433
Views
1
CrossRef citations to date
0
Altmetric
Research Articles

The Willmore flow with prescribed isoperimetric ratio

ORCID Icon
Pages 148-184 | Received 03 Mar 2023, Accepted 03 Jan 2024, Published online: 17 Jan 2024

Abstract

We introduce a non-local L2-gradient flow for the Willmore energy of immersed surfaces which preserves the isoperimetric ratio. For spherical initial data with energy below an explicit threshold, we show long-time existence and convergence to a Helfrich immersion. This is in sharp contrast to the locally constrained flow, where finite time singularities occur.

2020 Mathematics Subject Classification:

1 Introduction and main results

Finding the shape which encloses the maximal volume among surfaces of prescribed area is certainly one of the oldest and yet most prominent problems in mathematics and goes back to the legend of the foundation of Carthage. Since then generations of mathematicians have been studying isoperimetric problems, aiming to find the best possible shape in all kinds of settings. It turns out that—by the isoperimetric inequality—the optimal configuration in Euclidean space is given by a round sphere.

Likewise, the round spheres are the absolute minimizers for the Willmore energy, a functional measuring the bending of an immersed surface with various applications also beyond geometry, for instance in the study of biological membranes [Citation1, Citation2], general relativity [Citation3], nonlinear elasticity [Citation4] and image restoration [Citation5].

Note that the round spheres describe the optimal shape in both situations. In this article, we will study their relation using a gradient flow approach.

For an immersion f:ΣR3 of a closed oriented surface Σ, its Willmore energy is defined by W(f):=14Σ|H|2dμ.

Here μ=μf denotes the area measure induced by the pull-back of the Euclidean metric gf:=f*·,·, and H=Hf:=Hf,νf denotes the (scalar) mean curvature with respect to ν=νf:ΣS2, the unique unit normal along f induced by the chosen orientation on Σ, see Equation(2.1). A related quantity is the umbilic Willmore energy, given by W0(f):=Σ|A0|2dμ, where A0 denotes the trace-free part of the second fundamental form. As a consequence of the Gauss–Bonnet theorem, these two energies are equivalent from a variational point of view, since for a surface with fixed genus g, we have (1.1) W0(f)=2W(f)8π+8πg.(1.1)

Both energies are not only geometric, i.e. invariant under diffeomorphisms on Σ, but—remarkably—also conformally invariant, i.e. invariant with respect to smooth Möbius transformations of R3. By [Citation6, Theorem 7.2.2], we have W(f)4π with equality if and only if Σ=S2 and f:S2R3 parameterizes a round sphere.

The isoperimetric ratio of an immersion f:ΣR3 is defined as the quotient (1.2) I(f):=36πV(f)2A(f)3,where A(f):=ΣdμandV(f):=13Σf,νdμ(1.2) denote the area and the algebraic volume enclosed by f(Σ), respectively. Here, the normalizing constant is chosen such that by the isoperimetric inequality we always have I(f)=:σ[0,1] with σ = 1 if and only if Σ=S2 and f:S2R3 parameterizes a round sphere. Critical points of the isoperimetric ratio—or equivalently, critical points of the volume functional with prescribed area—are precisely the CMC-surfaces, i.e. the surfaces with constant mean curvature, which form an important generalization of minimal surfaces and naturally arise in the modeling of soap bubbles.

The problem of minimizing the Willmore energy among all immersions of a genus g surface Σg with prescribed isoperimetric ratio, i.e. the minimization problem (1.3) βg(σ):=inf{W(f)|f:ΣgR3 immersion with I(f)=σ},(1.3) naturally arises in mathematical biology in the Canham–Helfrich model [Citation1, Citation2] with zero spontaneous curvature and has already been studied mathematically in [Citation7–9]. While the genus zero case was solved in [Citation7], the results in [Citation8, Citation9] combined with recent findings in [Citation10] and [Citation11] show that the infimum in Equation(1.3) is always attained for any gN0 and σ(0,1); and satisfies βg(σ)<8π. The energy threshold 8π also plays an important role in the analysis of the Willmore energy, since by the famous Li–Yau inequality [Citation12], any immersion f of a compact surface with W(f)<8π has to be embedded.

A sufficiently smooth minimizer in Equation(1.3) is a Helfrich immersion, i.e. a solution to the Euler–Lagrange equation (1.4) ΔH+|A0|2Hλ1Hλ2=0 for some λ1,λ2R,(1.4) where Δ=Δgf denotes the Laplace–Beltrami operator on (Σ,gf). In [Citation13], solutions to Equation(1.4) with small umbilic Willmore energy have been classified, depending on the sign of the Lagrange-multipliers λ1 and λ2. We observe that for λ1,λ2R fixed, Equation(1.4) is also the Euler–Lagrange equation of the Helfrich energy given by (1.5) Hλ1,λ2(f):=W0(f)+λ1A(f)+λ2V(f),(1.5) where the energy either penalizes or favors large area or volume, depending on the sign of λ1 and λ2, respectively.

The L2-gradient flow of the Willmore energy was introduced and studied by Kuwert and Schätzle in their seminal works [Citation14–16].

Their methods are very robust and allow to handle also other situations, such as the surface diffusion flow [Citation17, Citation18] and the Willmore flow of tori of revolution [Citation19]. The locally constrained Helfrich flow, i.e. the L2-gradient flow for the energy Equation(1.5), and its asymptotic behavior have been studied in [Citation20, Citation21], where it was shown that finite time singularities must occur below a certain energy threshold. However, this flow does not preserve the isoperimetric ratio.

The goal of this article is to discuss a dynamic version of the minimization problem Equation(1.3). To this end, we introduce the Willmore flow with prescribed isoperimetric ratio, which decreases W as fast as possible while keeping I(f)I(f0)=σ fixed. This yields the evolution equation (1.6) tf=[ΔH|A0|2H+λ(3A(f)H2V(f))]ν,(1.6) where the Lagrange multiplier λ:=λ(t):=λ(ft) depends on ft:=f(t,·) and is given by (1.7) λ(f):=(ΔH+|A0|2H)(3A(f)H2V(f))dμ|3A(f)H2V(f)|2dμ.(1.7)

In Equation(2.9) we will justify the particular choice of λ, which yields that I is actually preserved along a solution of Equation(1.6)Equation(1.7).

Definition 1.1.

Let σ(0,1),T>0 and let Σg denote a connected, oriented and closed surface with genus gN0. A smooth family of immersions f:[0,T)×ΣgR3 satisfying Equation(1.6) with λ as in Equation(1.7) and I(f)σ is called a σ-isoperimetric Willmore flow with initial datum f0:=f(0,·).

Stationary solutions of the flow Equation(1.6)Equation(1.7) are solutions to the Helfrich Equationequation (1.4) for λ1=3A(f)λ and λ2=2V(f)λ. Conversely, any Helfrich immersion is also a stationary solution to Equation(1.6)Equation(1.7), see Lemma 2.6.

However, as the Lagrange multiplier λ defined in Equation(1.7) depends on the solution, the isoperimetric flow Equation(1.6) substantially differs from the L2-gradient flow of the Helfrich energy Equation(1.5), where the parameters λ1 and λ2 are fixed numbers and chosen a priori. On the analytic side, the integral nature of the Lagrange multiplier makes the evolution Equationequation (1.6) a non-local, quasilinear, degenerate parabolic PDE of 4th order. Also geometrically, the constraint I(f)σ causes new difficulties, as we cannot control the area and the volume independently along the flow (as in [Citation21], for instance), but only the isoperimetric ratio I.

The Willmore flow with a constraint on either the area or the enclosed volume has been studied in [Citation22] and a recent article by the author [Citation23]. However, the situation here is fundamentally different and several new challenges arise.

First, if only the area or the volume is prescribed (and nonzero), constrained critical points of the corresponding variational problem are in fact Willmore immersions, i.e. solutions of Equation(1.4) with λ1=λ2=0, due to the scaling invariance of the Willmore energy. Although still an active field of research, the classification of these Willmore immersions is much better understood than that of general solutions of Equation(1.4) and a crucial ingredient in classifying the blow-ups in [Citation23]. Second, in [Citation23] the different scaling of the energy and constraint has been used to represent the Lagrange multiplier in a way that allows for good a priori estimates. This neat trick is clearly not available for the flow Equation(1.6)Equation(1.7). Third, unlike in [Citation23], the Lagrange multiplier has a much more complicated algebraic structure and cannot be treated as a lower order term.

These obstructions are the reason for a new energy threshold in the following main result on global existence and convergence.

Theorem 1.2.

Let f0:S2R3 be a smooth immersion with I(f0)=σ(0,1) and such that W(f0)min{4πσ,8π}. Then there exists a unique σ-isoperimetric Willmore flow with initial datum f0. This flow exists for all times and, as t, it converges smoothly after reparametrization to a Helfrich immersion f with I(f)=σ solving Equation(1.4) with λ10 and λ20.

This shows a fundamentally different behavior of the isoperimetric Willmore flow and the Helfrich flow, where finite time singularities occur, cf. [Citation20, Citation21]. Consequently, despite its new analytic challenges, the introduction of the non-local Lagrange multiplier has a regularizing effect on the gradient flow, see also [Citation24] for a related result for the mean curvature flow.

The 4πσ-threshold in Theorem 1.2 is motivated by the following simple application of the triangle inequality in L2(dμ). With I(f)=σ and Equation(1.2), we have Σ3A(f)H2V(f)2dμ36A(f)2(4πσW(f0))2.

This estimate bounds the denominator in Equation(1.7) from below if W(f0)<4πσ. Moreover, it allows to control the Lagrange multiplier in the crucial estimates by essentially lower order quantities, see Sections 4.

We highlight that the assumption in Theorem 1.2 is not an implicit smallness of the initial energy, cf. [Citation15, Citation18], but the threshold is explicitly given, although very little is known about minimizers and critical points of Equation(1.3). Moreover, as σ1, the interval of admissible initial energies in Theorem 1.2 becomes arbitrarily small. This seems plausible, since if σ = 1, f0 is a round sphere and the denominator in Equation(1.7) vanishes. Thus, it is a priori unclear whether there exists an admissible immersion f0 in Theorem 1.2 if σ(12,1)—in fact, this is equivalent to the condition β0(σ)4πσ. In Theorem 7.1, we will prove β0(σ)<4πσ for σ(0,1), which is asymptotically sharp as σ1, and consequently the existence of a suitable f0 follows. We also point out that it is unknown if the energy threshold in Theorem 1.2 is optimal as it is for the classical Willmore flow [Citation19, Citation25].

The proof of Theorem 1.2 is based on the methods developed by Kuwert–Schätzle for the Willmore flow [Citation14–16]. Under a non-concentration assumption on the curvature, we use localized energy estimates to control the evolution, see Sections 3. However, as in [Citation23], these estimates depend on certain Lp-type bounds on λ. The key ingredient of this paper is that for locally small curvature and if the initial energy is below the threshold of Theorem 1.2, the Lagrange multiplier can be absorbed in the estimates, see Sections 4, in particular Lemmas 4.1 and 4.2. This is an essential observation, which we can use to prove a lower bound on the lifespan and to construct a blow-up limit in the spirit of [Citation15], see Sections 5. Using the control over the Lagrange multiplier in the energy regime of Theorem 1.2, we deduce a crucial rigidity result: either the blowup is a compact Helfrich immersion or a Willmore immersion, see Proposition 5.4. In the first case, we conclude global existence and convergence by an argument based on the Łojasiewicz–Simon inequality in the spirit of [Citation26], combined with recent progress on this inequality in the presence of constraints [Citation27]. Due to the rigidity of the blow-up, we can follow the inversion strategy in [Citation16] relying on the classification of compact Willmore spheres [Citation28] to exclude the second case.

This last step is also where we crucially make use of the assumption Σg=S2. In the case of higher genus, a classification result for Willmore surfaces as in [Citation28] is currently lacking. Even if such a classification were available, a precise comprehension of the behavior under inversion would be indispensable to extend the argument beyond the spherical case. However, since the blow-up analysis is also available if g1, we establish the following remarkable dichotomy result.

Corollary 1.3.

Let σ(0,1), let Σ be a closed, oriented and connected surface and suppose that f:[0,T)×ΣR3 is a maximal σ-isoperimetric Willmore flow such that W(f0)<4πσ. Then there exist ĉ(0,1),(tj)jN[0,T),tjT,(rj)N(0,) and (xj)jNR3 such that the sequence of immersions f̂j:=rj1(f(tj+rj4ĉ,·)xj) converges, as j, smoothly on compact subsets of R3 after reparametrization to a proper Helfrich immersion f̂:Σ̂R3 where Σ̂ is a complete surface without boundary. Moreover

  1. if Σ̂ is compact, then T= and, as t, the flow f converges smoothly after reparametrization to a Helfrich immersion f as t.

