C(γ)m∑i=1μVn(B2m(qi))≤C(γ)mC(I,γ,σ0)4m2. |
Therefore
lim supm→∞limn|∫⟨HVn,X⟩dμVn|≤‖Y‖∞lim supmC(I,γ,σ0)1√m=0. | (4.6) |
Hence setting X=ΛmY in (4.4) and letting m→∞ we obtain
∫⟨Y,ν0⟩d|σ0|=∫⟨Y,νV⟩d|σV|, |
for any Y∈C0c(Br(q0)). Since q0∈γ is arbitrary we conclude that σV=σ0, and thus V is a minimizer.
Theorem 4.2. Let γ=γ1∪...∪γα be a disjoint union of smooth embedded compact 1-dimensional manifolds with α∈N≥2.
Let m:γ→N≥1 by H1-measurable with m∈L∞(H1└γ).
Let Q be the minimization problem
Q:=min{W(V)|V=v(M,θV):|σV|≤mH1└γ,suppV∪γcompact, connected}. | (4.7) |
If infQ<4π, then Q has minimizers.
Proof. We adopt the same notation used in the proof of Theorem 4.1. In this case the generalized boundaries of the minimizing sequence Vn=v(Mn,θVn) are denoted by σVn=νVn|σVn|, and |σVn|≤mH1└γ. The very same strategy used in Theorem 4.1 shows that Vn converges up to subsequence in the sense of varifolds to a limit V=v(M,θV)≠0 with M∪γ compact and connected by Theorem 3.4 and Remark 3.7, and W(V)≤infQ. Hence, to see that V is a minimizer, we are left to show that |σV|≤mH1└γ. Calling μ:=mH1└γ, we find as in Theorem 4.1 that there exist constants C=C(infQ,γ,μ) and s=s(γ) such that
μVn(Bσ(q))≤Cσ2, |
for any q∈γ, any σ∈(0,s), and any n large.
For any X∈C0c(Br(q0)) for fixed q0∈γ and r∈(0,s) the convergence of the first variation of varifolds reads
limn−2∫⟨HVn,X⟩dμVn+∫⟨X,νVn⟩d|σVn|=−2∫⟨HV,X⟩dμV+∫⟨X,νV⟩d|σV|, | (4.8) |
where we wrote σV=νV|σV|. Now we set X=ΛmY in (4.8) for Y∈C0c(Br(q0)) and Λm as in (4.5). Estimating as in (4.6) and taking the limit m→∞ we obtain
limn∫⟨Y,νVn⟩d|σVn|=∫⟨Y,νV⟩d|σV|, |
that is σVn⋆⇀σV, and thus |σV|(A)≤lim infn|σVn|(A)≤μ(A) for any open set A. Hence |σV|≤μ and V is a minimizer of Q.
Remark 4.3. Assuming in the above existence theorems that the connected components of the boundary datum are at least two (i.e., α≥2) is technical, but it is also essential in order to obtain a non-trivial minimization problem, i.e., a problem that does not necessarily reduces to a Plateau's one. In fact if we consider a single closed embedded smooth oriented curve γ, Lemma 34.1 in [30] guarantees the existence of a minimizing integer rectifiable current T=τ(M,θ,ξ) with compact support and with boundary γ. Hence by Lemma 33.2 in [30] the integer rectifiable varifold V=v(M,θ) is stationary and suppσV⊂γ. Then we can take M=suppT, that is compact. Since ∂T=γ and T is minimizing, the set M∪γ is connected and W(V) is trivially zero.
The Existence Theorems 4.1 and 4.2 can be applied in different perturbative regimes, as discussed in the following corollaries and remarks.
Corollary 4.4. Let γ=γ1∪...∪γα be a disjoint union of smooth embedded compact 1-dimensional manifolds with α∈N≥2. Suppose that there exists a compact connected surface Σ⊂R3 with boundary ∂Σ=γ. Let ε∈R and fε:R3→R3 be a smooth family of diffeomorphisms with f0=id|R3. For any ε let
σε=cofε(Σ)H1└(fε(γ)), |
where cofε(Σ) is the conormal field of fε(Σ).
If W(Σ)<4π, there exists ε1>0 such that if ε0<ε1 the minimization problems
Pε:=min{W(V)|V=v(M,θV):σV=σε,suppV∪fε(γ)compact, connected}, | (4.9) |
Qε:=min{W(V)|V=v(M,θV):|σV|≤H1└(fε(γ)),suppV∪fε(γ)compact, connected}, | (4.10) |
have minimizers for any ε∈(−ε0,ε0).
Corollary 4.5. Let γ=γ1∪...∪γα be a disjoint union of smooth embedded compact 1-dimensional manifolds with α∈N≥2. Suppose that there exists a compact connected minimal surface Σ⊂R3 with boundary ∂Σ=γ. Let ε∈R and fε:R3→R3 be a smooth family of diffeomorphisms with f0=id|R3. For any ε let
σε=cofε(Σ)H1└(fε(γ)), |
where cofε(Σ) is the conormal field of fε(Σ).
Then there exists ε1>0 such that if ε0<ε1 the minimization problems
Pε:=min{W(V)|V=v(M,θV):σV=σε,suppV∪fε(γ)compact, connected}, | (4.11) |
Qε:=min{W(V)|V=v(M,θV):|σV|≤H1└(fε(γ)),suppV∪fε(γ)compact, connected}, | (4.12) |
have minimizers for any ε∈(−ε0,ε0).
Remark 4.6. Many examples in which the Existence Theorems 4.1 and 4.2 and Corollary 4.4 apply are given by defining the following boundary data. We can consider any compact smooth surface S without boundary such that W(S)<8π. Then the monotonicity formula (see also [15,16]) implies that S is embedded. We remark that there exist examples of such surfaces having any given genus ([2,31]). Considering any suitable plane π that intersects S in finitely many disjoint compact embedded curves γ1,...,γα, we get that one halfspace determined by π contains a piece Σ of S with W(Σ)<4π and ∂Σ=γ1∪...∪γα. Calling coΣ the conormal field of Σ we get that problems
P:=min{W(V)|V=v(M,θV):σV=coΣH1└∂Σ,suppV∪∂Σ compact, connected }, |
Q:=min{W(V)|V=v(M,θV):|σV|≤H1└∂Σ,suppV∪∂Σ compact, connected }, |
and suitably small perturbations Pε, Qε of them have minimizers.
