This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Rotating Drops with Helicoidal Symmetry

Bennett Palmer and Oscar M. Perdomo O. Perdomo
Department of Mathematics
Central Connecticut State University
New Britain, CT 06050 USA
e-mail: perdomoosm@ccsu.edu B. Palmer
Department of Mathematics
Idaho State University
Pocatello, ID, 83209 USA
e-mail: palmbenn@isu.edu
(Date: September 6, 2025)
Abstract.

We consider helicoidal immersions in 𝐑3{\bf R}^{3} whose axis of symmetry is the zz-axis that are solutions of the equation 2H=Λ0aR222H=\Lambda_{0}-a\frac{R^{2}}{2} where HH is the mean curvature of the surface, RR is the distance form the point in the surface to the zz-axis and aa is a real number. We refer to these surfaces as helicoidal rotating drops. We prove the existence of properly immersed solutions that contain the zz-axis. We also show the existence of several families of embedded examples. We describe the set of possible solutions and we show that most of these solutions are not properly immerse and are dense in the region bounded by two concentric cylinders. We show that all properly immersed solutions, besides being invariant under a one parameter helicoidal group, they are invariant under a cyclic group of rotations of the variables xx and yy.

The second variation of energy for the volume constrained problem with Dirichlet boundary conditions is also studied.

2000 Mathematics Subject Classification:
53C42, 53C10

[Uncaptioned image]

1. Introduction

In this paper we study the equilibrium shape of a rotating liquid drop or liquid film, which is invariant under a helicoidal motion of the three dimensional Euclidean space. The subject of rotating drops has been studied by many authors, including Chandrasekhar [2], Brown and Scriven [1], Solonikov [8] and many others. Our main objective here is to use a new construction, recently developed by the second author [4], to construct an abundant supply of examples. This construction is closely related to the Delaunay’s classical construction of the axially symmetric constant mean curvature surfaces whose generating curves are produced by rolling a conic section. A special case of the type of surface which we study here, occurs when the rotating drop is a cylinder over a plane curve. We treat that case in detail in [7].

If a rigid object is moved from one position in space to another, this repositioning can be realized via a helicoidal motion of 𝐑3{\bf R}^{3}. If then, this motion is successively repeated, one arrives at a configuration which is invariant under a helicoidal motion. The simple idea that helicoidal motions give all repeated motions of a rigid object, known in the physical sciences as Pauling’s Theorem, is behind many of the occurrences of helicoidal symmetry in nature, since it allows for extensive growth with a minimal amount of information.

We will consider the equilibrium shape of a liquid drop rotating with a constant angular velocity Ω\Omega about a vertical axis. The surface of the drop, which we denote by Σ\Sigma, is represented as a smooth surface. The bulk of the drop is assumed to be occupied by an imcomressible liquid of a constant mass density ρ1\rho_{1} while the drop is surrounded by a fluid of constant mass density ρ2\rho_{2}. Since the drop is liquid, its free surface energy is proportional to its surface area 𝒜{\mathcal{A}} and we take the constant of proportionality to be one. The rotation contributes a second energy term of the form Ω2Δ-\Omega^{2}\Delta{\mathcal{I}}, where Δ\Delta{\mathcal{I}} is difference of moments of inertia about the vertical axis,

Δ:=(ρ1ρ2)UR2𝑑v.\Delta{\mathcal{I}}:=(\rho_{1}-\rho_{2})\int_{U}R^{2}\>dv\>.

This term represents twice the rotational kinetic energy.

The total energy is thus of the form

(1.1) :=𝒜Ω22Δ+Λ0𝒱,{\mathcal{E}}:={\mathcal{A}}-\frac{\Omega^{2}}{2}{\Delta\mathcal{I}}+\Lambda_{0}{\mathcal{V}}\>,

where 𝒱{\mathcal{V}} denotes the volume of the drop and Λ0\Lambda_{0} is a Lagrange multiplier. Let Δρ:=ρ1ρ2\Delta\rho:=\rho_{1}-\rho_{2}, then by introducing a constant a:=(Δρ)Ω2a:=(\Delta\rho)\Omega^{2}, we can write the functional in the form

a,Λ0=𝒜a2UR2𝑑V+Λ0𝒱,{\mathcal{E}}_{a,\Lambda_{0}}={\mathcal{A}}-\frac{a}{2}\int_{U}R^{2}\>dV+\Lambda_{0}{\mathcal{V}}\>,

where UU is the three dimensional region occupied by the bulk of the drop, and R:=x2+y2R:=\sqrt{x^{2}+y^{2}}.

Since we want to consider both embedded and immersed surfaces, we will precisely define the last two terms in the energy in the following way. First define vector fields on 𝐑3{\bf R}^{3} by

W=R4/16=R24(x1,x2,0),W0=R2/4=(1/2)(x1,x2,0).W=\nabla^{\prime}R^{4}/16=\frac{R^{2}}{4}(x_{1},x_{2},0)\>,W_{0}=\nabla^{\prime}R^{2}/4=(1/2)(x_{1},x_{2},0)\>.

If \nabla^{\prime}\cdot denotes the divergence operator on 𝐑3{\bf R}^{3}, then it is easily checked that W0=1\nabla^{\prime}\cdot W_{0}=1 and W=R2\nabla^{\prime}\cdot W=R^{2} hold. We, then define

𝒱:=ΣW0ν𝑑Σ,UR2𝑑v:=ΣWν𝑑Σ.{\mathcal{V}}:=\int_{\Sigma}W_{0}\cdot\nu\>d\Sigma\>,\quad\int_{U}R^{2}\>dv:=\int_{\Sigma}W\cdot\nu\>d\Sigma\>.

The definitions are valid so long as Σ\Sigma is immersed and oriented.

We will next derive the first variation of the functional given above. Let Xϵ=X+ϵ(ψν+T)+X_{\epsilon}=X+\epsilon(\psi\nu+T)+... be a variation of XX, where ψ\psi is a smooth function, ν\nu is the unit normal to the surface and TT is a tangent vector field along Σ\Sigma. The first variation formula for the area gives,

δ𝒜\displaystyle\delta{\mathcal{A}} =\displaystyle= Σ2Hψ𝑑Σ+ΣTn𝑑s\displaystyle-\int_{\Sigma}2H\psi\>d\Sigma+\oint_{\partial\Sigma}T\cdot n\>ds
=\displaystyle= Σ2Hψ𝑑Σ+Σ𝑑X×νT.\displaystyle-\int_{\Sigma}2H\psi\>d\Sigma+\oint_{\partial\Sigma}dX\times\nu\cdot T\>.

We will show in the appendix that

(1.2) δΩR2𝑑V=ΣψR2𝑑Σ+Σ𝑑X×WδX,\delta\int_{\Omega}R^{2}\>dV=\int_{\Sigma}\psi R^{2}\>d\Sigma+\oint_{\partial\Sigma}dX\times W\cdot\delta X\>,

where WW is a vector field satisfying W=R2\nabla^{\prime}\cdot W=R^{2} on 𝐑3{\bf R}^{3}, and it is well known that the first variation of volume is

(1.3) δ𝒱=Σψ𝑑Σ+Σ𝑑X×W0δX.\delta{\mathcal{V}}=\int_{\Sigma}\psi\>d\Sigma+\oint_{\partial\Sigma}dX\times W_{0}\cdot\delta X\>.

By combining the last three formulas, we arrive at

(1.4) δa,Λ0\displaystyle\delta{\mathcal{E}}_{a,\Lambda_{0}} =\displaystyle= Σ(2Ha2R2+Λ0)ψ𝑑Σ\displaystyle\int_{\Sigma}(-2H-\frac{a}{2}R^{2}+\Lambda_{0})\psi\>d\Sigma
+Σ1𝑑X×(νa2W+Λ0W0)δX\displaystyle+\oint_{\partial\Sigma_{1}}dX\times(\nu-\frac{a}{2}W+\Lambda_{0}W_{0})\cdot\delta X

Regardless of the boundary conditions, a necessary condition for an equilibrium is that in the interior of Σ\Sigma, there holds

(1.5) 2H=a2R2+Λ0.2H=-\frac{a}{2}R^{2}+\Lambda_{0}\>.

If we assume that the surface has free boundary contained in a supporting surface SS having outward unit normal NN, then the admissible variations must satisfy the condition δXN0\delta X\cdot N\equiv 0 on Σ\partial\Sigma. In order that the boundary integral in (1.4) vanishes for all admissible variations, we must have that dX×νdX\times\nu parallel to NN along the boundary which means that the surface Σ\Sigma meets the supporting surface SS in a right angle.