  2. if Σ̂ is not compact, then f̂ is a Willmore immersion.

Hence, under the above assumptions, in the singular case (b) the influence of the (non-local) constraint vanishes after rescaling as t and the purely local term in Equation(1.6), coming from the Willmore functional, dominates.

We now outline the structure of this article. After a brief review of the most relevant analytic and geometric background in Sections 2, we start our analysis by carefully computing and estimating a localized version of the energy decay in Sections 3. In Sections 4, we control the Lagrange multiplier in the energy regime of Theorem 1.2 which then enables us to construct a blow-up limit in Sections 5. Finally, in Sections 6 we prove our convergence result, Theorem 1.2, and Corollary 1.3 before we show Theorem 7.1 in Sections 7, yielding that the set of admissible initial data in Theorem 1.2 is always non-empty.

2 Preliminaries

In this section, we will briefly review the geometric and analytic background and prove some first properties of the flow Equation(1.6), see also [Citation29] for a more detailed discussion.

2.1 Geometric and analytic background

In the following, Σg always denotes an abstract compact, connected and oriented surface of genus gN0 without boundary.

An immersion f:ΣgR3 induces the pullback metric gf=f·,· on Σg, which in local coordinates is given by gij:=if,jf, where ·,· denotes the Euclidean metric. The chosen orientation on Σg determines a unique smooth unit normal field ν:ΣgS2 along f, which in local coordinates in the orientation is given by (2.1) ν=1f×2f|1f×2f|.(2.1)

We will always work with this unit normal vector field.

The (scalar) second fundamental form of f is then given by Aij:=ijf,ν and the mean curvature and the tracefree part of the second fundamental form are defined as H:=gijAij and Aij0:=Aij12Hgij, where gij:=(gij)1. Important relations are (2.2) |A|2=|A0|2+12H2=2|A0|2+2K,(2.2) where K denotes the Gauss curvature. Consequently, using Equation(1.1), we find (2.3) Σ|A|2dμ=W0(f)+2W(f)=4W(f)8π+8πg.(2.3)

The Levi-Civita connection =f induced by the metric gf extends uniquely to a connection on tensors, which we also denote by . For an orthonormal basis {e1,e2} of the tangent space, the Codazzi–Mainardi equations yield (2.4) iH=(jA)(ei,ej)=2(jA0)(ei,ej),(2.4) cf. [Citation15, (5)].

Clearly, potential singularities for the flow Equation(1.6) occur if V(f) becomes zero or if the denominator in Equation(1.7) vanishes. Note that in the latter case Hconst, thus f is a constant mean curvature immersion.

Lemma 2.1.

Let σ(0,1) and let f:ΣgR3 be an immersion with I(f)=σ. Then

  1. V(f)0;

  2. if g = 0, i.e. Σg=S2, or if f is an embedding, then Hconst. In particular, the denominator in Equation(1.7) is nonzero.

Proof.

The first statement follows immediately from the definition of I. For (ii), we assume by contradiction that Hconst, so f:ΣgR3 is an immersion with constant mean curvature. If Σg=S2, then f has to parameterize a round sphere by a result of Hopf [Citation30, Theorem 2.1, Chapter VI]. In the second case, f has to parameterize a round sphere by the famous theorem of Aleksandrov [Citation31]. In both cases this contradicts σ1. □

Despite its geometric degeneracy, Equation(1.6) is still a parabolic equation. Thus, starting with a smooth nonsingular initial datum, it is possible to prove the following short-time existence result in similar fashion as it is outlined in [Citation32, Chapter 4, Proposition 2.1], after observing that we can integrate by parts in Equation(1.7) so that the numerator of the Lagrange-multiplier contains no second order derivatives of A any more.

Proposition 2.2.

Let f0:ΣgR3 be a smooth immersion with Hf0const and I(f0)=σ(0,1). Then there exist T(0,] and a unique, non-extendable σ-isoperimetric Willmore flow f:[0,T)×ΣgR3 with initial datum f(0):=f(0,·)=f0.

If Σg=S2, assumption Hf0const in Proposition 2.2 follows from σ(0,1) by Lemma 2.1 (ii).

2.2 Evolution of geometric quantities

In this subsection, we will briefly review the variations of the relevant geometric quantities and energies.

Lemma 2.3.

[Citation23, Lemma 2.3] Let f:[0,T)×ΣgR3 be a smooth family of immersions with normal velocity tf=ξν. For an orthonormal basis {e1,e2} of the tangent space, the geometric quantities induced by f satisfy (2.5) t(dμ)=Hξdμ,(2.5) (2.6) tH=Δξ+|A|2ξ,(2.6) (2.7) (tA)(ei,ej)=ij2ξAikAkjξ.(2.7)

As a consequence, we have the following first variation identities, cf. [Citation23, Lemma 2.4].

Proposition 2.4.

Let f:ΣgR3 be an immersion and let φC(Σg;R3). Then we have W0(f)φ=W0(f),φL2(dμ)=(ΔH+|A0|2H)ν,φdμ,A(f)φ=A(f),φL2(dμ)=Hν,φdμ,V(f)φ=V(f),φL2(dμ)=ν,φdμ

Moreover, if I(f)>0, we have I(f)φ=I(f),φL2(dμ)=σ3A(f)Hν2V(f)ν,φdμ.

Proof.

Since W0,A, and V are invariant under orientation-preserving diffeomorphisms of Σg, we only need to consider normal variations, as any tangential variation corresponds to a suitable orientation-preserving family of reparametrizations (see for instance [Citation33, Theorem 17.8]), which leaves the quantities unchanged.

The variation of A then follows immediately from Equation(2.5). For W0 and V consider [Citation23, Lemma 2.4], for instance. The variation of I then follows. □

The scaling behavior of the energies yields the following important identities.

Lemma 2.5.

Let f:ΣgR3 be an immersion. Then we have (ΔH+|A0|2H)ν,fdμ=0,Hν,fdμ=2A(f),ν,fdμ=3V(f).

Proof.

By the scaling invariance of the Willmore energy, we find W0(f),fL2(dμ)=ddα|α=1W(αf)=0, so Proposition 2.4 yields the claim. For A and V we may proceed similarly, using the scaling behavior A(αf)=α2A(f),V(αf)=α3V(f) for all f:ΣgR3,α>0. □

This yields that Helfrich immersions are precisely the stationary solutions of Equation(1.6)Equation(1.7).

Lemma 2.6.

Let f:ΣgR3 be an immersion with I(f)=σ(0,1) and Hfconst. Then f is a Helfrich immersion if and only if it is a stationary solution to the σ-isoperimetric Willmore flow.

Proof.

The “if” part of the statement is immediate. Suppose f is a Helfrich immersion. We multiply Equation(1.4) with f,ν, integrate and use Lemma 2.5 to conclude 2λ1A(f)+3λ2V(f)=0.

By Lemma 2.1(i) we have V(f)0. Hence, with λ:=λ2V(f)2, EquationEquation (1.4) reads (2.8) ΔH+|A0|2Hλ(3A(f)H2V(f))=0.(2.8)

We have I(f)0 by Proposition 2.4, so by testing Equation(2.8) with I(f)ν and integrating it follows that λ is given as in Equation(1.7), so f is indeed stationary. □

It is not difficult to see that along a solution of Equation(1.6) with I(f)>0, the isoperimetric ratio is indeed preserved, since by Proposition 2.4, Equation(1.6) and Equation(1.7) we have (2.9) ​ddtI(f)=I(f),tfdμ=I(f),W0(f)+λI(f)I(f)dμ=I(f),W0(f)dμ+λI(f)|I(f)|2dμ=0.(2.9)

On the other hand, the Willmore energy decreases since by Equation(2.9) (2.10) ddtW0(f)=(ΔH+|A0|2H)ν,tfdμ=|tf|2dμ0.(2.10)

EquationEquations (2.9) and Equation(2.10) are the key features in studying the flow Equation(1.6) and of vital importance for our further analysis. We highlight two immediate consequences.

Remark 2.7.

  1. The computation in Equation(2.10) implies that W0 is a strict Lyapunov function along the flow Equation(1.6), i.e. W0 is strictly decreasing unless tf=0, so f is stationary (by uniqueness of the solution). By Equation(1.1), this also holds for W.

  2. Since W is monotone, the limit limtTW(f(t,·))[βg(σ),W(f0)] exists.

As Equation(1.6) is a (degenerate) parabolic equation, the scaling behavior in time and space is central in understanding the problem. Therefore, we gather the scaling behavior of some important quantities in the following lemma. The powers appearing in the time integrals below will naturally appear later in our energy estimates, see Proposition 3.3.

Lemma 2.8.

Let σ(0,1),f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow and let r > 0. Let f˜:[0,r4T)×ΣgR3,f˜(t,p):=r1f(r4t,p). Then

  1. f˜ is a σ-isoperimetric Willmore flow;

  2. the Lagrange multiplier λ˜ of f˜ satisfies λ˜(t)=λ(r4t);

  3. 0Tλ2A(f)2dt=0r4Tλ˜2A(f˜)2dt and 0T|λ|43|V(f)|43dt=0r4T|λ˜|43|V(f˜)|43dt.

Proof.

Follows from the scaling behavior of the geometric quantities and a direct calculation. □

3 Localized energy estimates

As in [Citation15, Section 3] and [Citation23, Sections 2.3 and 3], we will start our analysis by localizing the energy decay Equation(2.10). The main goal of this section is to show that all derivatives of A can be bounded along the flow, if the energy concentration and a suitable time integral involving the Lagrange multiplier are controlled. Note that at this stage, we do not yet need to assume Σg=S2 or any restriction on the initial energy.

Lemma 3.1.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow. Let η˜Cc(R3) and define η:=η˜°f. Then we have t12H2ηdμ+|W0(f)|2ηdμ=3λA(f)ΔHHηdμ+(3λA(f)H2λV(f))|A0|2Hηdμ2W0(f),νH,ηgfdμW0(f),νHΔηdμ+12H2tηdμ. and t|A0|2ηdμ+|W0(f)|2ηdμ=6λA(f)2H,A0gfηdμ+(3λA(f)H2λV(f))|A0|2Hηdμ2W0(f),ν(H,ηgf+A0,2ηgf)dμ+|A0|2tηdμ.

Proof.

This computation is very similar to [Citation32, Chapter 4, Lemma 2.8] (see also [Citation15, Section 3]) if one replaces λν with (3λA(f)2λV(f))ν, so we will focus on the differences. We will use a local orthonormal frame {ei(t)}i=1,2 for our computations and find (3.1) t12H2ηdμ+|W0(f)|2ηdμ=(3λA(f)H2λV(f))ΔHηdμ+(3λA(f)H2λV(f))|A0|2Hηdμ+(2ξiHiη+HξΔη)dμ+12H2tηdμ,(3.1) writing tf=ξν. Moreover, we have (3.2) (2ξiHiη+HξΔη)dμ=2W0(f),νiHiηdμ+6λA(f)HiHiηdμ4λV(f)iHiηdμW0(f),νHΔηdμ+3λA(f)H2Δηdμ2λV(f)HΔηdμ.(3.2)

If we carefully combine the terms with λ in Equation(3.1) and Equation(3.2), the claim follows after integrating by parts, where the terms involving derivatives of H and the factor 2λV(f) cancel.

For the second identity, arguing similarly as in [Citation32, Chapter 4, Lemma 2.8] we have t(|A0|2dμ)=2i(jξA0(ei,ej))dμj(ξjH)dμ+W0(f),ξνdμ=2i(jξA0(ei,ej))dμj(ξjH)dμ|W0(f)|2dμ+λ(ΔH+|A0|2H)(3A(f)H2V(f))dμ.

Integrating by parts and using Equation(2.4) we conclude t|A0|2ηdμ+|W0(f)|2ηdμ=[2jξAij0iη+ξjHjη+λ(ΔH+|A0|2H)(3A(f)H2V(f))η]dμ+|A0|2tηdμ=2W0(f),νAij0ji2ηdμ+2λ(3A(f)H2V(f))Aij0ij2ηdμ2W0(f),νiHiηdμ+2λ(3A(f)H2V(f))iHiηdμ+λ(ΔH+|A0|2H)(3A(f)H2V(f))ηdμ+|A0|2tηdμ.

Now, using integration by parts and Equation(2.4) once again, we have (3λA(f)H2λV(f))(2iHiη+2Aij0ij2η+ΔHη)dμ=6λA(f)2H,A0gfηdμ.