Remark 4.7. Suppose that γ=γ1∪...∪γα is a disjoint union of compact smooth embedded 1-dimensional manifolds and that γ is contained in some sphere S2R(c). Up to translation let c=0. If there is a point N∈S2R(0) such that for any i the image πN(γi) via the stereographic projection πN:S2R(0)∖{N}→R2 is homotopic to a point in R2∖∪αi=1πN(γi), then the problem
Q:=min{W(V)|V=v(M,θV):|σV|≤H1└γ,suppV∪γ compact, connected }, |
has minimizers. In fact under such assumption there exists a connected submanifold Σ of S2R(0) with ∂Σ=γ, thus W(Σ)<4π and Theorem 4.2 applies.
Remark 4.8. For given R≥1 and h>0 consider the curves
ΓR,h={x2+y2=1,z=h}∪{x2+y2=R2,z=−h}. |
Suppose that h0>0 is the critical value for which a connected minimal surface Σ with ∂Σ=ΓR,h exists if and only if h≤h0. Let Σ0 be a minimal surface with ∂Σ0=ΓR,h0. Applying Corollary 4.5 we get that for ε>0 sufficiently small the minimization problem
Qε:=min{W(V)|V=v(M,θV):|σV|≤H1└ΓR,h0+ε,suppV∪ΓR,h0+ε compact, connected } |
has minimizers.
Let us anticipate that in the case of boundary data of the form ΓR,h we will see in Corollary 6.2 that actually existence of minimizers for the problem Qε is guaranteed for any ε>0, in fact we will see that the hypotheses implying existence of minimizers actually hold for boundary datum ΓR,h for any h>0.
As we recalled in Remark 2.2, it is proved in [15] that the infimum of the Willmore energy on closed surfaces coincide with the infimum taken over non-zero compact varifolds without boundary. First we prove that such infima are both achieved by spheres. This result is certainly expected by experts in the field, but up to the knowledge of the authors it has not been proved yet without appealing to highly non-trivial regularity theorems.
Proposition 5.1. Let V=v(M,θV) be an integer rectifiable varifold with σV=0 and such that suppV is compact. If W(V)=4π, then V=v(S2R(z),1) for some 2-sphere S2R(z)⊂R3.
Proof. Passing to the limits σ→0 and ρ→∞ in the monotonicity formula for varifolds we get that
4πθV(p0)+4∫M|→H2+(p−p0)⊥|p−p0|2|2dμV=4π, |
for any p0∈R3. Hence θV(p0)=1 for any p0∈M, and also
→H(p)=−2(p−p0)⊥|p−p0|2, | (5.1) |
for H2-ae p∈M and for every p0∈M.
Fix δ>0 small and two points p1,p2∈M with p2∉B2δ(p1). For H2-ae p∈M we can write
→H(p)={−2(p−p1)⊥|p−p1|2p∉Bδ(p1),−2(p−p2)⊥|p−p2|2p∉Bδ(p2). |
Since M is bounded, we get that →H∈L∞(μV). Therefore, since θV=1 on M, by the Allard Regularity Theorem ([30]) we get that M is a closed surface of class C1,α for any α∈(0,1).
Since M is closed, it is also compact, and thus it is connected, for otherwise W(V)≥8π.
Let p∈M be any fixed point such that (5.1) holds, and call νp the unit vector such that ν⊥p=TpM. Up to translation let p=0. Consider the axis generated by ν0 and any point p0∈M∖{0}. We can write p0=q+w with q=αν0 and ⟨w,ν0⟩=0. Writing analogously (q+w′)∈M∖{0} another point with the same component on the axis generated by ν0, (5.1) implies that
−2−⟨q,ν0⟩ν0|q|2+|w|2=−2(0−q−w)⊥0|q−w|2=→H(0)=−2(0−q−w′)⊥0|q−w′|2=−2−⟨q,ν0⟩ν0|q|2+|w′|2. |
Hence, whenever q≠0, we have that |w|=|w′|; that is points in M of the form αν0+w with α≠0 and w∈ν⊥0 lie on a circle. It follows that M is invariant under rotations about the axis {tν0|t∈R}.
This argument works at H2-almost any point of M. Therefore we have that for any p∈M, the set M is invariant under rotations about the axis p+{tνp|t∈R}.
Still assuming 0∈M, up to rotation suppose that ν0=(0,0,1). Let a∈M be such that νa=(1,0,0). There exists a point b∈M such that b=tν0=(0,0,t) for some t∈R∖{0}. We can write 0=q+w and b=q+w′ for the same q∈a+{tνa|t∈R} and some w,w′∈ν⊥a. Since |w|=|w′|, it follows that q≠0, otherwise b=0. Since q≠0, the rotation of the origin about the axis a+{tνa|t∈R} implies that M contains a circle C of radius r>0 passing through the origin, and the plane containing C is orthogonal to ν⊥0. Since M is of class C1, the circle C has to be tangent at 0 to the subspace ν⊥0. Thus by invariance with respect to the rotation about the axis {tν0|t∈R}, we have that M contains the sphere with positive radius given by the rotation of C about {tν0|t∈R}. Since the Willmore energy of a sphere is 4π, it follows that M coincide with such sphere.
Now we can prove the above mentioned result on the asymptotic behavior of connected varifolds.
Corollary 5.2. Let Vn=v(Mn,θVn) be a sequence of integer rectifiable curvature varifolds with boundary satisfying the hypotheses of Theorem A.2. Suppose that Mn is compact and connected for any n.
If
W(Vn)≤4π+o(1)as n→∞,diam(suppVn)→n→∞+∞,lim supn|σVn|(R3)diam(suppVn)=0, |
and suppσVn is a disjoint union of uniformly finitely many compact embedded 1-dimensional manifolds, then the sequence
˜Vn:=v(Mndiam(suppVn),˜θn) |
where ˜θn(x)=θVn(diam(suppVn)x), converges up to subsequence and translation to the varifold
V=v(S,1), |
where S is a sphere of diameter 1, in the sense of varifolds and in Hausdorff distance.