We now assume that an equilibrium surface Σ\Sigma, i.e. a surface satisfying (1.5) is invariant under a helicoidal motion:

(1.6) (x1+ix2,x3)(eiωt(x1+ix2),x3+t),(x_{1}+ix_{2},x_{3})\mapsto(e^{-i\omega t}(x_{1}+ix_{2}),x_{3}+t)\>,

and we will derive a conservation law which characterizes the equilibrium surfaces. We do not assume that the angular velocity ω\omega which determines the pitch of the helicoidal surface is the same as the angular velocity Ω\Omega appearing above,

Let Σ1\Sigma_{1} denote the compact region in Σ\Sigma bounded on the sides by two integral curves C1C_{1} and C2C_{2} of the Killing field 𝒦(X):=ωE3×X+E3{\mathcal{K}}(X):=-\omega E_{3}\times X+E_{3} and bounded below and above by the horizontal planes x3=0x_{3}=0 and x3=2π/ωx_{3}=2\pi/\omega. Then, Σ1\Sigma_{1} is a compact surface with oriented boundary C1+α1C2α2C_{1}+\alpha_{1}-C_{2}-\alpha_{2} where α1\alpha_{1} and α2\alpha_{2} are congruent arcs in the planes x3=2π/ωx_{3}=2\pi/\omega, x3=0x_{3}=0, respectively. By the calculations in the appendix, we have, using (1.5)

(1.7) δ[Σ0]=Σ1𝑑X×(νa2W+Λ0W0)δX.\delta{\mathcal{E}}[\Sigma_{0}]=\oint_{\partial\Sigma_{1}}dX\times(\nu-\frac{a}{2}W+\Lambda_{0}W_{0})\cdot\delta X\>.

If we take the variation with δX=E3\delta X=E_{3} then, since E3E_{3} generates a translation, the first variation will vanish. Consequently, we obtain

(1.8) 0=Σ1𝑑X×(νa2R416+Λ0R24)E3.0=\oint_{\partial\Sigma_{1}}dX\times(\nu-\frac{a}{2}\nabla^{\prime}\frac{R^{4}}{16}+\Lambda_{0}\nabla^{\prime}\frac{R^{2}}{4})\cdot E_{3}\>.

Note that the integration over the 1-chain α1α2\alpha_{1}-\alpha_{2} yields zero since the two arcs are congruent and are traversed in opposite directions. On CiC_{i}, i=1,2i=1,2, we have dX/dt=(1)i+1𝒦dX/dt=(-1)^{i+1}{\mathcal{K}}.

dX×(νa2R416+Λ0R24)E3\displaystyle dX\times(\nu-\frac{a}{2}\nabla^{\prime}\frac{R^{4}}{16}+\Lambda_{0}\nabla^{\prime}\frac{R^{2}}{4})\cdot E_{3} =\displaystyle= (1)i+1(ωE3×X+E3)×(νa2R416+Λ0R24)E3dt\displaystyle(-1)^{i+1}(-\omega E_{3}\times X+E_{3})\times(\nu-\frac{a}{2}\nabla^{\prime}\frac{R^{4}}{16}+\Lambda_{0}\nabla^{\prime}\frac{R^{2}}{4})\cdot E_{3}\>dt
=\displaystyle= (1)i(ωE3×X+E3)×E3(νa2R416+Λ0R24)dt\displaystyle(-1)^{i}(-\omega E_{3}\times X+E_{3})\times E_{3}\cdot(\nu-\frac{a}{2}\nabla^{\prime}\frac{R^{4}}{16}+\Lambda_{0}\nabla^{\prime}\frac{R^{2}}{4})\>dt
=\displaystyle= (1)i+1ω(x1,x2,0)(νa8R2(x1,x2,0)+Λ02(x1,x2,0))dt\displaystyle(-1)^{i+1}\omega(x_{1},x_{2},0)\cdot(\nu-\frac{a}{8}R^{2}(x_{1},x_{2},0)+\frac{\Lambda_{0}}{2}(x_{1},x_{2},0))\>dt
=\displaystyle= (1)i+1ω((Qx3ν3)a8R4+Λ0R22)dt\displaystyle(-1)^{i+1}\omega\bigl{(}(Q-x_{3}\nu_{3})-\frac{a}{8}R^{4}+\frac{\Lambda_{0}R^{2}}{2}\bigr{)}\>dt

where Q=XνQ=X\cdot\nu is the support function of the surface. Setting Q^:=Qx3ν3{\hat{Q}}:=Q-x_{3}\nu_{3}, We can conclude from this that the integral

Ci(Q^a8R4+Λ0R22)𝑑t,\int_{C_{i}}\bigl{(}{\hat{Q}}-\frac{a}{8}R^{4}+\frac{\Lambda_{0}R^{2}}{2}\bigr{)}\>dt\>\>,

is independent of ii. Also, it is easily checked that the integrand is, in fact, constant on each helix CiC_{i} and we obtain the result that

(1.9) 2Q^+Λ0R2aR44constant.2{\hat{Q}}+\Lambda_{0}R^{2}-a\frac{R^{4}}{4}\equiv{\rm constant}\>.
Proposition 1.1.

Let Σ\Sigma be a helicoidal surface. A necessary and sufficient condition that Σ\Sigma is a critical point for the functional a,Λ0{\mathcal{E}}_{a,\Lambda_{0}} is that (1.9) holds.

Proof.

The necessity was shown above, we now show that the condition is sufficient. We can assume that the helicoidal symmetry group of the surface fixes the vertical axis.

Any helicoidal surface arises as the orbit of a planar “ generating curve” α\alpha under a helicoidal motion. We let ss be the arc length coordinate of α\alpha and we let tt denote a coordinate for the helices which are the orbits of points in α\alpha. Local calculations which can be found in [4], show that the mean curvature HH and the third component of the normal ν3\nu_{3} are functions of ss alone. Also, it is clear that the function R2R^{2} only depends on ss.

It is easy to see that if ν3\nu_{3} vanishes on any arc of α\alpha, then this arc is necessarily circular. It is clear that the orbit of a circular arc satisfying 1.9 is a critical surface for the functional a,Λ0{\mathcal{E}}_{a,\Lambda_{0}}. Now consider a connected arc ηα\eta\subset\alpha on which (say) ν3>0\nu_{3}>0 holds almost everywhere. If (1.5) does not hold on α\alpha, we can assume, by replacing α\alpha with a sub-arc if necessary, that 2HaR2/2+Λ0>0-2H-aR^{2}/2+\Lambda_{0}>0 holds almost everywhere on α\alpha also. Let Σ1\Sigma_{1} denote the compact domain consisting of the orbit of the arc α\alpha, for 0t2π/ω0\leq t\leq 2\pi/\omega. The boundary of Σ1\Sigma_{1} consists of two helices C1C_{1}, C2C_{2} together with two arcs α1\alpha_{1}, α2\alpha_{2} both congruent to α\alpha.

We take the first variation of a,Λ0{\mathcal{E}}_{a,\Lambda_{0}} with the variation field being the constant vector E3E_{3}. Since E3E_{3} is the generator of a one parameter family of isometries, this first variation vanishes. We express the first variation as in (1.4). Since (1.9) holds, the contributions to the boundary integral is zero since it is given by the right hand side of (1.8)(\ref{bint})vanishes and the integrals over α1\alpha_{1} and α2\alpha_{2} cancel each other since these arcs are congruent and are traversed in opposite directions. We then obtain from the calculations given above, that

0=Σ1(2Ha2R2+Λ0)ν3𝑑Σ,0=\int_{\Sigma_{1}}(-2H-\frac{a}{2}R^{2}+\Lambda_{0})\nu_{3}\>d\Sigma\>,

which is a contradiction since the integrand is positive almost everywhere on Σ1\Sigma_{1}.

This result can easily be modified for axially symmetric surfaces. In that case, the Killing field used is simply E3×XE_{3}\times X and the helices are replaced by circles and the equation (1.9) still holds.

2. TreamillSled coordinates analysis

We will be considering immersions of the form

ϕ(s,t)=(x(s)cos(ωt)+y(s)sin(ωt),x(s)sin(ωt)+y(s)cos(ωt),t)\phi(s,t)=(x(s)\cos(\omega t)+y(s)\sin(\omega t),-x(s)\sin(\omega t)+y(s)\cos(\omega t),t)

with the curve α(s)=(x(s),y(s))\alpha(s)=(x(s),y(s)) parametrized by arc-length. We will refer to the curve α\alpha as the profile curve of the surface since the surface given as the image of ϕ\phi is the orbit of α\alpha under the helicoidal motion (1.6). For θ(s)\theta(s) defined by

x(s)=cos(θ(s))andy(s)=sin(θ(s))x^{\prime}(s)=\cos(\theta(s))\quad\hbox{and}\quad y^{\prime}(s)=\sin(\theta(s))

We define the TreadmillSled coordinates ξ1(s)\xi_{1}(s) and ξ2(s)\xi_{2}(s) by

(2.1) ξ1(s)=x(s)cos(θ(s))+y(s)sin(θ(s))andξ2(s)=x(s)sin(θ(s))y(s)cos(θ(s)).\displaystyle\xi_{1}(s)=x(s)\cos(\theta(s))+y(s)\sin(\theta(s))\quad\hbox{and}\quad\xi_{2}(s)=x(s)\sin(\theta(s))-y(s)\cos(\theta(s))\>.