The claim follows. □

We will now carefully estimate the integrals in Lemma 3.1. To this end, we choose a particular class of test functions. Let γ˜Cc(R3) with 0γ˜1 and assume ||Dγ˜||Λ,||D2γ˜||Λ2 for some Λ>0. Then setting (3.3) γ:=γ˜°f:[0,T)×ΣgR we find |γ|Λ and |2γ|Λ2+|A|Λ,(3.3) and note that γ(t,·) has compact support in Σg, which is compact, for all 0t<T, see also [Citation23, Equation(3.1)].

For the rest of this article, we denote by C a universal constant with 0<C< which may change from line to line.

Lemma 3.2.

Let σ(0,1), let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow and let γ be as in Equation(3.3). Then we have t|A|2γ4dμ+32|W0(f)|2γ4dμC|λ|A(f)(2H,Agfγ4dμ+|A|4γ4dμ+Λ|A|3γ3dμ)+C|λ||V(f)|(|A|3γ4dμ+Λ|A|2γ3dμ)+CΛ4[γ>0]|A|2dμ+CΛ2|A|4γ2dμ.

Proof.

We have 22φ,Agf=22φ,A0gf+HΔφ for any φC([0,T)×Σg) by a direct computation in a local orthonormal frame. Hence, using Lemma 3.1 and |A|2=|A0|2+12H2, cf. Equation(2.2), we find t|A|2γ4dμ+2|W0(f)|2γ4dμ=6λA(f)(2H,Agfγ4dμ+|A0|2H2γ4dμ)4λV(f)|A0|2Hγ4dμ4W0(f),νH,γ4gfdμ2W0(f),ν2γ4,Agfdμ+|A|2tγ4dμ.

The terms W0(f),νH,γ4gfdμ and W0(f),ν2γ4,Agfdμ can be estimated as in [Citation15, Lemma 3.2]. Since tγ4=4γ3γ˜°ftf we have by Equation(3.3) |tγ4|CΛγ3(|W0(f)|+|λ|A(f)|A|+|λ||V(f)|).

Consequently, we find |A|2tγ4dμε|W0(f)|2γ4dμ+C(ε)Λ2|A|4γ2dμ+CΛ|λ|A(f)|A|3γ3dμ+CΛ|λ||V(f)||A|2γ3dμ.

Choosing ε>0 small enough, the claim follows from the estimates above. □

Note that on the right hand side of Lemma 3.2, terms involving the Lagrange multiplier multiplied with powers of A up to 4-th order and even second derivatives of H appear. With the energy, we can only control the L2-norms of H and A. In the following Proposition 3.3 we will close this gap by using higher powers of the Lagrange multiplier, the area and the volume, see also [Citation23, Proposition 3.3]; these powers behave correctly under rescaling, cf. Lemma 2.8. We will combine this with the interpolation techniques from [Citation14, Citation15] to get control on the local W2,2-norm of A, in terms of the (localized) Willmore gradient, at least if the L2-norm of A is locally small.

Proposition 3.3.

There exist universal constants ε0,c0,C(0,) with the following property: Let σ(0,1), let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow and let γ be as in Equation(3.3). If we have (3.4) [γ>0]|A|2dμ<ε0for some time t[0,T),(3.4) then at time t we can estimate t|A|2γ4dμ+c0(|2A|2+|A|2|A|2+|A|6)γ4dμCΛ4[γ>0]|A|2dμ+C(λ2A(f)2+|λ|43|V(f)|43)|A|2γ4dμ.

Here [γ0>0]|A0|2dμ0:=[γ>0]|A|2dμ|t=0.

Proof.

Using the assumption and the interpolation inequality in [Citation15, Proposition 2.6] (see also [Citation23, Proposition 3.2]), we have at time t[0,T) (|2A|2+|A|2|A|2+|A|6)γ4dμC|W0(f)|2γ4dμ+CΛ4[γ>0]|A|2dμ.

Consequently, from Lemma 3.2, we find for some c0(0,) (3.5) t|A|2γ4dμ+2c0(|2A|2+|A|2|A|2+|A|6)γ4dμC|λ|A(f)(2H,Agfγ4dμ+|A|4γ4dμ+Λ|A|3γ3dμ)+C|λ||V(f)|(|A|3γ4dμ+Λ|A|2γ3dμ)+CΛ2|A|4γ2dμ+CΛ4[γ>0]|A|2dμ.(3.5)

For the first term on the right hand side of Equation(3.5), we infer using Young’s inequality (3.6) |λ|A(f)(2H,Agfγ4dμ+|A|4γ4dμ+Λ|A|3γ3dμ)ε|2A|2γ4dμ+ε|A|6γ4dμ+Λ2|A|4γ2dμ+C(ε)|λ|2A(f)2|A|2γ4dμ.(3.6)

The second term on the right hand side of Equation(3.5) can be estimated by using Young’s inequality with p = 4 and q=43 and γ1 to obtain (3.7) C|λ||V(f)|(|A|3γ4dμ+Λ|A|2γ3dμ)ε|A|6γ4dμ+C(ε)|λ|43|V(f)|43|A|2γ4dμ+CΛ4[γ>0]|A|2dμ.(3.7)

Moreover, we have the estimate CΛ2|A|4γ2dμε|A|6γ4dμ+C(ε)Λ4[γ>0]|A|2dμ using Young’s inequality. Combining this with Equation(3.5), Equation(3.6), and Equation(3.7) and choosing ε>0 sufficiently small, the claim follows. □

Assumption (3.4) means that the second fundamental form is small on the support of γ. Note that this will only be satisfied locally, since by Equation(2.3) we always have |A|2dμ[8π,4W(f)8π+8πg]. We will now study the situation, where Equation(3.4) is satisfied on all balls with a certain radius, yielding a control over the concentration of the Willmore energy in R3. Following [Citation16] we introduce the following notation.

Definition 3.4.

For a smooth family of immersions f:[0,T)×ΣgR3,t[0,T), r > 0, we define the curvature concentration function (3.8) κ(t,r):=supxR3Br(x)|A|2dμ.(3.8)

Here and in the rest of this article, we follow the notation of [Citation14], i.e. the integrals over balls Br(x)R3 have to be understood over the preimages under ft.

If Γ>1 denotes the minimal number of balls of radius 1/2 necessary to cover B1(0)R3, then (3.9) κ(t,ρ)Γ·κ(t,ρ/2)for all 0t<T.(3.9)

We now prove an integrated form of Proposition 3.3.

Proposition 3.5.

Let ε0>0 be as in Proposition 3.3. There exist universal constants ε1(0,ε0),c0,C>0 with the following property: Let σ(0,1), let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow and let ρ>0 be such that (3.10) κ(t,ρ)<ε1for all t[0,T).(3.10)

Then for all xR3 and t[0,T) we have Bρ/2(x)|A|2dμ|t+c00tBρ/2(x)(|2A|2+|A|2|A|2+|A|6)dμdτBρ(x)|A0|2dμ0+C(1+σ2)ρ40tBρ(x)|A|2dμdτ+C0tλ2A(f)2Bρ(x)|A|2dμdτ.

Proof.

Fix xR3. Let γ˜Cc(R3) be a cutoff function with χBρ/2(x)γ˜χBρ(x),||Dγ˜||Cρ and ||D2γ˜||Cρ2. Therefore, γ:=γ˜°f is as in Equation(3.3) with Λ=Cρ. Moreover, if we take ε1>0 small enough, we have the estimate (3.11) ρ2CA(f)(3.11) as a consequence of Simon’s monotonicity formula [Citation34], see also [Citation23, Lemma 4.1]. Now, since we have |V(f)|43=(σ36π)23A(f)2 by Equation(2.9) and Equation(1.2), we observe (3.12) |λ|43|V(f)|43=(36π)23A(f)2|λ|43σ23C(λ2+σ2)A(f)2C(λ2A(f)2+σ2ρ4),(3.12) where we used Young’s inequality (with p=32 and q = 3). The statement then immediately follows by integrating Proposition 3.3 in time. □

Remark 3.6.

If we directly integrate Proposition 3.3, we have to deal with two terms involving λ, both of whose time integrals behave correctly under parabolic rescaling, cf. Lemma 2.8. The estimate Equation(3.12) above reveals that if Equation(3.10) is satisfied, then it suffices to control merely the λ2A(f)2-term, since (3.13) |λ|43|V(f)|43C(λ2A(f)2+σ2ρ4).(3.13)

For the blow-up construction in Sections 5, we will need the following higher order estimates for the flow in the case of non-concentrated curvature, cf. [Citation15, Theorem 3.5], [Citation23, Proposition 3.5].

Proposition 3.7.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow. Suppose ρ>0 is chosen such that TT*ρ4 for some 0<T*< and (3.14) κ(t,ρ)ε<ε1for all 0t<T,(3.14) where ε1>0 is as in Proposition 3.5. Moreover, assume (3.15) 0Tλ2A(f)2dtL¯<.(3.15)

Then for all t(0,T),xR3 and mN0 we have the local estimates ||mA||L2(Bρ/8(x))C(m,T*,L¯,σ)εtm4,||mA||LC(m,T*,L¯,σ)εtm+14, and the global bounds (3.16) ||mA||L2(dμ)C(m,T*,L¯,σ)tm4(|A0|2dμ0)12.(3.16)

In contrast to [Citation15, Theorem 3.5] and [Citation23, Proposition 3.5], we do not only prove local bounds, but also the global L2-control Equation(3.16). Note that the global L2-norms could also be estimated by the L-bounds and the area. However, this is disadvantageous since the area cannot be controlled along the flow, and in fact is always expected to diverge in the blow-up process, cf. Lemma 6.1. The necessity for the finer estimates leading to Equation(3.16) is why we give full details on the proof here, even though the argument is very similar to [Citation15, Theorem 3.5].

Proof of Proposition 3.7.

After parabolic rescaling, cf. Lemma 2.8, we may assume ρ = 1. Let xR3 and define K(t):=B1(x)|A|2dμ and L(t):=λ2A(f)2. Then, for all t[0,T) using that Kε<ε1 by Equation(3.14), we deduce from Proposition 3.5 (3.17) 0tB1/2(x)|2A|2dμdτCB1(x)|A0|2dμ0+C(1+σ2)0tB1(x)|A|2dμdτ+C0tLB1(x)|A|2dμdτC(σ)(K(0)+0t(1+L)K​dτ).(3.17)

Moreover, as K(t)ε<ε1 by Equation(3.14) we can interpolate by combining [Citation15, Lemma 2.8] and [Citation14, Lemma 4.2] to find 0t||A||L(B1/4(x))4dτC(σ)ε1(K(0)+0t(1+L)K​dτ)C(T*,L¯,σ), where we used the assumptions (3.14), (3.15) and TT* in the last step. Thus, defining a(t):=||A||L(B1/4(x))4 and using TT* and Equation(3.15), we have the estimate (3.18) 0t(1+L+a)dτC(T*,L¯,σ).(3.18)

Now, we pick γ˜Cc(R3) with χB1/8(x)γ˜χB1/4(x) and γ:=γ˜°f. Note that Equation(3.3) is satisfied with a universal Λ>0, which we do not keep track of. As in [Citation15, Theorem 3.5], we define Lipschitz cutoff functions in time via ξj(t):={0for t(j1)Tm,mT(t(j1)Tm)for (j1)TmtjTm,1for tjTm, where mN and 0jm. We also define ξ1(t):=0 and ξ0(t):=1 for all tR if m = 0. We note that ξm(T)=1 and (3.19) 0ddtξjmTξj1for all jN0.(3.19)

Furthermore, for 0jm we define Ej(t):=|2jA|2γ4j+4dμ. Then, by Proposition B.3, using γ1 and Equation(3.13) we have ddtEj(t)+12Ej+1(t)C(j,m,σ)[(1+L(t)+a(t))Ej(t)+(1+L(t)+a(t))K(t)].

Therefore, if we define ej:=ξjEj this implies using Equation(3.19) and ξj1 (3.20) ​ddtej(t)mTξj1(t)Ej(t)+C(j,m,σ)(1+L(t)+a(t))ej(t)+C(j,m,σ)(1+L(t)+a(t))K(t)12ξj(t)Ej+1(t).(3.20)

We now claim that for all 0jm and t(0,T) we have (3.21) ej(t)+120tξjEj+1dsC(j,m,T*,L¯,σ)Tj(K(0)+K(t)+0t(1+L+a)K​dτ).(3.21)

We proceed by induction on j. For j = 0 we have ξ01 on (0,T). Therefore, we clearly have e0=|A|2γ4dμK. Moreover, by Equation(3.17) we find 0tE1(s)ds=0t|2A|2γ8dμdsC(σ)(K(0)+0t(1+L)K​dτ).