Proof. Up to translation let us assume that 0∈suppVn. Then supp˜Vn is uniformly bounded with diam(supp˜Vn)=1. We have that
2μ˜Vn(R3)=∫divT˜Vnpdμ˜Vn(p)≤CW(˜Vn)12(μ˜Vn(R3))12+C|σVn|(R3)diam(suppVn), |
and thus Theorem A.2 implies that ˜Vn converges to a limit varifold V (up to subsequence). Also σ˜Vn⋆⇀σV, and thus |σV|(R3)≤lim infn|σ˜Vn|(R3)≤lim supn|σVn|(R3)diam(suppVn)=0; hence V has compact support and no generalized boundary.
Let us say that suppσ˜Vn is the disjoint union of the smooth closed curves γ1n,...,γαn. By the uniform boundedness of supp˜Vn, we get that γin converges to some compact set Xi in dH up to subsequence. Each Xi is connected by Lemma 3.1, then by Golab Theorem we know that H1(Xi)≤lim infnH1(γin)=0, hence Xi={pi} for any i for some points p1,...,pα, and we can assume that pi≠0 for any i=1,...,α.
Using ideas from the proof of Theorem 3.4, we can show that V≠0.In fact suppose by contradiction that V=0. Fix N∈N with N≥4. By connectedness of Mn, since diam(supp˜Vn)→1, and the boundary curves converge to a discrete sets, for j=1,...,N there are points zn,j∈∂Bj2N(0)∩supp˜Vn for n large. We can also choose N so that d(zn,j,suppσ˜Vn)≥δ(N)>0 for n large. The open balls {B14N(zn,j)}Nj=1 are pairwise disjoint. Using Young inequality as in Theorem 3.4 in the monotonicity formula (2.1) applied on ˜Vn at points zn,j with σ→0 and ρ=14N gives
π≤32μ˜Vn(B14N(zn,j))(14N)2+34∫B14N(zn,j)|H˜Vn|2dμ˜Vn+12|∫(1|p−zn,j|2−1(14N)2)(p−zn,j)dσ˜Vn(p)|, | (5.2) |
for any n and j=1,...,N. Since V=0 we have that lim supnμ˜Vn(B14N(zn,j))≤lim supnμ˜Vn(¯B2(0))=0. Also
|∫(1|p−zn,j|2−1(14N)2)(p−zn,j)dσ˜Vn(p)|≤C(δ(N),N)|σ˜Vn|(R3)→n→∞0. |
Hence summing on j=1,...,N in (5.2) and passing to the limit n→∞ we get
4π≤Nπ≤34limnW(˜Vn)≤3π, |
that gives a contradiction.
Therefore we can apply Theorem 3.4 to conclude that supp˜Vn converges to M in dH. Finally, since V is a compact varifold without generalized boundary and
4π≤W(V)≤lim infnW(Vn)=4π, |
by Proposition 5.1 we conclude that V is a round sphere of multiplicity 1. By Lemma 3.2 the diameter of M is the limit limndiam(supp˜Vn)=1.
In this section we want to discuss how the Existence Theorems 4.1 and 4.2 and the asymptotic behavior described in Corollary 5.2 relate with the remarkable case that motivates our study, namely the immersions in the class FR,h.
First, the monotonicity formula provides the following estimates on immersions φ∈FR,h.
Lemma 6.1. Fix R≥1 and h>0. It holds that:
i)
inf{W(φ)|φ∈FR,h}≤4π4h2+R2−1√(4h2+R2−1)2+16h2<4π. | (6.1) |
ii)
limh→∞inf{W(φ)|φ∈FR,h}=4π. | (6.2) |
Proof. i) We can consider as competitor in FR,h the truncated sphere
Σ=S2√1+(z0−h)2(z0)∩{|z|≤h}, |
where z0=(0,0,1−R24h) is the point on the z-axis located at the same distance from the two connected components of ΓR,h. The surface Σ is contained in another truncated sphere Σ′ having the same center and radius and symmetric with respect to the plane {z=1−R24h}. The boundary of Σ′ is the disjoint union of two circles of radius 1. We have
W(Σ)≤W(Σ′)=4π4h2+R2−1√(4h2+R2−1)2+16h2 |
ii) Let φ∈FR,h and Σ=φ(C). By connectedness there is a point p∈Σ∖∂Σ lying in the plane z=0. Hence dH(Σ,∂Σ)≥h, and by (2.7) we have
4π≤W(Σ)+22π(1+R)h∀Σ. |
Then 4π≤inf{W(φ)|φ∈FR,h}+4π(1+R)h and the thesis follows by using i) by letting h→∞.
We already discussed in Remark 4.8 the existence of minimization problems arising by perturbations of minimal catenoids in some FR,h. By Lemma 6.1 we can complete the picture about existence of optimal connected elastic surfaces with boundary ΓR,h for any R≥1 and h>0, as well as the asymptotic behavior of almost optimal surfaces having such boundaries.
Corollary 6.2. Fix R≥1 and h>0.
1) Then the minimization problem
QR,h:=min{W(V)|V=v(M,θV):|σV|≤H1└ΓR,h,suppV∪ΓR,hcompact, connected} |
has minimizers.
2) Let hk→∞ be any sequence. Let Σk=φk(C) for φk∈FR,hk. Suppose that W(φk)≤4π+o(1) as k→∞. Let Sk=ΣkdiamΣk.
Then (up to subsequence) Sk converges in Hausdorff distance to a sphere S of diameter 1, and the varifolds corresponding to Sk converge to V=v(S,1) in the sense of varifolds.
Proof. 1) The result follows by point i) in Lemma 6.1 by applying Corollary 4.4.
2) Identifying Sk with the varifold it defines, we estimate the total variation of the boundary measure by |∂Sk|≤H1(ΓR,hk)diamΣk. Moreover, by the Gauss-Bonnet Theorem the L2-norm of the second fundamental form of Sk is uniformly bounded. Hence Corollary 5.2 applies and the thesis follows.
Using the notation of point 2) in Corollary 6.2, we remark that even if we know that the rescalings Sk converge to a sphere in dH and as varifolds, it remains open the question whether at a scale of order h the sequence Σk approximate a big sphere. More precisely it seems a delicate issue to understand if diamΣk∼2hk as k→∞.
We conclude with the following partial result: the monotonicity formula gives us some evidence in the case we assume that diamΣkhk→∞.
Proposition 6.3. Let Σk=φk(C) for φk∈FR,hk. Suppose that W(φk)≤4π+o(1) as k→∞. Let Mk=Σkhk.
Then Mk converges up to subsequence to Z=v(M,θZ) in the sense of varifolds. If also
diamΣkhk→∞, |
then M is a plane containing the z-axis and θZ≡1.