The Gauss map of the immersion ϕ\phi can be computed as

ν=11+w2ξ12(sin(θωt),cos(θωt),ωξ1),\nu=\frac{1}{\sqrt{1+w^{2}\xi_{1}^{2}}}\left(\sin(\theta-\omega t),-\cos(\theta-\omega t),-\omega\xi_{1}\right)\>,

and so by a direct calculation, we obtain Q^=ξ2/1+ω2ξ12{\hat{Q}}=\xi_{2}/\sqrt{1+\omega^{2}\xi_{1}^{2}}. Finally, using that R2=x2+y2=ξ12+ξ22R^{2}=x^{2}+y^{2}=\xi_{1}^{2}+\xi_{2}^{2}, we see from (1.9), that the immersion ϕ\phi represents a rotating helicoidal drop if and only if there holds

(2.2) G(ξ1,ξ2):=2ξ21+ω2ξ12+Λ0(ξ12+ξ22)a4(ξ12+ξ22)2constant=:C.G(\xi_{1},\xi_{2}):=\frac{2\xi_{2}}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}+\Lambda_{0}(\xi_{1}^{2}+\xi_{2}^{2})-\frac{a}{4}(\xi_{1}^{2}+\xi_{2}^{2})^{2}\equiv{\rm constant}=:C\>.

After direct computation shows that the equation 2H=Λ0aR222H=\Lambda_{0}-a\frac{R^{2}}{2} reduces to

(2.3) θ(s)=2w2ξ22Λ0(1+w2ξ12)3/2+a(ξ12+ξ22)(1+w2ξ12)3/21+w2(ξ12+ξ22)\displaystyle\theta^{\prime}(s)=\frac{2w^{2}\xi_{2}-2\Lambda_{0}\left(1+w^{2}\xi_{1}^{2}\right)^{3/2}+a(\xi_{1}^{2}+\xi_{2}^{2})\left(1+w^{2}\xi_{1}^{2}\right)^{3/2}}{1+w^{2}\left(\xi_{1}^{2}+\xi_{2}^{2}\right)}

From the definition of ξ1\xi_{1}, ξ2\xi_{2} and θ\theta we get that ξ1=1ξ2θ\xi_{1}^{\prime}=1-\xi_{2}\theta^{\prime} and ξ2=ξ1θ\xi_{2}^{\prime}=\xi_{1}\theta^{\prime}. Using Equation (2.3) we conclude that ξ1\xi_{1} and ξ2\xi_{2} must satisfy

(2.4) ξ1\displaystyle\xi_{1}^{\prime} =\displaystyle= f1(ξ1,ξ2)=(1+w2ξ12)(2+1+w2ξ12ξ2(2Λ0a(ξ12+ξ22)))2(1+w2(ξ12+ξ22))\displaystyle f_{1}(\xi_{1},\xi_{2})=\frac{\left(1+w^{2}\xi_{1}^{2}\right)\left(2+\sqrt{1+w^{2}\xi_{1}^{2}}\,\xi_{2}\left(2\Lambda_{0}-a(\xi_{1}^{2}+\xi_{2}^{2})\right)\right)}{2(1+w^{2}(\xi_{1}^{2}+\xi_{2}^{2}))}
(2.5) ξ2\displaystyle\xi_{2}^{\prime} =\displaystyle= f2(ξ1,ξ2)=ξ1(2w2ξ22Λ0(1+w2ξ12)32+a(ξ12+ξ22)(1+w2ξ12)32)2(1+w2(ξ12+ξ22))\displaystyle f_{2}(\xi_{1},\xi_{2})=\frac{\xi_{1}\,\left(2w^{2}\xi_{2}-2\Lambda_{0}\left(1+w^{2}\xi_{1}^{2}\right)^{\frac{3}{2}}+a(\xi_{1}^{2}+\xi_{2}^{2})\left(1+w^{2}\xi_{1}^{2}\right)^{\frac{3}{2}}\right)}{2(1+w^{2}(\xi_{1}^{2}+\xi_{2}^{2}))}

This system of ordinary differential equations for ξ1\xi_{1} and ξ2\xi_{2} provides a different proof of the fact that the G(ξ1(s),ξ2(s))G(\xi_{1}(s),\xi_{2}(s)) must be constant. Since we can check that

Gξ1=2+2w2(ξ12+ξ22)(1+w2ξ12)32f2andGξ2=2+2w2(ξ12+ξ22)(1+w2ξ12)32f1\frac{\partial G}{\partial\xi_{1}}=-\frac{2+2w^{2}(\xi_{1}^{2}+\xi_{2}^{2})}{\left(1+w^{2}\xi_{1}^{2}\right)^{\frac{3}{2}}}\,f_{2}\quad\hbox{and}\quad\frac{\partial G}{\partial\xi_{2}}=\frac{2+2w^{2}(\xi_{1}^{2}+\xi_{2}^{2})}{\left(1+w^{2}\xi_{1}^{2}\right)^{\frac{3}{2}}}\,f_{1}
Remark 2.1.

The level sets of GG are symmetric with respect to the ξ2\xi_{2}-axis, therefore in order to understand the level set of GG, it is enough to understand those points in the level set with ξ10\xi_{1}\geq 0.

In order to study the level sets of the function GG we change the variables ξ1\xi_{1} and ξ2\xi_{2} for the variables rr and ξ2\xi_{2} where

r=ξ12+ξ22.r=\xi_{1}^{2}+\xi_{2}^{2}\>.

Doing this change, we get that the equation G=CG=C reduces to

2ξ21+ω2rω2ξ22+Λ0ra4r2=C.\frac{2\xi_{2}}{\sqrt{1+\omega^{2}r-\omega^{2}\xi_{2}^{2}}}+\Lambda_{0}r-\frac{a}{4}\,r^{2}=C\>.

In fact, this equation is exactly the one appearing in (1.9).

Therefore,

ξ2=(4C+r(4Λ0+ar))1+rω264+(4C+r(4Λ0+ar))2ω2.\xi_{2}=\frac{(4C+r(-4\Lambda_{0}+ar))\sqrt{1+r\omega^{2}}}{\sqrt{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}}\>.

By Remark 2.1, it is enough to consider those points with ξ10\xi_{1}\geq 0. Since ξ1=rξ22\xi_{1}=\sqrt{r-\xi_{2}^{2}} we get that

ξ1=p(r,a,Λ,C)64+(4C+r(4Λ0+ar))2ω2,\xi_{1}=\frac{\sqrt{p(r,a,\Lambda,C)}}{\sqrt{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}},

where

(2.6) p(r,a,Λ,C)=16C2+64r+32CΛ0r16Λ02r28aCr2+8aΛ0r3a2r4.\displaystyle p(r,a,\Lambda,C)=-16C^{2}+64r+32C\Lambda_{0}r-16\Lambda_{0}^{2}r^{2}-8aCr^{2}+8a\Lambda_{0}r^{3}-a^{2}r^{4}\,\>.
Remark 2.2.

Since p(r,a,Λ,C)p(r,a,\Lambda,C) is a polynomial in rr of degree 4 with negative leading coefficient when a0a\neq 0 and pp is a polynomial of degree two when a=0a=0, we get that the values of rr for which p(r,a,Λ,C)p(r,a,\Lambda,C) is positive are bounded. Since r=ξ12+ξ22=x2+y2r=\xi_{1}^{2}+\xi_{2}^{2}=x^{2}+y^{2}, we conclude that the profile curve of any helicoidal rotating drop is bounded.

Definition 1.

Let r1r_{1} and r2r_{2} be two non negative values that satisfies p(r1)=p(r2)=0p(r_{1})=p(r_{2})=0 and p(r)>0p(r)>0 for all r(r1,r2)r\in(r_{1},r_{2}). We define ρ:[r1,r2]𝐑2\rho:[r_{1},r_{2}]\longrightarrow{\bf R}^{2} by

ρ(r)=(p(r,a,Λ,C)64+(4C+r(4Λ0+ar))2ω2,(4C+r(4Λ0+ar))1+rω264+(4C+r(4Λ0+ar))2ω2).\rho(r)=\big{(}\,\frac{\sqrt{p(r,a,\Lambda,C)}}{\sqrt{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}}\,,\,\frac{(4C+r(-4\Lambda_{0}+ar))\sqrt{1+r\omega^{2}}}{\sqrt{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}}\,\big{)}\,\>.
Remark 2.3.

As pointed out before, all the level sets of the function GG are bounded. We have that the map ρ\rho parametrizes half of the level set G=CG=C.

Definition 2.

In the case the level set G=CG=C is a regular closed curve or union of regular closed curves, we define a fundamental piece of the profile curve as a simple connected part of the profile curve such that the parametrized curve (ξ1,ξ2)(\xi_{1},\xi_{2}) given by equations (2.1) correspond to exactly one closed curve in the level set of G=CG=C

Remark 2.4.

From the definition of TreadmillSled given in [4], we obtain that the profile curve of the solutions of the helicoidal rotational drop equation are characterized by the property that their treadmillSled are the level sets of GG. In other words, using the notation of [4], we have that TS(α)=βTS(\alpha)=\beta where β\beta is a parametrization of a connected component of the level set of G=CG=C and α\alpha is the profile curve of the helicoidal rotating drop. We will see that, for a few exceptional examples, the profile curve is a bounded complete curve having a circle as a limit cycle. For the non exceptional examples we can define an initial and final point of the fundamental piece and we have that the whole profile curve is the union of rotated fundamental pieces. We also have that if R1=min{|m|:mTS(α)}R_{1}=\hbox{min}\{|m|:m\in TS(\alpha)\} and R2=max{|m|:mTS(α)}R_{2}=\hbox{max}\{|m|:m\in TS(\alpha)\} and Δθ~\Delta{\tilde{\theta}} is the variation of the angle between 0p1\overrightarrow{0p_{1}} and 0p2\overrightarrow{0p_{2}} where p1p_{1} and p2p_{2} are the initial and final points of a fundamental piece, then, the profile curve is properly immersed if Δθ~π\frac{\Delta\tilde{\theta}}{\pi} is a rational number, otherwise the profile curve is dense in the set {(x,y)𝐑2:R1|(x,y)|R2}\{(x,y)\in{\bf R}^{2}:R_{1}\leq|(x,y)|\leq R_{2}\}.