For j1, integrating Equation(3.20) on [0,t] and using ej(0)=0, we find ej(t)+120tξjEj+1dτC(j,m,σ)0t(1+L+a)ejdτ+C(j,m,σ)0t(1+L+a)K​dτ+mT0tξj1EjdτC(j,m,T*,L¯,σ)0t(1+L+a)ejdτ+C(j,m,T*,L¯,σ)Tj(K(0)+K(t)+0t(1+L+a)K​dτ), by the induction hypothesis and since TT*. Using Gronwall’s inequality and estimating the exponential term by Equation(3.18), we find ej(t)120tξjEj+1ds+C(j,m,T*,L¯,σ)Tj(K(0)+K(t)+0t(1+L+a)K​dτ)+C(j,m,T*,L¯,σ)Tj0t(K(0)+K(τ)+0τ(1+L+a)Kds)(1+L(τ)+a(τ))dτ.

The estimate Equation(3.21) then follows by using Equation(3.18) and estimating the double integral via 0t(0τ(1+L+a)Kds)(1+L(τ)+a(τ))dτC(T*,L¯,σ)0t(1+L+a)Kds, where we also used Equation(3.18) once again. Now, for the local estimates, we evaluate Equation(3.21) at t=T,j=m and use Equation(3.14), TT* and Equation(3.18). Recalling Kε by Equation(3.14), this yields (3.22) |2mA|2γ4m+4dμ|t=TC(m,T*,L¯,σ)Tmε.(3.22)

For the global L2-estimate, we observe that Equation(3.21) is linear in ej and K. Hence, as in [Citation23, Proposition 4.2], we can sum up the local bounds to get (3.23) |2mA|2dμC(m,T*,L¯,σ)Tm(|A0|2dμ0+|A|2dμ+0T(1+L+a)|A|2dμdτ)C(m,T*,L¯,σ)Tm|A0|2dμ0,(3.23) where we used Equation(2.3), Equation(2.10), and Equation(3.18). After renaming T into t, Equation(3.22) and Equation(3.23) are precisely the desired L2-estimate for even orders of derivatives. Exactly as in [Citation15, Theorem 3.5], the local and global L2-estimates for 2m+1A follow by interpolation. The L-estimate can then be deduced as in [Citation15, Theorem 3.5] as well. □

4 Controlling the Lagrange multiplier

In this section, we will provide some important estimates for the Lagrange multiplier under the assumption that the initial energy is not too large. In contrast to [Citation23], the crucial power of λ is not of lower order when compared to the left hand side of Proposition 3.3. Nevertheless, a first immediate feature of the energy regime from Theorem 1.2 is that we can uniformly bound the denominator of λ from below.

Lemma 4.1.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow with W(f0)<4πσ. Then 3A(f)H2V(f)2dμ36A(f)2(4πσW(f0))2.

Proof.

This follows from the reverse triangle inequality in L2(dμ), Equation(2.9) and Equation(2.10). □

While the scaling techniques from [Citation23, Lemma 4.3] are not available here, we still get the following key estimate, which gives a control over λ by quantities which will be suitably integrable.

Lemma 4.2.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow with W(f0)<4πσ. Then, we have |λ|A(f)6(4πσW(f))tf,νdμ+|A0|2Hdμ.

Proof.

We test the evolution Equationequation (1.6) with the normal ν and integrate to obtain tf,νdμ=|A0|2Hdμ+(3A(f)Hdμ2V(f)A(f))λ, where we used the divergence theorem. We now estimate the prefactor of λ by 3A(f)Hdμ2V(f)A(f)2|V(f)|A(f)3A(f)Hdμ6A(f)(4πσW(f)), using the fact that I(f)σ by Equation(2.9). By the assumption and Equation(2.10) this is strictly positive and the claim follows. □

We remark that the existence of f0:ΣgR3 with I(f0)=σ(0,1) satisfying the assumption W(f0)<4πσ is not yet known and—in general—not true. For the case g = 0, this will follow from Theorem 7.1. However, for tori we have W(f0)2π2 by [Citation35], and hence 2π2W(f0)<4πσ can only hold for σ<2π<1. On the other hand, for σ(0,12] and arbitrary genus, there exists f0 with W(f0)<4πσ since βg(σ)<8π by [Citation11, Theorem 1.2]. We now use Lemma 4.2 to deduce the time integrability Equation(3.15) for λ in the case of small curvature concentration, which enables us to bound all derivatives of the second fundamental form by Proposition 3.7.

Lemma 4.3.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow. Let W(f0)K<4πσ and let ρ>0 be such that κ(t,ρ)ε<ε1 for all t[0,T), where ε1>0 is as in Proposition 3.5. Then for all τ[0,T) we have 0τλ2A(f)2dtC(4πσK)4(W(f0)W(f(τ))+C(σ,g)(τ12ρ2+τρ4)).

Note that by the invariance of the Willmore energy and the isoperimetric ratio, this estimate is preserved under parabolic rescaling, cf. Lemma 2.8.

Proof of Lemma 4.3.

By the assumption we get the local control from Proposition 3.5. As in [Citation23, Proposition 4.2], we can sum up these local bounds to get the global estimate 0τ|A|6dμdtC|A0|2dμ0+C(1+σ2)ρ40τ|A|2dμdt+C0τλ2A(f)2|A|2dμdt.

Now, by Equation(2.3), the energy decay Equation(2.10) and the assumption, we have (4.1) |A|2dμC(σ,g).(4.1)

Thus, we obtain the estimate (4.2) 0τ|A|6dμdtC(σ,g)(1+τρ4+0τλ2A(f)2dt).(4.2)

By Equation(2.2) we have |A0|2|H|C|A|3. Therefore, using Lemma 4.2 we find (4.3) 0τλ2A(f)2dtCb2(0τ|tf|2dμdt+0τC(σ,g)A(f)|A|4dμdt),(4.3) by Cauchy–Schwarz and Equation(4.1), where b=b(K,σ):=4πσK>0. For the first term in Equation(4.3), by Equation(2.10) and Equation(1.1) we have 0τ|tf|2dμdt=W(f0)W(f(τ)). For the second term in Equation(4.3), we use Equation(3.11), Cauchy–Schwarz in time and space, Equation(4.1) and then Equation(4.2) to find C(σ,g)b20τ1A(f)|A|4dμdtC(σ,g)ρ2b2(0τ|A|6dμdt)12(0τ|A|2dμdt)12C(σ,g)b2τ12ρ2(1+τ12ρ2+(0τλ2A(f)2dt)12)C(δ,σ,g)b4(τ12ρ2+τρ4)+δ0τλ2A(f)2dt, for every δ>0 by Young’s inequality, and estimating b=4πσKC(σ) in the last step. The statement then follows from Equation(4.3) by taking δ>0 sufficiently small. □

5 The blow-up and its properties

In this section, we will rescale an isoperimetric Willmore flow as we approach the maximal existence time to obtain a limit immersion. Analyzing the properties of this limit will be the keystone in proving our main result, Theorem 1.2.

5.1 A lower bound on the existence time

As in [Citation14] and [Citation23], the first step is to prove a lower bound on the existence time of an isoperimetric flow which respects the parabolic rescaling in Section 5.2.

To that end, we state a general lifespan result for possible future reference, where the lower bound only depends on the radius of concentration ρ, the isoperimetric ratio σ and the behavior of the L2-norm of λA(f) near t = 0.

Proposition 5.1.

Let ε1>0 be as in Proposition 3.5. There exist universal constants δ¯>0 nd ε¯(0,min{8π,ε1}) with the following property: Let f0:ΣgR3 be an immersion with I(f0)=σ(0,1),Hf0const and W(f0)<4πσ. Let f be the σ-isoperimetric Willmore flow with initial datum f0. Assume that

  1. κ(0,ρ)ε<ε¯ for some ρ>0;

  2. there exists ω¯>0 with the following property: For any t0[0,min{T,ρ4ω¯}] with κ(t,ρ)<ε1 for all 0t<t0, we have 0t0λ2A2dtδ¯.

Then the maximal existence time of the flow satisfies T>ĉρ4 for some ĉ=ĉ(σ,ω¯)(0,1) and (5.1) κ(t,ρ)ĉ1εfor all t[0,ĉρ4].(5.1)

Note that we always have limt00tλ2A2dτ=0. The crucial insight here is that only the decay behavior of the L2-norm of λA under the assumption of small concentration allows control on the existence time in a way which transforms correctly under parabolic rescaling.

Proof of Proposition 5.1.

Without loss of generality, we may assume ρ = 1, otherwise we rescale as in Lemma 2.8, see also [Citation23, Proposition 3.5]. Let T denote the maximal existence time of the flow and let δ¯>0 to be chosen. We define ε¯:=min{ε13Γ,8π} where ε1>0 is as in Proposition 3.3 and Γ>1 is as in Equation(3.9) and set κ(t):=κ(t,1) for t[0,T). By compactness of f([0,t]×Σg) for t < T, the supremum in the definition of κ=κ(·,1) in Equation(3.8) is always attained and the function κ:[0,T)R is continuous with κ(0)ε<ε¯ by (a).

For a parameter ω(0,ω¯], to be chosen later, we now define (5.2) t0:=sup{0tmin{T,ω}|κ(τ)3Γε for all 0τ<t}[0,min{T,ω}].(5.2)

By continuity of tκ(t) and (a), we have t0>0. For t[0,t0), we have κ(t)3Γε<ε1 by Equation(5.2) and the definition of ε¯. Hence, by Proposition 3.5 and assumption (b) we find (5.3) B1/2(x)|A|2dμB1(x)|A0|2dμ0+3c1(1+σ2)Γεt+3c1Γεδ¯,(5.3) for all 0t<t0 where c1=C from Proposition 3.5. Now, if we choose δ¯:=(6c1Γ)1>0 and ω=ω(σ,ω¯)=min{(6c1(1+σ2)Γ)1,ω¯}>0 we find from Equation(5.3) (5.4) B1/2(x)|A|2dμB1(x)|A0|2dμ0+ε2ω1t+ε22εfor all 0t<t0.(5.4)

However, if t0<min{T,ω}, together with Equation(3.9), this implies κ(t)2Γε<ε1 for all 0t<t0 by our choice of ε¯. On the other hand, by Equation(5.2) and continuity, we must have κ(t0)=3Γε, a contradiction.

Consequently, t0=min{T,ω} has to hold. Assume t0=Tω. Then, as before, from Equation(5.4) and Equation(3.9) we find κ(t)2Γε<ε1for all 0t<T=t0, by the definition of ε¯. As Tωω¯ by assumption and 0ω¯λ2A(f)2dtδ¯ by (b), we can apply Proposition 3.7 to conclude that for any 0<ξ<T we have (5.5) ||mA||C(m,ξ,σ,ω¯) for all mN0,t[ξ,T),(5.5) and ||mA||L2C(m,ξ,W(f0),σ,ω¯). Consequently, for all t[ξ,T) we can estimate (5.6) λA(f)2C(ΔHL22+||A||L4AL22)A(f)23A(f)H2V(f)L22C(ξ,W(f0),σ,ω¯),(5.6) using Equation(1.7), Cauchy-Schwarz, Equation(2.2) and Lemma 4.1. Similarly, we find (5.7) λV(f)2C(σ)A(f)(|ΔH|2+|A|6)dμC(σ)(||ΔH||2+||A||6)C(ξ,σ,W(f0),ω¯),(5.7) where we used I(f)σ and Equation(5.5). Exactly with the same arguments as in [Citation14, pp. 330–332] (see also [Citation32, Chapter 4, Proof of Theorem 1.1 after Equation(5.8)]), we can deduce that f(t) smoothly converges to a smooth immersion f(T) as tT. By assumption and the energy decay, we infer from Lemma 4.1 that the denominator in Equation(1.7) is bounded away from zero for all t[0,T), so f(T) is not a constant mean curvature immersion. By Proposition 2.2, we can then restart the flow with initial datum f(T) which contradicts the maximality of T.

Hence, T>ω has to hold. The estimate Equation(5.1) then follows from Equation(5.4) and Equation(3.9) after choosing ĉ=ĉ(σ,ω¯)=min{ω,(2Γ)1,1}>0. □

Together with the integral estimate for the Lagrange multiplier in Lemma 4.3, this now implies the following

Proposition 5.2

(Lifespan bound for small energy gap). Let σ(0,1), let f:[0,T)×ΣgR3 be a maximal σ-isoperimetric Willmore flow such that

  1. W(f0)K<4πσ;

  2. κ(0,ρ)ε<ε¯, where ε¯>0 is as in Proposition 5.1;

  3. W(f0)limtTW(f(t))d¯, where d¯=d¯(K,σ,g)>0.

Then the maximal existence time is bounded from below by T>ĉρ4, where ĉ=ĉ(K,σ,g) and for all 0tĉρ4 we have κ(t,ρ)ĉ1ε.