Proof. We identify Mk with the varifold it defines. First we can establish the convergence up to subsequence in the sense of varifolds by using Theorem A.2. In fact we have that H1(∂Mk)→0, ∫Mk|IIMk|2 is scaling invariant and thus finite. Moreover, since d(0,∂Mk)≥1, by monotonicity (2.1) we get that
μMk(Bσ(0))σ2≤−1σ2∫Bσ(0)⟨HMk,p⟩dμMk(p)−12∫Bσ(0)∩∂Mk(1|p|2−1σ2)⟨p,coMk(p)⟩dH1(p)+limρ→∞AMk(ρ)≤π+o(1)+1σ2∫Bσ(0)|p||HMk|dμMk(p)+12∫∂Mk∖Bσ(0)dH1(p)|p|+12σ2∫∂Mk∩Bσ(0)|p|dH1(p)≤π+o(1)+1σμMk(Bσ(0))12W(Mk)12+12H1(∂Mk)+12σH1(∂Mk), |
where AMk(⋅) is the monotone quantity centered at 0 evaluated on Mk, and therefore μMk(Bσ(0))≤C(σ) for any σ≥1. Hence the hypotheses of Theorem A.2 are satisfied and we call Z=v(M,θZ) the limit varifold of Mk. Observe that σZ=0 and W(Z)<+∞.
From now on assume that diamΣk/hk→∞. Arguing as in the proof of Corollary 5.2 we can prove that Z≠0. In fact suppose by contradiction that Z=0. Fix N∈N with N≥4. By connectedness of Mk, for j=1,...,N there are points zk,j∈∂BjN(0,0,1)∩Mk and zk,j∉∂Mk for k large. The open balls {B12N(zk,j)}Nj=1 are pairwise disjoint. Hence the monotonicity formula (2.1) applied on Mk at points zk,j with σ→0 and ρ=12N gives
π≤32μMk(B12N(zk,j))(12N)2+34∫B12N(zk,j)|HMk|2dμMk, | (6.3) |
for any k and j=1,...,N. Since Z=0 we have that
lim supkμMk(B12N(zk,j))≤lim supkμMk(B2(0,0,1))=0. |
Hence, summing on j=1,...,N in (6.3) and passing to the limit k→∞ we get
4π≤Nπ≤34limkW(Mk)≤3π, |
that gives a contradiction.
Also the support of Z is unbounded. In fact suppose by contradiction that suppZ⊂⊂BR(0), and thus M is closed by Proposition. Since Mk is connected, there exists q′k∈Mk∩∂B2R(0) definitely in k for R sufficiently big. Up to subsequence q′k→q′. By Remark 3.5 we get that q′∈suppZ, that contradicts the absurd hypothesis.
Since M is unbounded, by Corollary (or equivalently (A.22) in [15]) we know that
limρ→∞μZ(Bρ(q))ρ2≥π. |
By construction
limk∫Bσ(0)∩∂Mk⟨p|p|2,coMk⟩dH1(p)=0, |
hence passing to the limit k→∞ in the monotonicity formula (2.1) evaluated on Mk we get that
AZ(σ)≤lim infkAMk(σ), |
for ae σ>0. By monotonicity
AZ(σ)≤lim infklimσ→∞AMk(σ)≤lim infkW(Mk)4+H1(∂Mk)≤π. |
On the other hand, by (A.14) in [15] we can write that
limσ→∞AZ(σ)=14W(Z)+limσ→∞μZ(Bσ(q))σ2≥14W(Z)+π. |
Hence Z is stationary, limρ→∞μZ(Bρ(q))ρ2=π, and M is closed.
If p0 is any point in M, the monotonicity formula for Z centered at p0 reads
μZ(Bσ(p0))σ2+∫Bρ(p0)∖Bσ(p0)|(p−p0)⊥|2|p−p0|4=μZ(Bρ(q))ρ2. | (6.4) |
In particular θZ(p0)=1, and thus we can apply Allard Regularity Theorem at p0. Thus we get that M is of class C∞ around p0 (and analogously everywhere), and thus there exists the limit
limσ→0∫Bρ(p0)∖Bσ(p0)|(p−p0)⊥|2|p−p0|4=∫Bρ(p0)|(p−p0)⊥|2|p−p0|4. |
Passing to the limits ρ→∞ and σ↘0 in (6.4), we get that
limρ→∞∫Bρ(p0)|(p−p0)⊥|2|p−p0|4=0. |
Therefore |(p−p0)⊥|=0 for any p∈M, where we recall that (⋅)⊥ is the orthogonal projection on TpM⊥. Since this is true for any p0∈M, we derive that M is a plane. Finally Remark 3.5 implies that M contains the vertical axis {(0,0,t)|t∈R}.
The authors declare no conflict of interest.
In this appendix we recall the definitions and the results about curvature varifolds with boundary that we need throughout the whole work. This section is based on [18] (see also [14,30]).
Let Ω⊂Rk be an open set, and let 1<n≤k. We identify a n-dimensional vector subspace P of Rk with the k×k-matrix {Pij} associated to the orthogonal projection over the subspace P. Hence the Grassmannian Gn,k of n-spaces in Rk is endowed with the Frobenius metric of the corresponding projection matrices. Moreover given a subset A⊂Rk, we define Gn(A)=A×Gn,k, endowed with the product topology. A general n-varifold V in an open set Ω⊂Rk is a non-negative Radon measure on Gn(Ω). The varifold convergence is the weak* convergence of Radon measures on Gn(Ω), defined by duality with C0c(Gn(Ω)) functions.
We denote by π:Gn(Ω)→Ω the natural projection, and by μV=π♯(V) the push forward of a varifold V onto Ω. The measure μV is called induced (weight) measure in Ω.
Given a couple (M,θ) where M⊂Ω is countably n-rectifiable and θ:M→N≥1 is Hn-measurable, the symbol v(M,θ) defines the (integer) rectifiable varifold given by
∫Gn(Ω)φ(x,P)dv(M,θ)(x,P)=∫Mφ(x,TxM)θ(x)dHn(x), | (A.1) |
where TxM is the generalized tangent space of M at x (which exists Hn-ae since M is rectifiable). The function θ is called density or multiplicity of v(M,θ). Note that μV=θHn└M in such a case.
From now on we will always understand that a varifold V is an integer rectifiable one.