Refer to captionRefer to captionRefer to captionRefer to caption

Figure 2.1. The first picture shows a helicoidal rotational drop, the second picture shows its profile curve emphasizing the fundamental piece, the third picture show the level set G=CG=C and the last picture shows how the TreadmillSled of the profile curve produces the level set G=CG=C, in this particular example the TreadmillSled of the profile curve will go over the level set G=CG=C two times.

We compute the variation Δθ~\Delta{\tilde{\theta}} in terms of the parameter rr. We assume that α(s)\alpha(s) is the profile curve of a helicoidal rotational drop. Recall that we are assuming that that ss is the arc-length parameter for the curve α\alpha. If β(s)=(ξ1(s),ξ2(s))\beta(s)=(\xi_{1}(s),\xi_{2}(s)), then we have that for some function s=σ(r)s=\sigma(r), β(σ(r))=ρ(r)\beta(\sigma(r))=\rho(r) holds. By the chain rule we have that

(2.7) dsdr=dσdr=|ρ(s)||β(s)|=|ρ(r)|2f12(s)+f22(s)=1264+(4C+r(4Λ0+ar))2ω2p(r,a,C).\displaystyle\frac{ds}{dr}=\frac{d\sigma}{dr}=\frac{|\rho^{\prime}(s)|}{|\beta^{\prime}(s)|}=\sqrt{\frac{|\rho^{\prime}(r)|^{2}}{f_{1}^{2}(s)+f_{2}^{2}(s)}}=\frac{1}{2}\,\sqrt{\frac{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}{p(r,a,C)}}\>.

and we also have that if θ~\tilde{\theta} denotes the polar angle of the profile curve, this is, if θ~(s)\tilde{\theta}(s) satisfies the equation α(s)=(x(s),y(s))=R(s)(cosθ(s)~,sinθ(s)~)\alpha(s)=(x(s),y(s))=R(s)(\cos\tilde{\theta(s)},\sin\tilde{\theta(s)}), then θ~(s)=ξ2(s)r\tilde{\theta}^{\prime}(s)=\frac{\xi_{2}(s)}{r}

(2.8) dθ~dr=dθ~dsdsdr=12(4C+ar24Λ0r)1+rω2rp(r,a,C)\displaystyle\frac{d\tilde{\theta}}{dr}=\frac{d\tilde{\theta}}{ds}\,\frac{ds}{dr}=\frac{1}{2}\,\frac{(4C+ar^{2}-4\Lambda_{0}r)\sqrt{1+r\omega^{2}}}{r\sqrt{p(r,a,C)}}

Since the map ρ:[r1,r2]𝐑2\rho:[r_{1},r_{2}]\longrightarrow{\bf R}^{2} parametrizes half of the TreadmillSled of the fundamental piece of the profile curve, we obtain the following expression for function Δθ~\Delta\tilde{\theta} defined in Remark 2.4

(2.9) Δθ~=Δθ~(C,a,ω,r1,r2)=r1r2(4C+ar24Λ0r)1+rω2rp(r,a,C)𝑑r\displaystyle\Delta\tilde{\theta}=\Delta\tilde{\theta}(C,a,\omega,r_{1},r_{2})=\,\int_{r_{1}}^{r_{2}}\frac{(4C+ar^{2}-4\Lambda_{0}r)\sqrt{1+r\omega^{2}}}{r\sqrt{p(r,a,C)}}\,dr
Remark 2.5.

If we have a helicoidal rotational drop Σ\Sigma and we multiply every point by a positive fixed number λ\lambda, this is, if we consider the surface λΣ\lambda\Sigma, then this new surface satisfies the equation of the rotating drop for some other values of Λ0\Lambda_{0} and aa. We also have that if we change the orientation of the profile curve of a surface Σ\Sigma that satisfies the equation of the rotating drop with values Λ0\Lambda_{0}, aa and HH, then the reparametrized surface satisfies the equation with values Λ0-\Lambda_{0}, a-a and H-H. With these two observations in mind, we have that in order to consider all the helicoidal rotational drops, up to parametrizations, rigid motions and dilations, it is enough to consider two cases: Case I, Λ0=0\Lambda_{0}=0 and a=1a=-1 and Case II, Λ0=1\Lambda_{0}=1 and aa is any real number.

2.1. Case I: Λ0=0\Lambda_{0}=0 and a=1a=-1

In this case the polynomial p(r,a,Λ,C)p(r,a,\Lambda,C) reduces to

q=q(r,C)=16C2+64r+8Cr2r4q=q(r,C)=-16C^{2}+64r+8Cr^{2}-r^{4}

Recall that we are interested in finding two positive consecutive roots of the polynomial qq. Notice that when CC is a negative large number then the polynomial qq has no roots and when CC is a positive large number then the polynomial qq has more than one root. In every case q(0)=16C20q(0)=-16C^{2}\leq 0 and the limit when rr\to\infty of q(r)q(r) is negative infinity. The following lemma was proven in [7] and provides the number of possible roots of q(r,C)q(r,C) in terms of the values of CC.

Lemma 2.6.

For any C>C0=3223C>C_{0}=-\frac{3}{2^{\frac{2}{3}}}, the polynomial q(r,C)q(r,C) has exactly two positive real roots. When C=C0C=C_{0}, 43\sqrt[3]{4} is the only real root of q(r,C)q(r,C) and when C<C0C<C_{0}, q(r,C)q(r,C) has not real roots.

Now we will compute the limit of Δθ~\Delta\tilde{\theta} when CC goes to C0C_{0}. We will use the following lemma from [5]

Lemma 2.7.

Let f(c,r)f(c,r) and g(r,c)g(r,c) be smooth functions such that g(C0,r0)=gr(C0,r0)=0g(C_{0},r_{0})=\frac{\partial g}{\partial r}(C_{0},r_{0})=0 and 2gr2(C0,r0)=2A\frac{\partial^{2}g}{\partial r^{2}}(C_{0},r_{0})=-2A where A>0A>0. If {Cn}\{C_{n}\}, {un}\{u_{n}\} and {vn}\{v_{n}\} are sequences such that CnC_{n} converges to C0C_{0}, unu_{n} and vnv_{n} converges to r0r_{0} with un<r0<vnu_{n}<r_{0}<v_{n}, and g(un)=g(vn)=0g(u_{n})=g(v_{n})=0 and g(r)>0g(r)>0 for all r(un,vn)r\in(u_{n},v_{n}), then

unvnf(c,r)drg(c,r)f(C0,r0)πAasn.\int_{u_{n}}^{v_{n}}\frac{f(c,r)\,dr}{\sqrt{g(c,r)}}\longrightarrow\,f(C_{0},r_{0})\,\frac{\pi}{\sqrt{A}}\quad\hbox{as}\quad n\longrightarrow\infty\>.

Notice that helicoidal rotating drops are defined when CC takes value from C0=343C_{0}=-\frac{3}{\sqrt[3]{4}} to \infty. When C=C0C=C_{0} the only root of the polynomial pp is r0=43r_{0}=\sqrt[3]{4}. If we apply Lemma 2.7 with f(r,c)=(4Cr2)1+rω2rf(r,c)=\frac{(4C-r^{2})\sqrt{1+r\omega^{2}}}{r} and g(r,c)=q(r,C)g(r,c)=q(r,C) to the integral given in (2.9), we obtained that

(2.10) limCC0+Δθ~=B(ω)=2π31+43ω2\displaystyle\lim_{C\to C_{0}^{+}}\Delta\tilde{\theta}=B(\omega)=-\frac{2\pi}{\sqrt{3}}\sqrt{1+\sqrt[3]{4}\,\omega^{2}}
Remark 2.8.

Recall that whenever Δθ~=n2πm\Delta\tilde{\theta}=\frac{n2\pi}{m} holds for some pair of integers mm and nn, then the entire profile curve is properly immersed and it is invariant under the group ZmZ_{m}.

Remark 2.9.

Up to dilations and rigid motions, the moduli space for all helicoidal rotating drops with Λ0=0\Lambda_{0}=0 is the region in the plane

{(C,ω):CC0=343,ω>0}\{(C,\omega):C\geq C_{0}=-\frac{3}{\sqrt[3]{4}},\quad\omega>0\}

Moreover, for any ω>0\omega>0, the surface associated with the point (C,ω)=(C0,ω)(C,\omega)=(C_{0},\omega) is a round cylinder of radius 23\sqrt[3]{2}, because it can be easily checked that when C=C0C=C_{0}, then, for any ω\omega, the level set G=CG=C reduces to the point {(0,2)}\{(0,-\sqrt{2})\}

Refer to caption

Figure 2.2. Moduli space of the helicoidal drops with a=1a=-1 and Λ=0\Lambda=0

2.2. Case II: Λ0=1\Lambda_{0}=1

First note that the case a=0a=0 corresponds to helicoidal surface with constant mean curvature. These surfaces were studied using similar techniques in [4] and for this reason we will assume here that a0a\neq 0. In this case the polynomial p(r,a,Λ,C)p(r,a,\Lambda,C) reduces to

(2.11) q=q(r,C,a)=16C2+64r+32Cr16r28aCr2+8ar3a2r4.\displaystyle q=q(r,C,a)=-16C^{2}+64r+32C\,r-16\,r^{2}-8aC\,r^{2}+8a\,r^{3}-a^{2}\,r^{4}\>.