Note that the limit in (iii) exists due to Remark 2.7 (ii).

Proof of Proposition 5.2.

We check that the assumptions in Proposition 5.1 are satisfied. Let ε¯,δ¯>0 be as in Proposition 5.1. Assumption (a) of Proposition 5.1 holds true by assumption (ii). We now verify assumption (b) in Proposition 5.1. To that end, let ω¯>0 to be chosen and assume that for some t0[0,min{T,ρ4ω¯}] we have κ(t,ρ)<ε1 for all 0t<t0. By (i), we may apply Lemma 4.3 and use (iii) to find the estimate 0t0λ2A2dtC(K,σ,g)(d¯+ω¯12+ω¯)δ¯, if we choose d¯=d¯(K,σ,g)>0 and ω¯=ω¯(K,σ,g)>0 small enough. The assumptions of Proposition 5.1 are thus fulfilled and the result follows with ĉ=ĉ(σ,ω¯)=ĉ(K,σ,g). □

5.2 Existence of a blow-up

In this section, we will rescale as we approach the maximal existence time T(0,] of a σ-isoperimetric Willmore flow f:[0,T)×ΣgR3 with σ(0,1). To that end, let (tj)jN[0,T),tjT,(rj)jN(0,),(xj)jNR3 be arbitrary. By translation invariance and Lemma 2.8 for all jN the flow (5.8) fj:[0,rj4(Ttj))×ΣgR3,fj(t,p):=rj1(f(tj+rj4t,p)xj)(5.8) is also a σ-isoperimetric Willmore flow with initial datum fj(0)=rj1(f(tj,·)xj) and maximal existence time rj4(Ttj). Throughout this section, we will denote all geometric quantities of the flow fj with a subscript j, such as Aj,λj,κj,μj for example. The next lemma guarantees the existence of suitable tj, rj and xj.

Lemma 5.3.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a maximal σ-isoperimetric Willmore flow with W(f0)K<4πσ. Let ĉ=ĉ(K,σ,g)(0,1) be as in Proposition 5.2. Then, there exist sequences (tj)jN[0,T),tjT,(rj)jN(0,) and (xj)jNR3 such that for all jN we have

  1. tj+rj4ĉ<T;

  2. κj(t,1)ε¯ for all t[0,ĉ], where ε¯>0 is as in Proposition 5.1;

  3. infjNB1(0)|Afj(ĉ,·)|2dμfj(ĉ,·)>0.

Proof.

Given any t[0,T), with essentially the same arguments as in [Citation23, Lemma 6.6], one finds a radius rt(0,) such that (5.9) ακ(t,rt)ĉε¯,(5.9) where α=α(K,σ,g)>0. One then argues as in [Citation16, p. 349] (see also [Citation23, Proposition 6.7]), to prove the existence of tjT and (xj)jNR3 such that choosing rj:=rtj, we find that (iii) is satisfied.

Now, the flow fj satisfies κj(0,1)=κ(tj,rj)=κ(tj,rtj)ĉε¯<ε¯ by Equation(5.9) and since ĉ(0,1). Moreover, by the invariances of the Willmore energy we have W(fj(0))K<4πσ for all jN and W(fj(0))limtrj4(Ttj)W(fj(t))=W(f(tj))limtTW(f(t))0,as j.

Consequently, for j sufficiently large, we can apply Proposition 5.2, to find that the maximal existence time of the flow fj is bounded from below by rj4(Ttj)>ĉ which proves (i) and κj(t,1)ε¯ for all t[0,ĉ] by Equation(5.1) which proves (ii). □

Proposition 5.4

(Existence and properties of the limit immersion). Let σ(0,1) and suppose f:[0,T)×ΣgR3 is a maximal σ-isoperimetric Willmore flow with W(f0)K<4πσ. Let ĉ(0,1),tjT,(rj)jN(0,) and (xj)jNR3 be as in Lemma 5.3. Then, there exists a complete, orientable surface Σ̂ without boundary and a proper immersion f̂:Σ̂R3 such that, after passing to a subsequence, rjr[0,] and

  1. as j,f̂j:=fj(ĉ,·)f̂ smoothly on compact subsets of R3, after reparametrization;

  2. we have B1(0)¯|Â|2dμ̂>0 and W(f̂)W(f0);

  3. f̂ is a Helfrich immersion, i.e. a solution to Equation(1.4);

  4. if A(f̂j), then f̂ is a Willmore immersion.

Any Helfrich immersion f̂:Σ̂R3 which arises from the process described above is called a concentration limit. More precisely, we call f̂a blow-up if rj0, a blow-down for rj and a limit under translation if rjr(0,). Note that by Lemma 5.3 (i) the last two can only occur if T=.

We highlight that Proposition 5.4 (iv) is particularly remarkable, since it means that under the assumption of diverging area, the constraint vanishes in the concentration limit, see also [Citation23, Theorem 6.2] for a similar rigidity result. This will be essential in the proof of Theorem 1.2.

Proof of Proposition 5.4.

After passing to a subsequence, we may assume rjr in [0,]. We have ε¯<ε1 and ĉ(0,1) by Proposition 5.1 and hence by Lemma 5.3 (ii) we find κj(t,1)<ε1 for all t[0,ĉ]. We may thus use Lemma 4.3 to bound 0tλj2A(fj)dτC(K,σ,g) for all t[0,ĉ] and for all jN. Consequently, using Proposition 3.7 we conclude that for any jN we have (5.10) ||mAj||C(m,K,σ,g)tm+14,(5.10) (5.11) ||mAj||L2(dμj)C(m,K,σ,g)tm4for 0<tĉ.(5.11)

Moreover, from Simon’s monotonicity formula, cf. [Citation34, Equation(1.3)], for any R > 0 we find R2μj(BR(0))CK<for all jN.

Thus, we may apply the localized version of Langer’s compactness theorem ([Citation15, Theorem 4.2], see also [Citation23, Appendix A]) to the sequence of immersions f̂j:=fj(ĉ,·). After passing to a subsequence, we thus find a proper limit immersion f̂:Σ̂R3, where Σ̂ is a complete (possibly empty) surface without boundary, diffeomorphisms ϕj:Σ̂(j)Uj, where UjΣg are open sets and Σ̂(j)={pΣ̂||f̂(p)|<j}, and functions ujC(Σ̂(j);R3) such that we have f̂j°ϕj=f̂+ujon Σ̂(j) as well as ||̂muj||L(Σ̂(j),ĝ)0 as j for all mN0, so (i) is proven.

Moreover, sending j in Lemma 5.3 (iii) and using the smooth convergence on compact subsets, it follows B1(0)¯|Â|2dμ̂>0 and hence in particular Σ̂. The second statement in (ii) follows from the scaling invariance and the lower semicontinuity of the Willmore energy with respect to smooth convergence on compact subsets of R3, see [Citation19, Appendix B] for instance.

Let ξ(0,ĉ) be arbitrary. Using Equation(5.11) and arguing as in Equation(5.6) and Equation(5.7), we find (5.12) λjA(fj)+λjV(fj)C(ξ,K,σ,g)for all t[ξ,ĉ],jN,(5.12) which when combined with Equation(1.6) and Equation(5.10) immediately yields (5.13) ||tfj||C(ξ,K,σ,g)for all t[ξ,ĉ],jN.(5.13)

Now, as a consequence of Lemma B.1, we find (5.14) tmAjC(m,ξ,K,σ,g),tmAjL2(dμj)C(m,ξ,K,σ,g)for all t[ξ,ĉ],mN0,jN,(5.14) using Equation(5.10), Equation(5.11), and Equation(5.12). Similarly, using Lemma B.2 instead we obtain (5.15) tmHjC(m,ξ,K,σ,g),tmHjL2(dμj)C(m,ξ,K,σ,g)for all t[ξ,ĉ],jN,mN0.(5.15)

We will now use this to bound the derivative of the Lagrange multiplier. To that end, we observe that using I(fj)σ and integration by parts, we find λjA(fj)=3|Hj|2dμj+3|Aj0|2Hj2dμj12πσA(fj)12|Aj0|2Hjdμj3Hj12πσA(fj)122dμj..

Note that by Lemma 4.1 the denominator is bounded from below by some C(K,σ)>0. Using Equation(2.5), Equation(5.10),(5.11), Equation(5.13), Equation(5.14), and Equation(5.15), by direct computation we find (5.16) tλjA(fj)C(ξ,K,σ,g)for all t[0,ĉ],jN.(5.16)

Now, using I(fj)σ and Equation(2.5) we infer tλjV(fj)=C(σ)(tλjA(fj)A(fj)1212λjV(fj)A(fj)32tA(fj))=C(σ)(tλjA(fj)A(fj)12+12λjA(fj)A(fj)32Hjtfj,νjdμj).

Since κj(t,1)<ε1 for t[0,ĉ], we can apply Equation(3.11) with ρ = 1 to obtain A(fj)1C and hence using Equation(5.16), Equation(5.12), Equation(5.13), and Equation(2.10) we have tλjV(fj)C(ξ,K,σ,g)for all t[0,ĉ],jN.

For jN, we now define the flows f˜j:=fj°ϕj:=fj(·,ϕj(·)):(0,ĉ]×Σ̂(j)R3 and observe that they satisfy the L-estimates Equation(5.10) with A˜j instead of Aj and the evolution equation (5.17) tf˜j=[ΔH˜j|A˜j0|2H˜j+λj(3A(fj)H˜j2V(fj))]νj°ϕj.(5.17)

As in [Citation23, Proof of Theorem 6.2], the estimates for f˜j together with the C1-estimates for λjA(fj) and λjV(fj) can then be used to deduce that, after passing to a subsequence, the flows f˜j converge in C1([ξ,ĉ];Cm(P;R3)) for all mN and all PΣ̂ compact to a limit flow flim:[ξ,ĉ]×Σ̂R3. Moreover, we may assume νj°ϕjνlim in C1([ξ,ĉ];Cm(P;R3)) for all mN and all PΣ̂ compact, where νlim(t,·) is a smooth normal vector field along flim(t,·) for all t[ξ,ĉ], as well as λ(fj)A(fj)λlim,1andλ(fj)V(fj)λlim,2in C0([ξ,ĉ];R) as j.

Now, let PΣ̂ be a fixed compact set and let jN be large enough. Then, using Equation(5.17), Equation(2.9), and Equation(2.10) we find ξĉP|tf˜j|2dμ˜jdtξĉΣW0(fj)+λjσ1I(fj),tfjdμjdt=ξĉtW0(fj)dt.

In particular, taking j and using f˜jflim in C1([ξ,ĉ];Cm(P;R3)) for all mN, we find by Remark 2.7 (ii) ξP|tflim|dμlimdtlimj(W0(f(tj+rj4ξ))W0(f(tj+rj4ĉ)))=0.

Consequently, we have flimflim(ĉ,·)=limjfj(ĉ,ϕj(·))=limjf̂j°ϕj=f̂ in Cm(P;R3) for all mN and PΣ̂ compact.

We observe that ν̂:=νlim(ĉ,·) is a global and smooth normal vector field along f̂ and hence Σ̂ is orientable. Setting λ̂1:=λlim,1(ĉ),λ̂2:=λlim,2(ĉ) and using Equation(5.17) we find (ΔĤ|Â0|2Ĥ+3λ̂1Ĥ2λ̂2)ν̂=limjtf˜j(ĉ,·)=tflim(ĉ,·)=0on Σ̂, so f̂ is a Helfrich immersion and (iii) is proven.