We say that a function →H∈L1loc(μV;Rk) is the generalized mean curvature of V=v(M,θ) and σV Radon Rk-valued measure on Ω is its generalized boundary if
∫divTMXdμV=−n∫⟨→H,X⟩dμV+∫XdσV, | (A.2) |
for any X∈C1c(Ω;Rk), where divTMX(p) is the Hn-ae defined tangential divergence of X on the tangent space of M. Recall that σV has the form σV=νVσ, where |νV|=1 σ-ae and σ is singular with respect to μV.
If V has generalized mean curvature →H, the Willmore energy of V is defined to be
W(V)=∫|H|2dμV. | (A.3) |
The operator X↦δV(X):=∫divTMXdμV is called first variation of V. Observe that for any X∈C1c(Ω;Rk), the function φ(x,P):=divP(X)(x)=tr(P∇X(x)) is continuous on Gn(Ω). Hence, if Vn→V in the sense of varifolds, then δVn(X)→δV(X).
By analogy with integration formulas classically known in the context of submanifolds, we say that a varifold V=v(M,θ) is a curvature n-varifold with boundary in Ω if there exist functions Aijk∈L1loc(V) and a Radon Rk-valued measure ∂V on Gn(Ω) such that
∫Gn(Ω)Pij∂xjφ(x,P)+Aijk(x,P)∂Pjkφ(x,P)dV(x,P)==n∫Gn(Ω)φ(x,P)Ajij(x,P)dV(x,P)+∫Gn(Ω)φ(x,P)d∂Vi(x,P), | (A.4) |
for any i=1,...,k for any φ∈C1c(Gn(Ω)). The rough idea is that the term on the left is the integral of a tangential divergence, while on the right we have integration against a mean curvature plus a boundary term. The measure ∂V is called boundary measure of V.
Theorem A.1 ([18]). Let V=v(M,θ) be a curvature varifold with boundary on Ω. Then the following hold true.
i) Aijk=Aikj, Aijj=0, and Aijk=PjrAirk+PrkAijr=PjrAikr+PkrAijr.
ii) Pil∂Vl(x,P)=∂Vi(x,P) as measures on Gn(Ω).
iii) PilAljk=Aijk.
iv) Hi(x,P):=1nAjij(x,P) satisfies that PilHl(x,P)=0 for V-ae (x,P)∈Gn(Ω).
v) V has generalized mean curvature →H with components Hi(x,TxM) and generalized boundary σV=π♯(∂V).
We call the functions IIkij(x):=PilAjkl components of the generalized second fundamental form of a curvature varifold V. Observe that IIkjj=PjlAjlk=Ajjk−PklAjjl=Ajkj−PklAjlj=nHk−nPklHl=nHk, and Aijk=IIkij+IIjki.
In conclusion we state the compactness theorem that we use in this work.
Theorem A.2 ([18]). Let p>1 and Vl a sequence of curvature varifolds with boundary in Ω. Call A(l)ijk the functions Aijk of Vl. Suppose that A(l)ijk∈Lp(V) and
supl{μVl(W)+∫Gn(W)|∑i,j,k|A(l)ijk||pdVl+|∂Vl|(Gn(W))}≤C(W)<+∞ | (A.5) |
for any W⊂⊂Gn(Ω), where |∂Vl| is the total variation measure of ∂Vl. Then:
i) up to subsequence Vl converges to a curvature varifold with boundary V in the sense of varifolds. Moreover A(l)ijkVl→AijkV and ∂Vl→∂V weakly* as measures on Gn(Ω);
ii) for every lower semicontinuous function f:Rk3→[0,+∞] it holds that
∫Gn(Ω)f(Aijk)dV≤lim infl∫Gn(Ω)f(A(l)ijk)dVl. | (A.6) |
It follows from the above theorem that the Willmore energy is lower semicontinuous with respect to varifold convergence of curvature varifolds with boundary satisfying the hypotheses of Theorem A.2.
The monotonicity formula on varifolds with locally bounded first variation is a fundamental identity proved in [31], with important consequences on the structure of varifolds with bounded Willmore energy, collected for example in [15]. Such consequences usually concern varifolds without generalized boundary: σV=0. So, in this section we are interested in extending some of these results in the case of curvature varifold with boundary. The strategy is analogous to the one of [15] and the following results are probably expected by the experts in the field, however we prove them here for the convenience of the reader.
Let V=v(M,θV) be a 2-dimensional curvature varifold with boundary with finite Willmore energy. Denote by σV the generalized boundary. Let 0<σ<ρ and p0∈R3. Integrating the tangential divergence of the field X(p)=(1|p−p0|2σ−1ρ2)+(p−p0), where |p−p0|2σ=max{σ2,|p−p0|2}, with respect to the measure μV (see also [31] and [24]) one gets that
A(σ)+∫Bρ(p0)∖Bσ(p0)|→H2+(p−p0)⊥|p−p0|2|2dμV(p)=A(ρ), | (B.1) |
where
A(ρ):=μV(Bρ(p0))ρ2+14∫Bρ(p0)|H|2dμV(p)+Rp0,ρ, | (B.2) |
and
Rp0,ρ:=∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)+12∫Bρ(p0)(1|p−p0|2−1ρ2)(p−p0)dσV(p)=:∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)+Tp0,ρ. | (B.3) |
In particular the function ρ↦A(ρ) is non-decreasing.
From now on, let us assume that the support suppσV⊂S, where S is compact and |σV|(S)<+∞. We also assume that
lim supR→∞μV(BR(0))R2≤K<+∞. |
We have that
|∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)|≤(μV(Bρ(p0))ρ2)12(∫Bρ(p0)|H|2dμV)12≤ε2μV(Bρ(p0))ρ2+2ε∫Bρ(p0)|H|2dμV. | (B.4) |
If d(p0,S)≥δ we have that
|∫Bρ(p0)(1|p−p0|2−1ρ2)(p−p0)dσV(p)|≤(1δ+1ρ)|σV|(S∩Bρ(p0)). | (B.5) |
In particular the monotone function A(ρ) evaluated at p0∉S is bounded below and there exists finite the limit limρ↘0A(ρ).