Recall that we are interested in finding two positive consecutive roots of the polynomial qq. The roots of the polynomial qq given in (2.11) were analyzed in [7]. In order to describe the roots of qq we need to define the following functions.

Definition 3.

Let h(R)=2(R1)R3h(R)=\frac{2(R-1)}{R^{3}}, and define R1:(,0)(0,)𝐑R_{1}:(-\infty,0)\cup(0,\infty)\to{\bf R}, R2:(0,8/27)𝐑R_{2}:(0,8/27)\to{\bf R} and R3:(0,8/27)𝐑R_{3}:(0,8/27)\to{\bf R} by

R1(a)=Rsuch thath(R)=awithR<1,R2(a)=Rsuch thath(R)=awith1<R<3/2,R3(a)=Rsuch thath(R)=awith3/2<R<.\begin{array}[]{cccc}R_{1}(a)=R\quad\hbox{such that}&h(R)=a&\quad\hbox{with}&R<1\>,\\ R_{2}(a)=R\quad\hbox{such that}&h(R)=a&\quad\hbox{with}&1<R<3/2\>,\\ R_{3}(a)=R\quad\hbox{such that}&h(R)=a&\quad\hbox{with}&3/2<R<\infty\>.\end{array}

We also define the functions r1:(,0)(0,)𝐑r_{1}:(-\infty,0)\cup(0,\infty)\to{\bf R}, r2:(0,8/27)𝐑r_{2}:(0,8/27)\to{\bf R} and r3:(0,8/27)𝐑r_{3}:(0,8/27)\to{\bf R} by

r1(a)=R12(a),r2(a)=R22(a),r3(a)=R32(a).r_{1}(a)=R_{1}^{2}(a),\quad r_{2}(a)=R_{2}^{2}(a),\quad r_{3}(a)=R_{3}^{2}(a)\>.

The following lemma was proven in [7] and provides the number of roots of qq depending on the values aa and CC.

Lemma 2.10.

Let r1r_{1}, r2r_{2} and r3r_{3} be as in Definition 3 and for i=1,2,3i=1,2,3 define

Ci=168ri+6ari2a2ri34(2+ari)C_{i}=\frac{16-8r_{i}+6a{r_{i}}^{2}-a^{2}{r_{i}}^{3}}{4(-2+ar_{i})}

and

q=q(r,a,C)=16C2+64r+32Cr16r28aCr2+8ar3a2r4.q=q(r,a,C)=-16C^{2}+64r+32C\,r-16\,r^{2}-8aC\,r^{2}+8a\,r^{3}-a^{2}\,r^{4}\>.

Recall that the domain of CiC_{i} is the same domain of rir_{i}. That is, the domain of C1(a)C_{1}(a) is {a0}\{a\neq 0\} and the domain of C2(a)C_{2}(a) and C3(a)C_{3}(a) is the interval (0,827](0,\frac{8}{27}]. For any a0a\neq 0 and any CC, the polynomial qq has non negative real roots whose multiplicities are given in the following table.

range of aa range of CC number of distinct real roots of qq
a<0C1(a)<0a<0\Rightarrow C_{1}(a)<0 C<C1(a)C<C_{1}(a) 0
C=C1(a)C=C_{1}(a) 11
C>C1(a)C>C_{1}(a) 2
a(0,827)C2(a)<0a\in(0,\frac{8}{27})\Rightarrow C_{2}(a)<0, C1(a)>0C_{1}(a)>0,
C2(a)<C3(a)<C1(a)C_{2}(a)<C_{3}(a)<C_{1}(a) C>C1(a)C>C_{1}(a) 0
C=C1(a)C=C_{1}(a) 11
C3(a)<C<C1(a)C_{3}(a)<C<C_{1}(a) 22
C=C3(a)C=C_{3}(a) 33(****)
C2(a)<C<C3(a)C_{2}(a)<C<C_{3}(a) 44
C=C2(a)C=C_{2}(a) 33(***)
C<C2(a)C<C_{2}(a) 33
a=827C2(a)=C3(a)=9/8a=\frac{8}{27}\Rightarrow C_{2}(a)=C_{3}(a)=-9/8,
C1(827)=9C_{1}(\frac{8}{27})=9 C<9/8C<-9/8 22
C=9/8C=-9/8 22(*)
9/8<C<9-9/8<C<9 22
C=9C=9 11(**)
9>C9>C 0
a>8/27C1(a)>0a>8/27\Rightarrow C_{1}(a)>0 C>C1(a)C>C_{1}(a) 0
C=C1(a)C=C_{1}(a) 11
C<C1(a)C<C_{1}(a) 22

(*) In this case the roots are 9/49/4 with multiplicity 33 and 81/481/4 with multiplicity one.
(**) In this case the only real root is 99 with multiplicity 22.
(***) In this case the first root has multiplicity 2.
(****) In this case the second root has multiplicity 2.

Now that we have discussed the roots of the polynomial qq we can describe the moduli space of all helicoidal rotating drops with Λ=1\Lambda=1.

Theorem 2.11.

Let Λ0=1\Lambda_{0}=1 and let Δθ~\Delta\tilde{\theta} be the function defined in (2.9). Let

Ω1={(a,C,ω):C>C1(a),a<0,w>0}Ω2={(a,C,ω):C2(a)<C<C3(a), 0<a<827,ω>0}\Omega_{1}=\{(a,C,\omega):C>C_{1}(a),\,a<0,\,w>0\}\quad\Omega_{2}=\{(a,C,\omega):C_{2}(a)<C<C_{3}(a),\,0<a<\frac{8}{27}\,,\,\omega>0\}\,
Ω3={(a,C,ω):C<C1(a),a>0,ω>0}Ω=Ω1Ω3Ω2\Omega_{3}=\{(a,C,\omega):C<C_{1}(a),\,a>0,\,\omega>0\}\quad\Omega=\Omega_{1}\cup\Omega_{3}\setminus\Omega_{2}
β1={(a,C,ω):C=C1(a),a0,ω>0}\beta_{1}=\{(a,C,\omega)\,:\,C=C_{1}(a),\,a\neq 0,\,\omega>0\}
β2={(a,C,ω):C=C2(a), 0<a<827,ω>0}\beta_{2}=\{(a,C,\omega)\,:\,C=C_{2}(a),\,0<a<\frac{8}{27},\,\omega>0\}
β3={(a,C,ω):C=C3(a), 0<a<827,ω>0}\beta_{3}=\{(a,C,\omega)\,:\,C=C_{3}(a),\,0<a<\frac{8}{27},\,\omega>0\}

Under the convention that a point (a,C,ω)(a,C,\omega) represents a helicoidal rotating drop if the TreadmillSled of its profile curve is contained in the level set G=CG=C, we have:

i.) Every point (a,C,ω)(a,C,\omega) in the interior of Ω\Omega represents a helicoidal rotating drop with its fundamental piece having finite length. The TreadmillSled of the profile curve of these surfaces are parametrized by ρ\rho defined for values of rr between the only two roots of the polynomial q(r,a,C)q(r,a,C).

ii.) Every point in Ω2\Omega_{2} represents two helicoidal rotating drops, both having fundamental pieces of finite length. The TreadmillSleds of the profile curves of these surfaces are parametrized by ρ\rho defined for those values of rr that lie between the first and second root of the polynomial q(r,a,C)q(r,a,C) and the third and fourth root of the polynomial q(r,a,C)q(r,a,C) respectively.

iii.) Every point (a,C,ω)(a,C,\omega) in the set β1\beta_{1} represents a circular helicoidal rotating drop. This cylinder is the same for all values of ω\omega.

iv.) Every point (a,C,ω)(a,C,\omega) in the set β2\beta_{2} represents two helicoidal rotating drops: a circular cylinder and a non circular cylinder with bounded length of its fundamental piece. The circular cylinder is the same for all values of ww.

v.) Every point in the set β3\beta_{3} represents three helicoidal rotating drops. One is a circular cylinder, which is the same for all values of ww. The second one has a TreadmillSled parametrized by ρ\rho defined for those values of rr that lie between the first and second root of the polynomial q(r,a,C)q(r,a,C). Recall that the second root has multiplicity 22. The third surface has a TreadmillSled parametrized by ρ\rho defined for those values of rr that lie between the second and third root of the polynomial q(r,a,C)q(r,a,C). The second and third surfaces are not properly immersed and their profile curves have a circle as a limit cycle and they have infinite winding number with respect to a point interior to this circle. Solutions similar to these second or third types will be called helicoidal drops of exceptional type.

vi.) Points of the form (a,C,ω)=(827,98,ω)(a,C,\omega)=(\frac{8}{27},-\frac{9}{8},\omega) represent two helicoidal rotating drops: a circular cylinder, which is the same for all values of ω\omega, and one helicoidal drop of exceptional type.

vii.) Up to a rigid motion, every helicoidal drop falls into one of the cases above.

viii.) Every helicoidal drop that is not exceptional is either properly immersed (when Δθ~(a,C,ω)π\frac{\Delta\tilde{\theta}(a,C,\omega)}{\pi} is a rational number) or it is dense in the region bound by two round cylinders (when Δθ~(a,C,ω)π\frac{\Delta\tilde{\theta}(a,C,\omega)}{\pi} is an irrational number).