For (iv), we now assume A(f̂j) as j. By Lemma 4.2, for all jN and ξ(0,ĉ), we have by Cauchy–Schwarz (5.18) ξĉλjA(fj)2dtC(K,σ)ξĉ(|tfj|2dμj+A(fj)1(|Aj0|2|Hj|dμj)2)dtC(K,σ,g)[W(f(tj+rj4ξ,·))W(f(tj+rj4ĉ,·))+ξĉA(fj)1dt],(5.18) where we estimated |Aj|3dμjC(ξ,K,σ,g) for all t[ξ,ĉ], using Equation(5.10) and Equation(2.10). Moreover, as a consequence of Equation(2.5) and Equation(2.10), for all t[0,ĉ] we have A(fj(t,·))A(fj(ĉ,·))tĉHjνj,tfjdμjdτ2ĉW(f0)+120ĉ|tfj|2dμjdτC(K,σ,g), so that A(fj(t,·))A(fj(ĉ))C(K,σ,g) for all t[0,ĉ] and hence the last term on the right hand side of Equation(5.18) goes to zero as j. Since tjT, the first term in Equation(5.18) converges to zero for j. Consequently 0=limjξĉλjA(fj)2dt=ξĉ|λlim,1|2dt, so that in particular, λ̂1=λlim,1(ĉ)=0. Moreover, as I(fj)σ, from Equation(1.2) we obtain λj(ĉ)V(fj(ĉ,·))=C(σ)λj(ĉ)A(fj(ĉ,·))A(fj(ĉ,·))120,j, so λ̂2=0 and thus f̂ is a Willmore immersion. □

5.3 The constrained Łojasiewicz–Simon inequality

In this section, we establish a Łojasiewicz–Simon inequality [Citation36–38]. While the unconstrained Willmore energy satisfies such an inequality [Citation26], the constraint of fixed isoperimetric ratio requires us to prove a refined estimate. To that end, we rely on the general framework of constrained or refined Łojasiewicz–Simon inequalities on submanifolds of Banach spaces [Citation27], see also [Citation32, Chapter 1, Section 1.2].

Theorem 5.5

(Constrained Łojasiewicz–Simon inequality). Let f:ΣgR3 be a Helfrich immersion with I(f)=σ(0,1) such that Hfconst. Then, there exist C,r>0 and θ(0,12] such that for all immersions hW4,2(Σg;R3) with ||hf||W4,2r and I(h)=σ we have |W0(h)W0(f)|1θCW0(h)λ(h)σ1I(h)L2(dμh).

The proof of this result is very similar to [Citation23, Section 7.1], so we will only provide full details on the differences. Throughout this section we will fix a smooth immersion f:ΣgR3 with I(f)=σ(0,1). The normal Sobolev spaces along f are Wk,2(Σg;R3):={ϕWk,2(Σg;R3)|Pϕ=ϕ}, for kN0, with L2(Σg;R3):=W0,2(Σg;R3). Here, the L2-inner product always has to be understood with respect to the measure μf and P denotes the normal projection along f, given by PX:=X,νfνf for any vector field X along f.

Let r > 0 be sufficiently small and let U˜:={ϕW4,2(Σg;R3)|ϕ=uνf for ||u||W4,2(Σg;R)<r}.

Consider the shifted energies, defined by W:U˜R,W(ϕ):=W0(f+ϕ),I:U˜R,I(ϕ):=I(f+ϕ).

Note that this is well-defined, since f+ϕ is an immersion for all ϕU˜ with r > 0 small enough, cf. [Citation23, Lemma 7.5 (i)]. The first main ingredient toward proving Theorem 5.5 is the analyticity of the energy and the constraint.

Lemma 5.6.

For r > 0 small enough, the following maps are analytic.

  1. the function U˜R,ϕW(ϕ);

  2. the function U˜L2(Σg;R3),ϕW0(f+ϕ)ρf+ϕ, where dμf+ϕ=ρf+ϕdμf;

  3. the function U˜R,ϕI(ϕ);

  4. the function U˜L2(Σg;R3),ϕI(f+ϕ)ρf+ϕ.

Proof.

Statements (i) and (ii) are exactly as in [Citation23, Lemma 7.6 (ii) and (iii)]. By [Citation23, Lemma 7.6 (i) and (iv)], the maps U˜R,ϕA(f+ϕ) and U˜R,ϕV(f+ϕ) are analytic and hence so is I by definition of the isoperimetric ratio and since A(f+ϕ)>0 for all ϕU˜. For statement (iv) recall from Proposition 2.4 that for ϕU˜ we have I(f+ϕ)=I(f+ϕ)(3A(f+ϕ)Hf+ϕνf+ϕ2V(f+ϕ)νf+ϕ).

We note that U˜C0(Σg;R3),ϕνf+ϕ is analytic by [Citation23, Lemma 7.5 (ii)] and U˜L2(Σg;R3),ϕHf+ϕνf+ϕ is analytic by [Citation26, Lemma 3.2 (iv)]. We have A(f+ϕ)>0 for all ϕU˜ and V(f+ϕ)0 by continuity for r > 0 sufficiently small, since I(f)=σ>0. This implies (iv). □

We now compute the first and second variations of W and I in terms of their H-gradients, see [Citation27, Section 5].

Lemma 5.7.

Let H:=L2(Σg;R3) and let r > 0 be sufficiently small. For each ϕU˜, the H-gradients of W and I are given by HW(ϕ)=PW0(f+ϕ)ρf+ϕ,HI(ϕ)=I(ϕ)(3A(f+ϕ)Hf+ϕ2V(f+ϕ))Pνf+ϕρf+ϕ.

Moreover, the Fréchet-derivatives of the H-gradient maps of W and I at u = 0 satisfy (HW)(0):W4,2(Σg;R3)L2(Σg;R3) is a Fredholm operator with index zero,(HI)(0):W4,2(Σg;R3)L2(Σg;R3)is compact.

Proof.

For ϕ,ψU˜, we have by Proposition 2.4 ddt|t=0W(ϕ+tψ)=W0(f+ϕ),ψdμf+ϕ=PW0(f+ϕ)ρf+ϕ,ψdμf=PW0(f+ϕ)ρf+ϕ,ψH.

Similarly, the statement for HI can be shown. The Fredholm property of (HW)(0) follows from Equation(1.1) and [Citation26, Lemma 3.3 and p. 356]. For the last statement, we observe that for all ϕW4,2(Σg;R3) we have ddt|t=0νf+tϕ,νf=12ddt|t=0|νf+tϕ|2=0. Hence, using Equation(2.6) with ξ=νf,ϕ and Proposition 2.4 we find (HI)(0)ϕ=ddt|t=0(I(tϕ)(3A(f+tϕ)Hf+tϕ2V(f+tϕ))ρf+tϕ)νf=I(f),ϕL2(dμf)(3A(f)Hf2V(f))νf3σA(f)2A(f),ϕL2(dμf)Hfνf+3σA(f)(Δνf,ϕ+|Af|2νf,ϕ)νf+2σV(f)2V(f),ϕL2(dμf)νfσ(3A(f)Hf2V(f))Hfνf,ϕνf, where we used I(f)=σ and ddt|t=0ρf+tϕ=Hfνf,ϕ by Equation(2.5). Since this only involves terms of order two or less in ϕW4,2(Σg;R3), the claim follows from the Rellich–Kondrachov Theorem, see for instance [Citation39, Theorem 2.34]. □

Proof of Theorem 5.5.

From the assumption Hfconst, it follows that I(f)0 and hence HI(0)0. As in [Citation23, Proposition 7.4], we can thus apply [Citation27, Corollary 5.2] to deduce that Theorem 5.5 is satisfied in normal directions, i.e. for the functional W with the constraint I=σ. With the methods from [Citation26, p. 357], one can then use the invariance of the energies under diffeomorphisms to conclude that Theorem 5.5 holds in all directions. □

As in [Citation26, Lemma 4.1], the Łojasiewicz–Simon inequality yields the following asymptotic stability result, see also [Citation23, Lemma 7.9] and [Citation40, Theorem 2.1] for related results in the context of constrained gradient flows in Hilbert spaces.

Lemma 5.8.

Let fW:ΣgR3 be a Helfrich immersion with I(fW)=σ(0,1),HfWconst. Let kN,k4,δ>0. Then there exists ε=ε(fW)>0 such that if f:[0,T)×ΣgR3 is a σ-isoperimetric Willmore flow satisfying

  1. ||f0fW||Ck,α<ε for some α>0;

  2. W0(f(t))W0(fW) whenever ||f(t)°Φ(t)fW||Ckδ for diffeomorphisms Φ(t):ΣgΣg;

then the flow exists globally, i.e. we may take T=. Moreover, as t, it converges smoothly after reparametrization by some diffeomorphisms Φ˜(t):ΣgΣg to a Helfrich immersion f, satisfying W0(fW)=W0(f) and ||ffW||Ckδ.

Note that by Lemma 2.6, fW above is a stationary solution to Equation(1.6)Equation(1.7). Consequently, the proof of Lemma 5.8 is a straightforward adaptation of [Citation23, Lemma 7.9], applying our Łojasiewicz–Simon inequality in Theorem 5.5 and can be safely omitted. As an important consequence one then finds the following convergence result by following the lines of [Citation26, Section 5] (see also [Citation23, Theorem 7.1]), which yields that in the case where Σ̂ is compact, below the explicit energy threshold no blow-ups or blow-downs may occur.

Theorem 5.9.

Let σ(0,1), let f:[0,T)×ΣgR3 be a maximal σ-isoperimetric Willmore flow with W(f0)<4πσ and let f̂:Σ̂R3 be a concentration limit as in Proposition 5.4. If Σ̂ has a compact component and Hf̂const, then f̂ is a limit under translation. Moreover, the flow exists for all times, i.e. T=, and, as t, converges smoothly after reparametrization to a Helfrich immersion f with W(f)=W(f̂).

Proof.

Let ĉ(0,1),tjT,(rj)jN(0,) and (xj)jNR3 be as in Proposition 5.4. By arguing as in [Citation15, Lemma 4.3], we may assume Σ̂=Σg and, by Proposition 5.4 (i), we hence have f̂j°Φjf̂ smoothly on Σg, where Φj:ΣgΣg are diffeomorphisms. Let ε=ε(f̂)>0 be as in Lemma 5.8. Then for some fixed j0N sufficiently large and any α(0,1), we may assume ||f̂j0°Φj0f̂||C4,α<ε. Moreover, the σ-isoperimetric Willmore flow f˜j0(t,·):=rj01(f(tj0+rj04t,·)xj0)°Φj0,t[0,rj04(Ttj0)), satisfies f˜j0(ĉ)=f̂j0°Φj0. Using Equation(2.10) and the invariance of the Willmore energy, for any t[0,rj04(Ttj0)), we find from Remark 2.7 (ii) W0(f˜j0(t))limsrj04(Ttj0)W0(f(tj0+rj04s))=limsTW0(f(s))=limkW0(f̂k)=W0(f̂), where we used the smooth convergence f̂k°Φkf̂ in the last step. Also note that we have I(f̂j0°Φj0)=I(f̂)=σ. Thus, by Lemma 5.8 the flow f˜j0 exists globally and, as t, converges smoothly after reparametrization by appropriate diffeomorphisms Φ˜(t):ΣgΣg to a Helfrich immersion f with W0(f)=W0(f̂), so W(f)=W(f̂) by Equation(1.1). Consequently, T= and for all ttj0 we have f(t,Φj0°Φ˜(rj04(ttj0)))=rj0f˜j0(rj04(ttj0),Φ˜(rj04(ttj0)))+xj0rj0f+xj0, as t smoothly on Σg. It remains to prove r(0,). To that end, we choose times sk:=rj04(tktj0+ĉrk4) for kN, such that we find sk as k, since tkT=. We thus obtain rj01(f(tk+ĉrk4,·)xj0)°Φj0°Φ˜(sk)=f˜j0(sk)°Φ˜(sk)fsmoothly as k.

Consequently, the diameters converge, so dk:=diamf(tk+ĉrk4)(Σg)rj0diamf(Σg) as k, whence limkdk(0,) since Σg is compact.

On the other hand, since f̂k°Φkf̂ smoothly and rkr[0,] the limit limkrk1dk=limkdiamf̂k=diamf̂(Σg)(0,) exists, and consequently rkr(0,) has to hold. □

6 Convergence for spheres

The goal of this section is to prove Theorem 1.2. To that end, we want to use the fact that compactness of the concentration limit Σ̂ yields convergence of the flow by Theorem 5.9. We first note that the desired compactness follows, if the area along the sequence f̂j in Proposition 5.4 remains bounded.

Lemma 6.1.

Let σ(0,1), let f:[0,T)×ΣgR3 be a maximal σ-isoperimetric Willmore flow with W(f0)<4πσ and let f̂j be as in Proposition 5.4. If supjNA(f̂j)<, then Σ̂ is compact.

Proof.

By Lemma 5.3 (iii), we have f̂j(Σg)B1(0) for all jN, where f̂j=fj(ĉ,·) with fj as in Equation(5.8). We now use the diameter bound [Citation34, Lemma 1.1] to estimate diamf̂j(Σg)CA(f̂j)W(f̂j), such that using the assumption, the invariances of the Willmore energy and the energy decay Equation(2.10), we find supjNdiamf̂j(Σg)<. Consequently, there exists R(0,) such that f̂j(Σg)BR(0)¯ for all jN. Letting j and using Proposition 5.4 (i) and the definition of smooth convergence on compact subset of R3, one then easily deduces f̂(Σ̂)BR(0)¯ and then, since f̂ is proper, compactness of Σ̂. □

We will now use Lemma 6.1 and Proposition 5.4 to conclude that if the concentration limit is non-compact, then it is not only a Helfrich, but even a Willmore immersion. In the spherical case, the classification in [Citation28] and the inversion strategy from [Citation16] will then yield a contradiction. Combined with Theorem 5.9, this will prove our main result.