Keeping p0∉S (B.1) implies that
μV(Bσ(p0))σ2≤μV(Bρ(p0))ρ2+14∫Bρ(p0)|H|2dμV(p)+Rp0,ρ−Rp0,σ≤μV(Bρ(p0))ρ2+14W(V)+(μV(Bρ(p0))ρ2)12W(V)12−Tp0,σ+(1δ+1ρ)|σV|(S∩Bρ(p0))+ε2μV(Bσ(p0))σ2+2εW(V) | (B.6) |
Letting ρ→∞ and σ<δ in (B.6) we get that Tp0,σ=0 and
μV(Bσ(p0))σ2≤C(δ,K,W(V))<+∞∀0<σ<δ, | (B.7) |
Letting ρ→0 in (B.4) and using (B.7) we get that
limρ→0|∫Bρ(p0)⟨→H,p−p0⟩ρ2dμV(p)|=0. | (B.8) |
Therefore we see that if p0∈R3∖S, then
∃limσ↘0μV(Bσ(p0))σ2=πθV(p0)≤C(δ,|σV|(S),K,W(V)). | (B.9) |
Moreover, consider p0∈R3∖S and a sequence pk→p0; let ρ∈(0,d(p0,S)/2) and call ρ0=d(p0,S)/2, then by (B.1) we have that
μV(¯Bρ(p0))ρ2≥lim supkμV(Bρ(pk))ρ2≥lim supkπθV(pk)−Rpk,ρ−14∫Bρ(pk)|H|2dμV≥lim supkπθV(pk)−∫B2ρ(p0)|H|ρdμV−14∫B2ρ(pk)|H|2dμV≥lim supkπθV(pk)−(μV(B2ρ(p0))ρ2)12(∫B2ρ(p0)|H|2dμV)12−14∫B2ρ(pk)|H|2dμV≥lim supkπθV(pk)−(C(2ρ0,|σV|(S),K,W(V))+14)(∫B2ρ(p0)|H|2dμV)12, | (B.10) |
and thus letting ρ↘0 suitably we get
θV(p0)≥lim supkθV(pk), | (B.11) |
i.e., the multiplicity function θV is upper semicontinuous on R3∖S. Since θV is integer valued, the set {p∈R3∖S|θv(p)≥12} is closed in R3∖S. Therefore we can take the closed set M={p∈R3∖S|θv(p)≥12}∪S as the support of V.
A particular case of our analysis can be summarized in the following statement.
Proposition B.1. Let V be a 2-dimensional integer rectifiable curvature varifold with boundary. Denote by σV the generalized boundary and by S a compact set containing the support suppσV. Assume that
W(V)<+∞,lim supR→∞μV(BR(0))R2≤K<+∞, |
and S is a compact 1-dimensional manifold with H1(S)<+∞. Then the limit
limρ↘0μV(Bρ(p))ρ2 |
exists at any point p∈R3∖S, the multiplicity function θV(p)=limρ↘0μV(Bρ(p))ρ2 is upper semicontinuous on R3∖S and bounded by a constant C(d(p,S),|σV|(S),K,W(V)) depending only on the distance d(p,S), |σV|(S), K and W(V). Moreover V=v(M,θV) where M={p∈R3∖S|θv(p)≥12}∪S is closed.
Also, we can derive the following consequence.
Corollary B.2. Let V=v(M,θV) be a 2-dimensional integer rectifiable curvature varifold with boundary with W(V)<+∞. Denote by σV the generalized boundary and by S a compact set containing the support suppσV. Assume that S is a compact 1-dimensional manifold with H1(S)<+∞. Then
Mess. unbounded⇔lim supρ→∞μV(Bρ(0))ρ2≥π, | (B.12) |
where M essentially unbounded means that for every R>0 there is Br(x)⊂R3∖BR(0) such that μV(Br(x))>0.
Moreover, in any of the above cases the limit limρ→∞μV(Bρ(0))ρ2≥π exists.
Proof. Suppose that M is essentially unbounded. We can assume that lim supρ→∞μV(Bρ(0))ρ2≤K<+∞. Then
|∫Bρ(0)1ρ2⟨→H,p⟩dμV|≤1ρ2(∫Bσ(0)|H||p|dμV(p)+∫Bρ(0)∖Bσ(0)|H||p|dμV(p))≤σρ2√∫Bσ(0)|H|2dμV√μV(Bσ(0))+√μV(Bρ(0))ρ2√∫Bρ(0)∖Bσ(0)|H|2dμV |
for any 0 < \sigma < \rho < +\infty . Passing to the \limsup_{ \rho\to\infty} and then to \sigma\to\infty , we conclude that
\lim\limits_{ \rho\to\infty} \left| \int_{B_ \rho(0)} \frac{1}{ \rho^2} \langle \vec{H}, p \rangle \, d\mu_V \right| = 0. |
Hence, assuming without loss of generality that 0\not\in S , the monotone quantity A(\rho) evaluated on V with base point 0 gives
\exists\, \lim\limits_{ \rho\to\infty} A( \rho) = \mathcal{W}(V) + \frac12 \int \frac{p}{|p|^2}\, d \sigma_V(p) + \limsup\limits_{ \rho\to\infty} \frac{\mu_V(B_ \rho(0))}{ \rho^2}, |
and thus \exists\, \lim_{ \rho\to\infty} \frac{\mu_V(B_ \rho(0))}{ \rho^2} \le K < +\infty . Also the assumptions of Proposition B.1 are satisfied and we can assume that M is closed.
We can prove that M has at least one unbounded connected component. In fact any compact connected component N of M defines a varifold \mathbf v(N, \theta_V|_N) with generalized mean curvature; now if S\cap N = \emptyset then \mathcal{W}(N)\ge4\pi , and thus there are finitely many compact connected components without boundary, if instead S\cap N\neq\emptyset , S \subset B_{R_0}(0) by compactness, and \exists\, p_0\in N \setminus B_r(0) for r > R_0 but N is compact, then the monotonicity formula applied on \mathbf v(N, \theta_V|_N) at point p_0 gives
\begin{equation} \pi\le \lim\limits_{ \sigma\to0} A_{ \mathbf v(N, \theta_V|_N)}( \sigma) \le \lim\limits_{ \rho\to\infty} A_{ \mathbf v(N, \theta_V|_N)}( \rho) \le \frac14 \mathcal{W}( \mathbf v(N, \theta_V|_N)) +\frac12 \frac{| \sigma_V|(S)}{r-R_0}. \end{equation} | (B.13) |
Since M is essentially unbounded, if any connected component of M is compact we would find infinitely many compact connected components N , points p_0\in N , and r arbitrarily big in (B.13) so that the Willmore energy of any such N is greater than 2\pi , implying that \mathcal{W}(V) = +\infty .