Refer to caption

Figure 2.3. The picture above shows the graph of the functions C1C_{1}, C2C_{2} and C3C_{3}, these graphs are used in Proposition 2.11 to describe the moduli space of all helicoidal rotational drops with Λ0=1\Lambda_{0}=1
Proof.

We already know that the TreadmillSled of the profile curve of any helicoidal rotating drop satisfies the equation

G(ξ1,ξ2)=2ξ21+ω2ξ12+Λ0(ξ12+ξ22)a4(ξ12+ξ22)2=CG(\xi_{1},\xi_{2})=\frac{2\xi_{2}}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}+\Lambda_{0}(\xi_{1}^{2}+\xi_{2}^{2})-\frac{a}{4}(\xi_{1}^{2}+\xi_{2}^{2})^{2}=C

We also know that, up to rigid motions, the TreadmillSled of a curve determines the curve, see [4]. Since any level set of GG can be parametrized using the map ρ\rho given in Definition 1, and every parametrization of a level set of GG is defined for values of rr where the polynomial qq is positive, it then follows from Lemma 2.10 that every helicoidal rotating drop can be represented as one of the cases i, ii, iii, iv, v and vi. It is worth recalling, see Remark 2.3, that the parametrization ρ\rho only covers half of the level set of the map GG. Each one of these level sets is symmetric with respect to the ξ2\xi_{2} axis, and the parametrization ρ\rho covers the half on the right.

Notice that when the profile curve is a circle, the level set G=CG=C reduces to a point. When the profile curve is a circle we will take the parametrization ρ\rho to be defined just in a point, a root with multiplicity 2 of the polynomial qq.

When case (i) occurs, qq has only two simple roots x1x_{1} and x2x_{2} with x1<x2x_{1}<x_{2}. We can check that the derivative of qq at x1x_{1} is positive while the derivative of qq at x2x_{2} is negative, so the length of the fundamental piece, according to Equation (2.7), reduces to x1x264+(4C+r(4Λ0+ar))2ω2q(r,a,C)𝑑r\int_{x_{1}}^{x_{2}}\,\sqrt{\frac{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}{q(r,a,C)}}\,dr which converges. Therefore the length of the fundamental piece is finite.

For values of CC, ω\omega and aa that fall into case (ii), the polynomial q has 4 roots x1<x2<x3<x4x_{1}<x_{2}<x_{3}<x_{4} and it is positive from x1x_{1} to x2x_{2} and from x3x_{3} to x4x_{4}. Also, the level set of GG has two connected components. Half of each connected components of G=CG=C can be parametrized using the map ρ\rho. One half of the connected component of G=CG=C uses the domain (x1,x2)(x_{1},x_{2}) for ρ\rho and the half of the other connected component of G=CG=C uses the domain (x3,x4)(x_{3},x_{4}) for ρ\rho. The proof that the length of the fundamental piece of each surface is finite follows as in the proof in case (i).

For values of (a,C,ω)(a,C,\omega) that satisfies the case (iii), the polynomial qq has only one root x1=r1x_{1}=r_{1} with multiplicity two. We take R=x1R=\sqrt{x_{1}}. A direct calculation shows that if a>0a>0, then R1(a)=RR_{1}(a)=-R and if we consider the profile curve α(s)=(Rsin(sR),Rcos(sR))\alpha(s)=(R\sin(\frac{s}{R}),-R\cos(\frac{s}{R})), then ξ1=0\xi_{1}=0, ξ2=R\xi_{2}=R and G(ξ1,ξ2)=2R+Λ0R2a4R4G(\xi_{1},\xi_{2})=2R+\Lambda_{0}R^{2}-\frac{a}{4}R^{4}. Using the definition of C1C_{1} and the fact that R1(a)=RR_{1}(a)=-R, we can check that the expression G(ξ1,ξ2)G(\xi_{1},\xi_{2}) reduces to C=C1(a)C=C_{1}(a), which was our goal in order to show that the point (a,C,ω)(a,C,\omega) represents a round cylinder. Similarly, a direct verification shows that if a<0a<0, then R1(a)=RR_{1}(a)=R and if we consider the profile curve α(s)=(Rsin(sR),Rcos(sR))\alpha(s)=(R\sin(\frac{s}{R}),R\cos(\frac{s}{R})), then ξ1=0\xi_{1}=0, ξ2=R\xi_{2}=-R and G(ξ1,ξ2)=2R+Λ0R2a4R4G(\xi_{1},\xi_{2})=-2R+\Lambda_{0}R^{2}-\frac{a}{4}R^{4}. Using the definition of C1C_{1} and the fact that R1(a)=RR_{1}(a)=R, we can check that the expression G(ξ1,ξ2)G(\xi_{1},\xi_{2}) reduces to C=C1(a)C=C_{1}(a). Since r1r_{1} is independent of ww we have that these cylinders are independent of the value of ww. This finish the proof of part (iii).

For values of (a,C,ω)(a,C,\omega) that fall into case (iv), the polynomial qq has three roots x1<x2<x3x_{1}<x_{2}<x_{3}, where x1=r1x_{1}=r_{1} has multiplicity two and x2x_{2} and x3x_{3} are simple. The polynomial qq is positive for values of rr between x2x_{2} and x3x_{3}. In this case the level set G=CG=C is the union of the point (0,x1)(0,-\sqrt{x_{1}}) and a closed curve. If we consider the cylinder or radius x1\sqrt{x_{1}} oriented by the inward pointing normal, then, a direct computation shows that its mean curvature is 2H=1a2r12H=1-\frac{a}{2}r_{1}. Therefore this circular cylinder is a helicoidal rotating drop for the given parameters. Note that this cylinder is independent of ww. The TreamillSled of the profile curve of the other rotating drop is the level closed curve component of G=CG=C; half of this part can be parametrized by the map ρ\rho with domain those values of rr between x2x_{2} and x3x_{3}.

For values of (a,C,ω)(a,C,\omega) in case (v), the polynomial qq has only three roots x1<x2<x3x_{1}<x_{2}<x_{3}, where x2=r2x_{2}=r_{2} has multiplicity two and x1x_{1} and x3x_{3} are simple. The polynomial qq is positive for values of rr between x1x_{1} and x2x_{2} and for values of rr between x2x_{2} and x3x_{3}. In this case the level set G=CG=C is connected but it self-intersects at the point (0,x2)(0,-\sqrt{x_{2}}). Any part of a curve that crosses the ξ2\xi_{2}-axis non-horizontally cannot be the TreadmillSled of a regular curve (see Proposition 2.11 in [6]). Therefore the correct way to view the level set G=CG=C in this case is not as a connected closed curve that self intersects but as the union of two curves and a point. Figure 2.4 shows one of these level sets.

Refer to caption

Figure 2.4. For helicoidal drops in case (v) of Proposition 2.11, the level set of GG should be regarded as the union of two curves and a point. Each curve is the TreadmillSled of an exceptional helicoidal rotating drop and the point is the treadmillSled of a circular cylinder .

.

One can check that the circular cylinder with radius r2\sqrt{r_{2}}, oriented by the inward pointing normal satisfies the 2H=1a2r22H=1-\frac{a}{2}r_{2}, therefore this circular cylinder is a helicoidal rotating drop for the given parameters. The TreadmillSled associated with the profile curve of this round cylinder reduces to the point (0,x2)(0,-\sqrt{x_{2}}). The set G=CG=C minus the point (0,x2)(0,-\sqrt{x_{2}}) has two connected components. One of these connected components can be parametrized using the map ρ\rho with values of rr between x1x_{1} and x2x_{2} and the other using the map ρ\rho with values of rr between x2x_{2} and x3x_{3}. Each of these connected components is the TreamillSled of the fundamental curve for a rotating helicoidal drop whose length is unbounded. Specifically, their lengths are given respectively by the divergent integrals x1x264+(4C+r(4Λ0+ar))2ω2q(r,a,C)𝑑r\int_{x_{1}}^{x_{2}}\sqrt{\frac{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}{q(r,a,C)}}\,dr and x2x364+(4C+r(4Λ0+ar))2ω2q(r,a,C)𝑑r\int_{x_{2}}^{x_{3}}\sqrt{\frac{64+(4C+r(-4\Lambda_{0}+ar))^{2}\,\omega^{2}}{q(r,a,C)}}\,dr. Moreover, using the definition of TreadmillSled, we notice that the function giving the distance to the origin of the profile curve, |(x(s),y(s)||(x(s),y(s)|, agrees with the function distance to the origin of the level set G=CG=C given by |(ξ1(s),ξ2(s))|=|ρ(σ1(s))||(\xi_{1}(s),\xi_{2}(s))|=|\rho(\sigma^{-1}(s))|. Therefore as rr aproaches r2r_{2}, s=σ(r)s=\sigma(r) goes to -\infty and the function |(x(s),y(x)||(x(s),y(x)| approaches r2\sqrt{r_{2}}. Since polar angle of the profile curve can be calculated by integrating the expression in (2.8), we conclude that θ~(r)\tilde{\theta}(r) also goes to -\infty as rr approaches r2r_{2}. We conclude that the profile curve has a circle of radius r2\sqrt{r_{2}} as a limit cycle and it has infinite winding number with respect to a point interior to this circle.