Proof of Theorem 1.2.

Since Σg=S2 and σ(0,1), the existence of a unique, non-extendable σ-isoperimetric Willmore flow with initial datum f0 follows from Proposition 2.2 and Lemma 2.1 (ii). Moreover, by Remark 2.7 (i), the Willmore energy strictly decreases unless the flow is stationary, in which case global existence and convergence to a Helfrich immersion follow trivially. Thus, we may assume W(f0)<min{4πσ,8π}.

Let f̂:Σ̂R3 be a concentration limit as in Lemma 5.4. If Σ̂ is compact, we find Σ̂=S2 by [Citation15, Lemma 4.3] and long-time existence and convergence follow from Theorem 5.9 and the fact that Hf̂const by Lemma 2.1 (ii).

For the sake of contradiction, we assume that Σ̂ is not compact. Then we may assume A(f̂j) by Lemma 6.1. Consequently, by Proposition 5.4 we find that f̂:Σ̂R3 is a Willmore immersion with W(f̂)W(f0)<8π. The rest of the argument is as in [Citation23, Proof of Theorem 1.2]: Denote by I the inversion in a sphere with radius 1 centered at x0f̂(Σ̂) and let Σ¯:=I(f̂(Σ̂)){0}. Then Σ¯ is compact. By [Citation16, Lemma 5.1], Σ¯ is a smooth Willmore sphere with W(Σ¯)<8π and hence, using Bryant’s classification result [Citation28], has to be a round sphere. Thus, f̂(Σ̂) has to be either a plane or a sphere. Since Σ̂ is non-compact by assumption, this yields that f̂ has to parametrize a plane, a contradiction to Proposition 5.4 (ii).

Now the limit immersion f:S2R3 satisfies W(f)8π and solves Equation(1.4) for some λ1,λ2R. It remains to prove λ1,λ20. Arguing as in the proof of Lemma 2.6, we infer 2λ1A(f)+3λ2V(f)=0.

Now, V(f)0 by Lemma 2.1 (i) as σ(0,1) and also A(f)>0. Consequently, if one of λ1,λ2 is zero, then so is the other. In this case f is a Willmore sphere with W(f)8π. By Bryant’s result [Citation28], it then has to be a round sphere, so I(f)=1, a contradiction and hence λ1,λ20. □

Corollary 1.3 is an immediate consequence of the previous results.

Proof of Corollary 1.3.

By the assumption on the initial energy, Proposition 5.4 yields the existence of a suitable blow-up sequence and a concentration limit f̂ with the desired properties. If f̂ has constant mean curvature Ĥc, using Equation(2.2) EquationEquation (1.4) reads 12Ĥ32K̂Ĥλ1Ĥλ2=0.

If Σ̂ is compact, we conclude Ĥc0 and hence K̂ also has to be constant. But then f̂ has to parametrize a round sphere (see for instance [Citation30, Chapter V.Citation1]), a contradiction to I(f̂)=σ(0,1). Therefore, statement (a) follows from Theorem 5.9. If Σ̂ is not compact, we may assume A(f̂j) by Lemma 6.1 after passing to a subsequence. In this case, f̂ is a Willmore immersion by Proposition 5.4 (iv), yielding statement (b). □

7 An upper bound for β0

In this section, we will prove an upper bound for the minimal Willmore energy of spheres with isoperimetric ratio σ(0,1).

Theorem 7.1.

For every σ(0,1) we have β0(σ)<4πσ.

We remark that this estimate becomes sharp for σ1 since β0(1)=4π. On the other hand for σ(0,12], the statement follows since by [Citation7, Lemma 1] we have β0(σ)<8π for all σ(0,1). We will prove Theorem 7.1 by comparing energy and isoperimetric ratio of an ellipsoid. To that end, for a(0,1], we define the half-ellipse ca(t):=(0,acost,sint)T,t[π2,π2], in the y-z-plane in R3. By rotating the curve ca around the z-axis we obtain a particular type of ellipsoid, a prolate spheroid. More explicitly, we define fa(t,θ)=(acostcosθ,acostsinθ,sint)Tfor t[π2,π2],θ[0,2π].

Fortunately, its area, volume and also its Willmore energy can be explicitly computed without the use of elliptic integrals.

Lemma 7.2.

Let a(0,1). Then we have

  1. V(fa)=4π3a2;

  2. A(fa)=2πa(a+arcsin1a21a2);

  3. W(fa)=7π3+2π3a2+πaarcsin1a21a2.

Proof.

(i) and (ii) are standard formulas, see for instance [Citation41, Section 4.8]. For (iii), we observe that the mean curvature and the surface element of fa are given by dμfa=acosta2sin2t+cos2tdtdθ,Hfa=(1+a2)cos2t+2a2sin2ta(cos2t+a2sin2t)32, by standard formulas for surfaces of revolution, see for instance [Citation42, Section 3C]. In order to compute the Willmore energy, we thus have to evaluate the integral W(fa)=π2π2π2((1+a2)cos2t+2a2sin2t)2costa(cos2t+a2sin2t)52dt.

Substituting u=sint, this integral can then be explicitly computed yielding (iii). □

Proof of Theorem 7.1.

Clearly, we have β0(I(fa))W(fa). Moreover, by Lemma 7.2 and a short computation we have (7.1) I(fa)=8a(a+arcsin1a21a2)3for all a(0,1).(7.1)

An elementary computation yields I(fa)1 as a1 and similarly I(fa)0 as a0. Consequently, we have {I(fa)|a(0,1)}=(0,1) by a continuity argument.

Now, by Equation(7.1), we find for all a(0,1) W(fa)4πI(fa)=π6(14+4a2+6arcsin1a2a1a23(a+arcsin1a21a2)3a)=π6F(a), where the function F is negative for a(0,1) by Lemma 7.3. □

Lemma 7.3.

The function F:(0,1)R defined by F(a):=14+4a2+6arcsin1a2a1a23(a+arcsin1a21a2)3a satisfies F(a) < 0 for all a(0,1).

We will prove Lemma 7.3 in Sections A. A quick glimpse at the plot of F in illustrates that the statement of Lemma 7.3 is true. However, a rigorous proof seems to be surprisingly difficult, since the function combines trigonometric functions with polynomials and its graph becomes very flat near F(1)=0.

Fig. 1 The function F(a) in Lemma 7.3.

Fig. 1 The function F(a) in Lemma 7.3.

Acknowledgments

The author would like to thank Anna Dall’Acqua for helpful discussions and comments. In addition, the author is grateful to the referees for their careful reading and their valuable comments on the original manuscript.

Disclosure statement

The author reports there are no competing interests to declare.

Additional information

Funding

This work was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Grant 404870139 and the Austrian Science Fund (FWF) under Grant P 32788-N.

References

  • Canham, P. (1970). The minimum energy of bending as a possible explanation of the biconcave shape of the human red blood cell. J. Theor. Biol. 26(1):61–81. DOI: 10.1016/s0022-5193(70)80032-7.
  • Helfrich, W. (1973). Elastic properties of lipid bilayers: Theory and possible experiments. Zeitschrift für Naturforschung C 28(11):693–703. DOI: 10.1515/znc-1973-11-1209.
  • Hawking, S. W. (1968). Gravitational radiation in an expanding universe. J. Math. Phys. 9(4):598–604. DOI: 10.1063/1.1664615.
  • Friesecke, G., James, R. D., Müller, S. (2002). A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Commun. Pure Appl. Math. 55(11):1461–1506. DOI: 10.1002/cpa.10048.
  • Droske, M., Rumpf, M. (2004). A level set formulation for Willmore flow. Interfaces Free Bound. 6(3):361–378. DOI: 10.4171/IFB/105.
  • Willmore, T. J. (1993). Riemannian Geometry. Oxford Science Publications. New York: The Clarendon Press, Oxford University Press.
  • Schygulla, J. (2012). Willmore minimizers with prescribed isoperimetric ratio. Arch. Ration. Mech. Anal. 203(3):901–941. DOI: 10.1007/s00205-011-0465-4.
  • Keller, L. G. A., Mondino, A., Rivière, T. (2014). Embedded surfaces of arbitrary genus minimizing the Willmore energy under isoperimetric constraint. Arch. Ration. Mech. Anal. 212(2):645–682. DOI: 10.1007/s00205-013-0694-9.
  • Mondino, A., Scharrer, C. (2021). A strict inequality for the minimization of the Willmore functional under isoperimetric constraint. Adv. Calc. Var. 16(3):529–540. DOI: 10.1515/acv-2021-0002.
  • Scharrer, C. (2022). Embedded Delaunay tori and their Willmore energy. Nonlinear Anal., 223:Paper No. 113010, 23. DOI: 10.1016/j.na.2022.113010.
  • Kusner, R., McGrath, P. (2023). On the Canham problem: bending energy minimizers for any genus and isoperimetric ratio. Arch. Ration. Mech. Anal. 247(1):Paper No. 10, 14.
  • Li, P., Yau, S. T. (1982). A new conformal invariant and its applications to the Willmore conjecture and the first eigenvalue of compact surfaces. Invent. Math. 69(2):269–291.
  • McCoy, J., Wheeler, G. (2013). A classification theorem for Helfrich surfaces. Math. Ann. 357(4):1485–1508. DOI: 10.1007/s00208-013-0944-z.
  • Kuwert, E., Schätzle, R. (2002). Gradient flow for the Willmore functional. Commun. Anal. Geom. 10(2):307–339. DOI: 10.4310/CAG.2002.v10.n2.a4.
  • Kuwert, E., Schätzle, R. (2001). The Willmore flow with small initial energy. J. Differ. Geom. 57(3):409–441.
  • Kuwert, E., Schätzle, R. (2004). Removability of point singularities of Willmore surfaces. Ann. Math. (2) 160(1):315–357. DOI: 10.4007/annals.2004.160.315.
  • McCoy, J., Wheeler, G., Williams, G. (2011). Lifespan theorem for constrained surface diffusion flows. Math. Z. 269(1–2):147–178. DOI: 10.1007/s00209-010-0720-7.
  • Wheeler, G. (2012). Surface diffusion flow near spheres. Calc. Var. Partial Differ. Equ. 44(1–2):131–151. DOI: 10.1007/s00526-011-0429-4.
  • Dall’Acqua, A., Müller, M., Schätzle, R., Spener, A. The Willmore flow of tori of revolution, 2020. arXiv:2005.13500. To appear in Anal. PDE.
  • McCoy, J., Wheeler, G. (2016). Finite time singularities for the locally constrained Willmore flow of surfaces. Commun. Anal. Geom. 24(4):843–886. DOI: 10.4310/CAG.2016.v24.n4.a7.
  • Blatt, S. (2019). A note on singularities in finite time for the L2 gradient flow of the Helfrich functional. J. Evol. Equ. 19(2):463–477. DOI: 10.1007/s00028-019-00483-y.
  • Jachan, F. (2014). Area preserving Willmore flow in asymptotically Schwarzschild manifolds. PhD thesis, FU Berlin.
  • Rupp, F. (2023). The volume-preserving Willmore flow. Nonlinear Anal. 230:Paper No. 113220, 30. DOI: 10.1016/j.na.2023.113220.
  • Huisken, G. (1987). The volume preserving mean curvature flow. J. Reine Angew. Math. 382:35–48.
  • Blatt, S. (2009). A singular example for the Willmore flow. Analysis (Munich) 29(4):407–430. DOI: 10.1524/anly.2009.1017.
  • Chill, R., Fašangová, E., Schätzle, R. (2009). Willmore blowups are never compact. Duke Math. J. 147(2):345–376. DOI: 10.1215/00127094-2009-014.
  • Rupp, F. (2020). On the Łojasiewicz–Simon gradient inequality on submanifolds. J. Funct. Anal. 279(8):108708. DOI: 10.1016/j.jfa.2020.108708.
  • Bryant, R. L. (1984). A duality theorem for Willmore surfaces. J. Differ. Geom. 20(1):23–53.
  • Kuwert, E., Schätzle, R. (2012). The Willmore functional. In: Topics in Modern Regularity Theory, volume 13 of CRM Series. Pisa: Ed. Norm., 1–115.
  • Hopf, H. (1983). Differential Geometry in the Large, volume 1000 of Lecture Notes in Mathematics. Berlin: Springer-Verlag.
  • Aleksandrov, A. D. (1962). Uniqueness theorems for surfaces in the large. V. Amer. Math. Soc. Transl. (2) 21:412–416.
  • Rupp, F. (2022). Constrained gradient flows for Willmore-type functionals. PhD thesis, Universität Ulm.
  • Lee, J. M. (2013). Introduction to Smooth Manifolds, 2nd ed., volume 218 of Graduate Texts in Mathematics. New York: Springer.
  • Simon, L. (1993). Existence of surfaces minimizing the Willmore functional. Commun. Anal. Geom. 1(2):281–326. DOI: 10.4310/CAG.1993.v1.n2.a4.
  • Marques, F. C., Neves, A. (2014). Min-max theory and the Willmore conjecture. Ann. Math. (2) 179(2):683–782. DOI: 10.4007/annals.2014.179.2.6.
  • Łojasiewicz, S. (1963). Une propriété topologique des sous-ensembles analytiques réels. In: Les Équations aux Dérivées Partielles (Paris). Paris: Éditions du Centre National de la Recherche Scientifique, 87–89.
  • Łojasiewicz, S. (1965). Sur les ensembles semi-analytiques. I.H.E.S., Bures-sur-Yvette.
  • Simon, L. (1983). Asymptotics for a class of nonlinear evolution equations, with applications to geometric problems. Ann. Math. (2) 118(3):525–571. DOI: 10.2307/2006981.
  • Aubin, T. (1982). Nonlinear Analysis on Manifolds. Monge–Ampère Equations, volume 252 of Grundlehren der Mathematischen Wissenschaften. New York: Springer-Verlag.
  • Lengeler, D. (2018). Asymptotic stability of local Helfrich minimizers. Interfaces Free Bound. 20(4):533–550. DOI: 10.4171/IFB/411.
  • Zwillinger, D. (2012). CRC Standard Mathematical Tables and Formulae, 32nd ed. Boca Raton, FL: CRC Press.
  • Kühnel, W. (2002). Differential Geometry, volume 16 of Student Mathematical Library. Providence, RI: American Mathematical Society. Curves—surfaces—manifolds, Translated from the 1999 German original by Bruce Hunt.
  • Cohn, P. M. (2003). Basic Algebra. Groups, Rings and Fields. London: Springer-Verlag.
  • Simons, J. (1968). Minimal varieties in Riemannian manifolds. Ann. Math. (2) 88:62–105. DOI: 10.2307/1970556.