As M has a connected unbounded component, for any \rho sufficiently large there is x_ \rho\in M\cap B_{2 \rho}(0) . Applying the monotonicity formula on V at x_ \rho for \rho sufficiently big so that S \subset B_ \rho(0) we get that
\begin{split} \pi \le \lim\limits_{ \sigma\to0} A( \sigma) &\le \frac{\mu_V(B_ \rho(x_ \rho))}{ \rho^2} + \frac14 \int_{B_ \rho(x_ \rho)} |H|^2\, d\mu_V + \frac{1}{ \rho}\int_{B_ \rho(x_ \rho)} |H|\, d\mu_V \\ &\le 9\frac{\mu_V(B_{3 \rho}(0))}{(3 \rho)^2} +\frac14 \int_{ \mathbb{R}^3 \setminus B_ \rho(0)} |H|^2\, d\mu_V + \varepsilon \frac{\mu_V(B_ \rho(x_ \rho))}{ \rho^2} + C_ \varepsilon \int_{B_ \rho(x_ \rho)} |H|^2\, d\mu_V , \end{split} |
that implies that
\lim\limits_{ \rho\to\infty} \frac{\mu_V(B_ \rho(0))}{ \rho^2} \ge \frac{\pi}{9+ \varepsilon}, |
for any \varepsilon > 0 .
Consider now any sequence R_n\to\infty and the sequence of blow-in varifolds given by
V_n = \mathbf v\left(\frac{M}{R_n}, \theta_n\right), |
where \theta_n(x) = \theta_V(R_nx) . Since
\mu_{V_n}(B_R(0)) = \frac{1}{R_n^2}\mu_V(B_{R_nR}(0)) = \frac{1}{(RR_n)^2}\mu_V(B_{RR_n}(0))R^2\le K'R^2 |
is bounded for any R > 0 , \mathcal{W}(V_n) = \mathcal{W}(V) , and | \sigma_{V_n}|(\mathbb{R}^3)\to0 , by the classical compactness theorem of rectifiable varifolds (Theorem 42.7 in [30]) we get that V_n converges to an integer rectifiable varifold W (up to subsequence). Also W\neq0 , in fact 0\in \mbox{supp} W by the fact that
\mu_W(\overline{B_1(0)}) \ge \liminf\limits_n \mu_{V_n} (B_1(0)) = \liminf\limits_n \frac{\mu_V(B_{R_n}(0))}{R_n^2}\ge \frac\pi9. |
We have that W is stationary, in fact for any r > 0 we have that
\begin{split} \int_{ \mathbb{R}^3 \setminus\overline{B_r(0)}} |H_W|^2\, d\mu_W \le \liminf\limits_n \int_{ \mathbb{R}^3 \setminus\overline{B_r(0)}} |H_{V_n}|^2\, d\mu_{V_n} = \liminf\limits_n \int_{ \mathbb{R}^3 \setminus\overline{B_{R_nr}(0)}} |H_V|^2\, d\mu_V = 0. \end{split} |
Also \sigma_W = 0 , in fact for any X\in C^0_c(\mathbb{R}^3) the convergence of the first variation reads
\lim\limits_n -2 \int \langle H_{V_n}, X \rangle\, d\mu_{V_n} + \int X\, d \sigma_{V_n} = \lim\limits_n -2 \int \langle H_{V_n}, X \rangle\, d\mu_{V_n} = \int X\, d \sigma_V, |
and \mbox{supp} \sigma_V \subset\{0\} . Taking X = \Lambda_m Y for Y\in C^0_c(\mathbb{R}^3) and
\Lambda_m(p) = \begin{cases} 1-md(p, 0) & d(p, 0)\le\frac1m, \\ 0 & d(p, 0) \gt \frac1m, \end{cases} |
we see that
\left| \int \langle H_{V_n}, X \rangle\, d\mu_{V_n} \right| = \left| \int_{B_{\frac1m}(0)} \langle H_{V_n} , \Lambda_m Y \rangle \, d\mu_{V_n} \right| \le \|Y\|_\infty \mathcal{W}(V)^{\frac12}\left(K'\frac{1}{m^2}\right)^{\frac12}, |
and thus
\int Y\, d \sigma_V = \lim\limits_n -2 \int \langle H_{V_n}, \Lambda_m Y \rangle\, d\mu_{V_n} = \lim\limits_{m\to\infty} \lim\limits_n -2 \int \langle H_{V_n}, \Lambda_m Y \rangle\, d\mu_{V_n} = 0, |
for any Y\in C^0_c(\mathbb{R}^3) .