For values of (a,C,w)(a,C,w) that satisfies the case (vi) the polynomial qq has only two roots x1<x2x_{1}<x_{2} where x1=94x_{1}=\frac{9}{4} has multiplicity three and x2=814x_{2}=\frac{81}{4} is simple. The polynomial qq is positive for values of rr between x1x_{1} and x2x_{2}. In this case the level set G=CG=C is connected but it has a singularity at the (0,32)(0,-\frac{3}{2}). We can check that the circular cylinder with radius 32\frac{3}{2} oriented by the inward pointing normal satisfies the 2H=1a2r12H=1-\frac{a}{2}r_{1}, therefore this circular cylinder is a helicoidal rotating drop. The TreadmillSled associated of the profile curve of this round cylinder reduces to the point (0,32)(0,-\frac{3}{2}). The set G=CG=C minus the point (0,32)(0,-\frac{3}{2}) is connected and half of it can be parametrized using the map ρ\rho with values of rr between 94\frac{9}{4} and 814\frac{81}{4}. This part of the set G=CG=C is the TreamillSled of the fundamental curve of a rotating helicoidal drop whose length is not bounded.

Since we know that the profile curve of every rotating helicoidal drop satisfies the integral equation G=CG=C and cases (i)-(vi) cover all the possible level sets for the level sets of GG then every rotating helicoidal drop fall into one of the first 6 cases of this proposition. This proves (vii).

In order to prove (viii) we notice that when a helicoidal rotating drop is not exceptional, it has a fundamental piece with finite length whose TreadmillSled is a closed regular curve (a connected component of the set G=CG=C). By the properties of the TreadmillSled operator (in particular the one that states that the TreadmillSled inverse is unique up to rotations about the origin), we have the the whole profile curve is a union of rotations of the fundamental piece. The angle of rotation is given by Δθ~=Δθ~(C,a,ω,x1,x2)\Delta\tilde{\theta}=\Delta\tilde{\theta}(C,a,\omega,x_{1},x_{2}). We therefore have that the profile curve is invariant under the group of rotations

(2.12) 𝔾={(y1,y2)(cos(nΔθ~)y1+sin(nΔθ~)y2,sin(nΔθ~)y1+cos(nΔθ)y2)|n}.\displaystyle\mathbb{G}=\{(y_{1},y_{2})\rightarrow(\cos(n\Delta\tilde{\theta})y_{1}+\sin(n\Delta\tilde{\theta})y_{2},-\sin(n\Delta\tilde{\theta})y_{1}+\cos(n\Delta\theta)y_{2})\,|\,n\in\mathbb{Z}\}\>.

It is clear that if Δθ~π\frac{\Delta\tilde{\theta}}{\pi} is a rational number then the group 𝔾\mathbb{G} is finite and the helicoidal surface is properly immerse. Moreover, if Δθ~π\frac{\Delta\tilde{\theta}}{\pi} is not a rational number, then the group 𝔾\mathbb{G} is not finite and the helicoidal drop is dense in the region bounded by the two cylinders of radius r1\sqrt{r_{1}} and r2\sqrt{r_{2}}. A more detailed explanation of this last statement can be found in [4].

2.3. Embedded and properly embedded examples

In this subsection we will find some embedded examples and we will show their profile curves. As pointed out before, when the helicoidal drop is not exceptional, its profile curve is a union rotations of fundamental pieces that ends up being invariant under the group 𝔾\mathbb{G} define in (2.12). It is not difficult to see that a necessary condition for the helicoidal drop to be embedded is that Δθ~=2πm\Delta\tilde{\theta}=\frac{2\pi}{m} for some integer mm. We will show that this condition is not sufficient. In order to catch the potentially embedded examples we need to understand the function Δθ~(C,a,ω,x1,x2)\Delta\tilde{\theta}(C,a,\omega,x_{1},x_{2}). As a very elementary technique to solve the equation Δθ~=2πm\Delta\tilde{\theta}=\frac{2\pi}{m} we will use the intermediate value theorem. We know that for any aa, there is a first (or last) value of CC, C0(a,x1,x2)C_{0}(a,x_{1},x_{2}), for which the function Δθ~\Delta\tilde{\theta} is defined. We will compute the limit of Δθ~\Delta\tilde{\theta} when CC goes to C0C_{0} using Lemma 2.7. The graphs shown in this paper were generated using the software Mathematica 8 .

2.4. Embedded examples with Λ0=0\Lambda_{0}=0 and a=1a=-1

From Lemma 2.6 we know that the polynomial pp has two positive root if and only if C>C0=343C>C_{0}=-\frac{3}{\sqrt[3]{4}}. A direct application of Lemma 2.7 shows:

Proposition 2.12.

If Λ0=0\Lambda_{0}=0, a=1a=-1 and for any C>C0C>C_{0}, x1x_{1} and x2x_{2} denote the two roots of the polynomial q(r,C)=16C2+64r+8Cr2r4q(r,C)=-16C^{2}+64r+8Cr^{2}-r^{4}, then

limCC0+Δθ(C,ω,x1,x2)=x1x2(4Cr2)1+rω2rq(r,C)𝑑r=B(ω)=2π1+43ω23\lim_{C\to C_{0}^{+}}\Delta\theta(C,\omega,x_{1},x_{2})=\,\int_{x_{1}}^{x_{2}}\frac{(4C-r^{2})\sqrt{1+r\omega^{2}}}{r\sqrt{q(r,C)}}\,dr=B(\omega)=-\frac{2\pi\sqrt{1+\sqrt[3]{4}\,\omega^{2}}}{\sqrt{3}}

holds.

Refer to captionRefer to captionRefer to captionRefer to caption

Figure 2.5. The last graph shows how the beginning of the graphs on the left change when ω\omega changes. The highlighted points in the last graph (ω=0\omega=0, 11 and 1.51.5) correspond to the highlighted points in the three graphs to the left. For ω=0\omega=0 there is no solution of the equation Δθ~=2πm\Delta{\tilde{\theta}}=-\frac{2\pi}{m} with negative values of CC. We see that, for some values of ω\omega, the equation Δθ~=2π\Delta{\tilde{\theta}}=-2\pi has a solution with CC negative, which is responsible for the existence of embedded examples with Λ0=0\Lambda_{0}=0 and a=1a=-1.

Refer to captionRefer to captionRefer to caption

Figure 2.6. The profile curves of some embedded helicoidal rotating drops when Λ0=0\Lambda_{0}=0 and a=1a=-1. When CC is close to the critical value C0,C_{0}, the embedded examples is close to a round cylinder. As CC increases, the shape develops a self-intersection.

Refer to captionRefer to captionRefer to caption

Figure 2.7. Surfaces associated with the profile curve on Figure 2.6 .

Using the Intermediate Value Theorem, we can numerically solve the equation Δθ~=2π\Delta\tilde{\theta}=-2\pi for values of ww and CC. The images in Figures 2.6 and 2.7 show some of the resulting profile curves and the corresponding surfaces. They also show the values of CC and ww that solve the equation Δθ~=2π\Delta\tilde{\theta}=-2\pi.

2.5. Embedded examples with Λ0=1\Lambda_{0}=1 and a0a\neq 0

We now show some embedded examples in this case. Again the Intermediate Value Theorem is used to numerically solve the equation Δθ~=2πm\Delta\tilde{\theta}=\frac{2\pi}{m}. A direct application of Lemma 2.7 shows:

Proposition 2.13.

Let r1r_{1} r2r_{2} and r3r_{3} be the expressions given in Definition 3 and let C1C_{1}, C2C_{2} and C3C_{3} be the expression defined in Lemma 2.10. For i=1,2i=1,2 let us defined the following two bounds.

bi(a,w)=π(4Ci+ari24ri)1+riw2ri16+8a(C3ri)+6a2ri2.b_{i}(a,w)=\frac{\pi(4C_{i}+ar_{i}^{2}-4r_{i})\sqrt{1+r_{i}w^{2}}}{r_{i}\sqrt{16+8a(C-3r_{i})+6a^{2}r_{i}^{2}}}\>.

a.) If a<0a<0 then limCC1(a)+Δθ~(C,a,x1,x2)=b1(a)\lim_{C\to C_{1}(a)^{+}}\Delta\tilde{\theta}(C,a,x_{1},x_{2})=b_{1}(a). Here x1x_{1} and x2x_{2} are the first two roots of the polynomial q(r,C,a)q(r,C,a)

b.) If a>0a>0 then limCC1(a)Δθ~(C,a,x1,x2)=b1(a)\lim_{C\to C_{1}(a)^{-}}\Delta\tilde{\theta}(C,a,x_{1},x_{2})=b_{1}(a). Here x1x_{1} and x2x_{2} are the first two roots of the polynomial q(r,C,a)q(r,C,a)

c.) If a>0a>0 then limCC2(a)+Δθ~(C,a,x1,x2)=b2(a)\lim_{C\to C_{2}(a)^{+}}\Delta\tilde{\theta}(C,a,x_{1},x_{2})=b_{2}(a). Here x1x_{1} and x2x_{2} are the first two roots of the polynomial q(r,C,a)q(r,C,a).