A

Proof of Lemma 7.3

This section is devoted to proving Lemma 7.3. The idea is to make a change of variables, such that the problem is equivalently formulated in terms of a polynomial in x,cosx and sinx. Then, we use the power series representation of the Cosine and Sine functions, to reduce the problem to the question if a certain polynomial has roots in a given interval. This last point can then be discussed by studying the Sturm chain of the polynomial.

Proof of Lemma 7.3.

For x(0,π2) we consider the function G(x):=F(cosx)sin3xcosx, so that expanding we find G(x)=3x39x2cosxsinx+6xsin2x9xcos2xsin2x+14cosxsin3x+cos3xsin3x.

We observe that F(a) < 0 for all a(0,1) is equivalent to G(x) < 0 for all x(0,π2).

Using the power series expansion of the Cosine and Taylor’s theorem with the Lagrange form of the remainder, for any NN we infer cosx=k=0N(1)k(2k)!x2k+1(2N+1)!cos(2N+1)(ξ)x2N+1, for some ξ=ξ(N)(0,x)(0,π2). An induction argument yields cos(2N+1)=(1)N+1sin, so the remainder has a sign, depending on the parity of N. Hence, denoting TcosN(x):=k=0N(1)k(2k)!x2k, we infer (A.1) Tcos2n+1(x)<cosx<Tcos2n(x)for all x(0,π2),nN0.(A.1)

By similar arguments, defining TsinN(x):=k=0N(1)k(2k+1)!x2k+1 and using sin(2N+2)=(1)N+1sin we find (A.2) Tsin2n+1(x)<sinx<Tsin2n(x)for all x(0,π2),nN0.(A.2)

We will now use Equation(A.1) and Equation(A.2) to estimate G. For x(0,π2), we have G(x)<3x39x2Tcos3(x)Tsin3(x)+6xTsin2(x)29xTcos3(x)2Tsin3(x)2+14Tcos2(x)Tsin2(x)3+Tcos2(x)3Tsin2(x)3=:k(x).

Now, we observe that k(x) is polynomial of degree 27. Using Mathematica, we find that this can be simplified to k(x)=x95852528640000q(x), for the degree 18 polynomial q(x)=984711168000+660770611200x2209922048000x4+40156646400x65069859840x8+437184000x1025717120x12+994464x1422944x16+241x18.

By substituting x2=z, in order to prove G(x) < 0 for x(0,π2) it thus suffices to show that p(z)=984711168000+660770611200z209922048000z2+40156646400z35069859840z4+437184000z525717120z6+994464z722944z8+241z9<0 for all z(0,π24). To this end, one may compute the Sturm chain of the polynomial p (see [Citation43, Theorem 8.8.15] for instance), to find that there exist no real roots of p in the interval [0,3][0,π24]. Consequently, since p(0)<0, we find p(z) < 0 for all z[0,π24], and hence the claim follows. □

B

Higher order evolution

In this section, we will prove a higher order version of Proposition 3.3. To this end, we follow [Citation14, Citation15] and denote by ϕ*ψ any multilinear form, depending on ϕ and ψ in a universal bilinear way, where ϕ,ψ are tensors on Σg. In particular, we have |ϕ*ψ|C|ϕ||ψ| for a universal constant C > 0 and (ϕ*ψ)=ϕ*ψ+ϕ*ψ. Moreover, for mN0 and rN,r2 we denote by Prm(A) any term of the type Prm(A)=i1++ir=mi1A**irA.

In addition, for r = 1 we extend this definition by denoting by P1m(A) any contraction of mA with respect to the metric g.

With this notation, we observe that along an isoperimetric Willmore flow the covariant derivatives of the second fundamental form A also satisfy a 4-th order evolution equation.

Lemma B.1.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow. Then for all mN0 we have t(mA)+Δ2(mA)=P3m+2(A)+P5m(A)+λA(f)(P1m+2(A)+P3m(A))+λV(f)P2m(A).

Proof.

We observe tf=ξν with (B.1) ξ=ΔH+P30(A)+λA(f)P10(A)2λV(f).(B.1)

For m = 0, we thus find by Equation(2.7) tA=2ξ+A*A*ξ=Δ2A+P32(A)+P50(A)+λA(f)(P12(A)+P30(A))+λV(f)P20(A), where we used 2ΔH=Δ2A+P32(A) as a consequence of Simons’ identity [Citation44]. Assume the statement is true for m1. By [Citation14, Lemma 2.3] with ϕ=mA and the fact that we are in codimension one, we find tm+1A+Δ2m+1A=(P3m+2(A)+P5m(A)+λA(f)(P1m+2(A)+P3m(A))+λV(f)P2m(A))+i+j+k=3iA*jA*k+mA+A*ξ*mA+A*ξ*mA=P3m+3(A)+P5m+1(A)+λA(f)(P1m+3(A)+P3m+1(A))+λV(f)P2m+1(A), using Equation(B.1) in the last step. □

With similar computations as above, one finds the following

Lemma B.2.

Let σ(0,1) and let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow. The for all mN0 we have t(mH)+Δ2(mH)=P3m+2(A)+P5m(A)+λA(f)(P1m+2(A)+P3m(A))+λV(f)P2m(A).

We can now prove the following higher order analogue of Proposition 3.3.

Proposition B.3.

Let σ(0,1), let f:[0,T)×ΣgR3 be a σ-isoperimetric Willmore flow and let γ be as in Equation(3.3). Then for all mN0,s2m+4 and ϕ=mA we have ​ddt|ϕ|2γsdμ+12|2ϕ|2γsdμC(λ2A(f)2+|λ|43|V(f)|43+||A||L([γ>0])4)|ϕ|2γsdμ+C(1+λ2A(f)2+|λ|43|V(f)|43+||A||L([γ>0])4)[γ>0]|A|2dμ, where C=C(s,m,Λ)>0.

In order to prove Proposition B.3, we first recall the following

Lemma B.4.

[Citation14, Lemma 3.2] Let f:[0,T)×ΣgR3 be a normal variation, tf=ξν. Let ϕ be a (l0)-tensor satisfying tϕ+Δ2ϕ=Y. Then for any γC2([0,T)×Σg) and s4 we have ​ddt|ϕ|2γsdμ+|2ϕ|2γsdμ2Y,ϕγsdμ+A*ϕ*ϕ*ξγsdμ+|ϕ|2sγs1tγdμ+C|ϕ|2γs4(|γ|4+γ2|2γ|2)dμ+C|ϕ|2(|A|2+|A|4)γsdμ, where C=C(s).

Proof of Proposition B.3.

In the following, note that the value of C=C(s,m,Λ) is allowed to change from line to line. We apply Lemma B.4 with Y=tϕ+Δ2ϕ,ξ=P12(A)+P30(A)+λA(f)P10(A)2λV(f) by Equation(B.1) and estimate the terms on the right hand side. Using ϕ=mA and Lemma B.1, we thus have (B.2) 2Y,ϕγsdμ+A*ϕ*ϕ*ξγsdμ+C|ϕ|2(|A|2+|A|4)γsdμ=(P3m+2(A)+P5m(A))*ϕγsdμ+λA(f)(P1m+2(A)+P3m(A))*ϕγsdμ+λV(f)P2m(A)*ϕγsdμ.(B.2)

Moreover, by Equation(3.3) we find (B.3) |ϕ|2γs1tγdμ=|ϕ|2γs1Dγ˜°f,ν(ΔH|A0|2H+3λA(f)H2λV(f))dμ.(B.3)

We proceed by estimating all the terms involving λ in Equation(B.2) and Equation(B.3). For the first λ-term in Equation(B.2), since |P1m+2(A)|C|2ϕ| we find for every ε>0 λA(f)P1m+2(A)*ϕγsdμε|2ϕ|2γsdμ+C(ε)λ2A(f)2|ϕ|2γsdμ.

For the second term, we use [Citation14, Corollary 5.5] with k = m, r = 4 to obtain λA(f)P3m(A)*ϕγsdμCλA(f)||A||L([γ>0])2(|ϕ|2γsdμ+[γ>0]|A2|dμ).

The last λ-term in Equation(B.2) can be estimated by [Citation14, Corollary 5.5] with k = m and r = 3, yielding λV(f)P2m(A)*ϕγsdμCλV(f)||A||L([γ>0])(|ϕ|2γsdμ+[γ>0]|A|2dμ).

Now for the first λ-term in Equation(B.3), we use Young’s inequality twice to obtain λA(f)|ϕ|2γs1Dγ˜°f,νHdμCλ2A(f)2|ϕ|2γsdμ+C|ϕ|2|A|2γs2dμCλ2A(f)2|ϕ|2γsdμ+C||A||L([γ>0]))4|ϕ|2γsdμ+C|ϕ|2γs4dμ.

For the second λ-term in Equation(B.3), we can use Young’s inequality with p=43 and q = 4 to estimate λV(f)|ϕ|2γs1Dγ˜°f,νdμC|λ|43|V(f)|43|ϕ|2γsdμ+C|ϕ|2γs4dμ.

Choosing ε>0 sufficiently small and absorbing, by Lemma B.4, Equation(B.2) and Equation(B.3) (B.4) ​ddt|ϕ|2γsdμ+34|2ϕ|γsdμ(P3m+2(A)+P5m(A))*ϕγsdμ+|ϕ|2γs1Dγ˜°f,ν(ΔH|A0|2H)dμ+C(λ2A(f)2+|λ|43|V(f)|43+||A||L([γ>0])4)(|ϕ|2γsdμ+[γ>0]|A|2dμ)+C|ϕ|2γs4+|ϕ|2γs4(|γ|4+γ2|2γ|2)dμ,(B.4) where we used Young’s inequality to obtain the correct powers of λ and ||A||L([γ>0]). Now, all the terms involving λ on the right hand side of Equation(B.4) are as in the statement. For the second and the last term in Equation(B.4), one may proceed exactly as in the proof of [Citation14, Proposition 3.3]. This way, one creates additional terms which can be estimated by |ϕ|2γs4dμ+|ϕ|2γs2dμε|2ϕ|2γsdμ+Cε[γ>0]|A|2γs42mdμ, for every ε>0, using twice the interpolation inequality [Citation14, Corollary 5.3] (which trivially also holds in the case k=m=0). The first term on the right hand side of Equation(B.4) can then be estimated by means of [Citation14, (4.15)]. After choosing ε>0 small enough and absorbing, the claim follows since s2m+4 and γ1. □