Finally the monotonicity formula applied on W gives
\lim\limits_n \frac{\mu_V(R_n(0))}{R_n^2}\ge\liminf\limits_n \mu_{V_n}(B_1(0))\ge\mu_W(B_1(0))\ge \lim\limits_{ \sigma\to0} A_W( \sigma) \ge \pi. |
[1] |
Alessandroni R, Kuwert E (2016) Local solutions to a free boundary problem for the Willmore functional. Calc Var Partial Dif 55: 1-29. doi: 10.1007/s00526-015-0942-y
![]() |
[2] | Bauer M, Kuwert E (2003) Existence of minimizing Willmore surfaces of prescribed genus. Int Math Res Notices 10: 553-576. |
[3] |
Bergner M, Dall'Acqua A, Fröhlich S (2010) Symmetric Willmore surfaces of revolution satisfying natural boundary conditions. Calc Var Partial Dif 39: 361-378. doi: 10.1007/s00526-010-0313-7
![]() |
[4] |
Bergner M, Dall'Acqua A, Fröhlich S (2013) Willmore surfaces of revolution with two prescribed boundary circles. J Geom Anal 23: 283-302. doi: 10.1007/s12220-011-9248-2
![]() |
[5] |
Bergner M, Jakob R (2014) Sufficient conditions for Willmore immersions in \mathbb{R}^3 to be minimal surfaces. Ann Glob Anal Geom 45: 129-146. doi: 10.1007/s10455-013-9391-z
![]() |
[6] | Dall'Acqua A, Deckelnick K, Grunau H (2008) Classical solutions to the Dirichlet problem for Willmore surfaces of revolution. Adv Calc Var 1: 379-397. |
[7] |
Dall'Acqua A, Deckelnick K, Wheeler G (2013) Unstable Willmore surfaces of revolution subject to natural boundary conditions. Calc Var Partial Dif 48: 293-313. doi: 10.1007/s00526-012-0551-y
![]() |
[8] |
Dall'Acqua A, Fröhlich S, Grunau H, et al. (2011) Symmetric Willmore surfaces of revolution satisfying arbitrary Dirichlet boundary data. Adv Calc Var 4: 1-81. doi: 10.1515/acv.2010.022
![]() |
[9] | Deckelnick K, Grunau H (2009) A Navier boundary value problem for Willmore surfaces of revolution. Analysis 29: 229-258. |
[10] |
Eichmann S (2016) Nonuniqueness for Willmore surfaces of revolution satisfying Dirichlet boundary data. J Geom Anal 26: 2563-2590. doi: 10.1007/s12220-015-9639-x
![]() |
[11] |
Eichmann S (2019) The Helfrich boundary value problem. Calc Var Partial Dif 58: 1-26. doi: 10.1007/s00526-018-1462-3
![]() |
[12] |
Elliott CM, Fritz H, Hobbs G (2017) Small deformations of Helfrich energy minimising surfaces with applications to biomembranes. Math Mod Meth Appl Sci 27: 1547-1586. doi: 10.1142/S0218202517500269
![]() |
[13] | Gazzola F, Grunau H, Sweers G (2010) Polyharmonic boundary value problems. Lect Notes Math 1991: xviii+423. |
[14] |
Hutchinson J (1986) Second fundamental form for varifolds and the existence of surfaces minimizing curvature. Indiana U Math J 35: 45-71. doi: 10.1512/iumj.1986.35.35003
![]() |
[15] |
Kuwert E, Schätzle R (2004) Removability of point singularities of Willmore surfaces. Ann Math 160: 315-357. doi: 10.4007/annals.2004.160.315
![]() |
[16] |
Li P, Yau ST (1982) A new conformal invariant and its applications to the Willmore conjecture and the first eigenvalue on compact surfaces. Invent Math 69: 269-291. doi: 10.1007/BF01399507
![]() |
[17] |
Mandel R (2018) Explicit formulas, symmetry and symmetry breaking for Willmore surfaces of revolution. Ann Glob Anal Geom 54: 187-236. doi: 10.1007/s10455-018-9598-0
![]() |
[18] |
Mantegazza C (1996) Curvature varifolds with boundary. J Differ Geom 43: 807-843. doi: 10.4310/jdg/1214458533
![]() |
[19] |
Marques FC, Neves A (2014) Min-Max theory and the Willmore Conjecture. Ann Math 179: 683-782. doi: 10.4007/annals.2014.179.2.6
![]() |
[20] | Morgan F (2008) Geometric Measure Theory: A Beginners's Guide, 4 Eds., Academic Press. |
[21] | Pozzetta M (2017) Confined Willmore energy and the Area functional. arXiv:1710.07133. |
[22] | Pozzetta M (2018) On the Plateau-Douglas problem for the Willmore energy of surfaces with planar boundary curves. arXiv:1810.07662. |
[23] |
Rivière T (2008) Analysis aspects of Willmore surfaces. Invent Math 174: 1-45. doi: 10.1007/s00222-008-0129-7
![]() |
[24] |
Rivière T (2013) Lipschitz conformal immersions from degenerating Riemann surfaces with L2-bounded second fundamental forms. Adv Calc Var 6: 1-31. doi: 10.1515/acv-2012-0108
![]() |
[25] | Rivière T (2014) Variational principles for immersed surfaces with L2-bounded second fundamental form. J Reine Angew Math 695: 41-98. |
[26] |
Schätzle R (2010) The Willmore boundary problem. Calc Var 37: 275-302. doi: 10.1007/s00526-009-0244-3
![]() |
[27] |
Schoen R (1983) Uniqueness, symmetry, and embeddedness of minimal surfaces. J. Differ Geom 18: 791-809. doi: 10.4310/jdg/1214438183
![]() |
[28] |
Schygulla J (2012) Willmore minimizers with prescribed isoperimetric ratio. Arch Ration Mech Anal 203: 901-941. doi: 10.1007/s00205-011-0465-4
![]() |
[29] |
Seguin B, Fried E (2014) Microphysical derivation of the Canham-Helfrich free-energy density. J Math Biol 68: 647-665. doi: 10.1007/s00285-013-0647-9
![]() |
[30] | Simon L (1984) Lectures on Geometric Measure Theory, Proceedings of the Centre for Mathematical Analysis of Australian National University. |
[31] |
Simon L (1993) Existence of surfaces minimizing the Willmore functional. Commun Anal Geom 1: 281-326. doi: 10.4310/CAG.1993.v1.n2.a4
![]() |
[32] | Willmore TJ (1965) Note on embedded surfaces. Ann Al Cuza Univ Sect I 11B: 493-496. |
[33] | Willmore TJ (1993) Riemannian Geometry, Oxford Science Publications. |
1. | Klaus Deckelnick, Marco Doemeland, Hans-Christoph Grunau, Boundary value problems for a special Helfrich functional for surfaces of revolution: existence and asymptotic behaviour, 2021, 60, 0944-2669, 10.1007/s00526-020-01875-6 | |
2. | L. De Luca, M. Ponsiglione, Variational models in elasticity, 2021, 3, 2640-3501, 1, 10.3934/mine.2021015 | |
3. | Marco Pozzetta, On the Plateau–Douglas problem for the Willmore energy of surfaces with planar boundary curves, 2021, 27, 1292-8119, S2, 10.1051/cocv/2020049 | |
4. | Anthony Gruber, Magdalena Toda, Hung Tran, Stationary surfaces with boundaries, 2022, 62, 0232-704X, 305, 10.1007/s10455-022-09850-4 | |
5. | Manuel Schlierf, On the convergence of the Willmore flow with Dirichlet boundary conditions, 2024, 241, 0362546X, 113475, 10.1016/j.na.2023.113475 | |
6. | Anna Kubin, Luca Lussardi, Marco Morandotti, Direct Minimization of the Canham–Helfrich Energy on Generalized Gauss Graphs, 2024, 34, 1050-6926, 10.1007/s12220-024-01564-2 | |
7. | Manuel Schlierf, Global existence for the Willmore flow with boundary via Simon’s Li–Yau inequality, 2025, 1864-8258, 10.1515/acv-2024-0018 |