Proof.

Since Δθ~=x1x2(4C+ar24r)1+rω2rq𝑑r\Delta\tilde{\theta}=\int_{x_{1}}^{x_{2}}\frac{(4C+ar^{2}-4r)\sqrt{1+r\omega^{2}}}{r\sqrt{q}}\,dr, in every case, when CC approaches the limit value, the two roots approach rir_{i} (i=1 or 2), which is a root of qq with multiplicity 22. Therefore Lemma 2.7 applies and the proposition follows. Notice that the value AA in Lemma 2.7 is given by

A=12q′′(ri)=16+8a(C3ri)+6a2ri2.A=-\frac{1}{2}q^{\prime\prime}(r_{i})=16+8a(C-3r_{i})+6a^{2}r_{i}^{2}\>.

Remark 2.14.

When 0<a<8270<a<\frac{8}{27} the domain of the function Δθ~\Delta\tilde{\theta} starts at C=C2(a)C=C_{2}(a) and the limit of the function Δθ~(C)\Delta\tilde{\theta}(C) when CC2(a)+C\to C_{2}(a)^{+} is b2(a,w)b_{2}(a,w). Moreover, this function has a vertical asymptote at C=C3(a)C=C_{3}(a), and then it has a jump discontinuity at C=0C=0. Finally, this function is define for all values of CC smaller than C1(a)C_{1}(a) and the limit of Δθ~(C)\Delta\tilde{\theta}(C) when CC1(a)C\to C_{1}(a)^{-} is b1(a,w)b_{1}(a,w)

Refer to caption

Figure 2.8. The graph of the function Δθ~\Delta\tilde{\theta} when a=0.2a=0.2 and ω=0.15\omega=0.15. In this case C21.065C_{2}\approx-1.065, b27.66b_{2}\approx-7.66, C30.698C_{3}\approx-0.698, C111.76C_{1}\approx 11.76 and b12.23b_{1}\approx 2.23. The points (C2,b2)(C_{2},b_{2}) and (C1,b1)(C_{1},b_{1}) have been highlighted.

Taking a look at Figure 2.8 we notice that, when ω=0.15\omega=0.15 and a=0.2a=0.2, we have that for any integer m>2m>2, the equation Δθ~=2πm\Delta\tilde{\theta}=\frac{2\pi}{m} has a solution. We have numerically solved this equation for m=4m=4 and m=8m=8. Figures 2.9 and 2.10 provides a picture of the profile curves of the properly immerse examples.

Refer to caption  Refer to caption  Refer to caption

Figure 2.9. The embedded helicoidal rotational drop obtained by solving the equation Δθ~=π2\Delta\tilde{\theta}=\frac{\pi}{2} when a=0.2a=0.2 and ω=0.15\omega=0.15.

Refer to caption  Refer to caption

Figure 2.10. The properly immerse helicoidal rotational drop obtained by solving the equation Δθ~=π4\Delta\tilde{\theta}=\frac{\pi}{4} when a=0.2a=0.2 and ω=0.15\omega=0.15.

We have that if we decrease the value of aa while keeping the value of ω\omega constant, we can again solve the equation Δθ~=π4\Delta\tilde{\theta}=\frac{\pi}{4} but this time the helicoidal rotational drop is embedded. See Figure 2.11.

Refer to caption  Refer to caption

Figure 2.11. The embedded helicoidal rotational drop obtained by solving the equation Δθ~=π4\Delta\tilde{\theta}=\frac{\pi}{4} when a=0.05a=0.05 and ω=0.15\omega=0.15.

Finally we would like to show that if we increase ω\omega then it is possible to solve the equation Δθ~=2π\Delta\tilde{\theta}=2\pi, (see Figure 2.12) .

Refer to caption  Refer to caption  Refer to caption

Figure 2.12. The embedded helicoidal rotational drop obtained by solving the equation Δθ~=2π\Delta\tilde{\theta}=2\pi when a=0.2a=0.2 and ω=2\omega=2.

3. Second variation

For any sufficiently smooth surface, we define an invariant =2H+(a/2)R2\ell=2H+(a/2)R^{2}. The first variation formula (1.4) restricted to compactly supported variations can then be expressed

δa,Λ0=Σ(Λ0)ψ𝑑Σ,\delta{\mathcal{E}}_{a,\Lambda_{0}}=-\int_{\Sigma}(\ell-\Lambda_{0})\psi\>d\Sigma\>,

where ψ:=δXν\psi:=\delta X\cdot\nu. We assume that the surface is in equilibrium so that Λ00\ell-\Lambda_{0}\equiv 0 holds. The second variation is thus

δ2a,Λ0=Σψ(δ)𝑑Σ.\delta^{2}{\mathcal{E}}_{a,\Lambda_{0}}=-\int_{\Sigma}\psi(\delta\ell)\>d\Sigma\>.

A well known formula for the pointwise variation of the mean curvature is

(3.1) 2δH=L^[ψ]+2HδX,2\delta H={\hat{L}}[\psi]+2\nabla H\cdot\delta X\>,

where L^=Δ+|dν|2{\hat{L}}=\Delta+|d\nu|^{2}. Also

δR2=2i=1,2xiδXEi\displaystyle\delta R^{2}=2\sum_{i=1,2}x_{i}\delta X\cdot E_{i} =\displaystyle= 2i=1,2xi(ψνi+(δX)TEi)\displaystyle 2\sum_{i=1,2}x_{i}(\psi\nu_{i}+(\delta X)^{T}\cdot E_{i})
=\displaystyle= 2ψQ^+2R2(δX)T,\displaystyle 2\psi{\hat{Q}}+2\nabla^{\prime}R^{2}\cdot(\delta X)^{T}\>,

where Q^=x1ν1+x2ν2{\hat{Q}}=x_{1}\nu_{1}+x_{2}\nu_{2}. Combining this with (3.1), we have

(3.2) δ=L[ψ]+T,\delta\ell=L[\psi]+\nabla\ell\cdot T\>,

where L[ψ]=Δψ+(|dν|2+aQ^)ψL[\psi]=\Delta\psi+(|d\nu|^{2}+a{\hat{Q}})\psi\>. Since we are assuming Λ0=\ell\equiv\Lambda_{0}= constant, the second term above vanishes and the second variation formula for variations vanishing on Σ\partial\Sigma then reads

(3.3) δa,Λ0=ΣψL[ψ]dΣ=Σψ(Δψ+(|dν|2+aQ^ψ)dΣ=Σ|ψ|2(|dν|2+aQ^)ψ2dΣ.\delta{\mathcal{E}}_{a,\Lambda_{0}}=-\int_{\Sigma}\psi L[\psi]\>d\Sigma=-\int_{\Sigma}\psi(\Delta\psi+(|d\nu|^{2}+a{\hat{Q}}\psi)\>d\Sigma=\int_{\Sigma}|\nabla\psi|^{2}-(|d\nu|^{2}+a{\hat{Q}})\psi^{2}\>d\Sigma\>.

This formula can be found in [3]. As usual, an equilibrium surface will be called stable if the second variation is non negative for all compactly supported variations satisfying the additional condition

(3.4) Σψ𝑑Σ=0.\int_{\Sigma}\psi\>d\Sigma=0\>.

This is just the first order condition which is necessary and sufficient for the variation to be volume preserving.

For a part of the surface of the form α×[h/2,h/2]\alpha\times[-h/2,h/2] this can be written

δ2a,Λ0\displaystyle\delta^{2}{\mathcal{E}}_{a,\Lambda_{0}} =\displaystyle= αh/2h/2(11+ω2ξ12([1+ω2R2]ψs22ωξ2ψsψt+ψt2)\displaystyle\int_{\alpha}\int_{-h/2}^{h/2}\biggl{(}\frac{1}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}([1+\omega^{2}R^{2}]\psi_{s}^{2}-2\omega\xi_{2}\psi_{s}\psi_{t}+\psi_{t}^{2})
(4H22K+aξ2)1+ω2ξ12ψ2)dtds,\displaystyle-(4H^{2}-2K+a\xi_{2})\sqrt{1+\omega^{2}\xi_{1}^{2}}\psi^{2}\biggr{)}\>dt\>ds\>,

where KK denotes the Gaussian curvature. Choosing ψ=sin((2πt)/h)\psi=\sin((2\pi t)/h), gives

Σ|ψ|2𝑑Σ=2π2hα11+ω2ξ12𝑑s.\int_{\Sigma}|\nabla\psi|^{2}\>d\Sigma=\frac{2\pi^{2}}{h}\int_{\alpha}\frac{1}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}\>ds\>.

In addition, for this choice of ψ\psi, we have ψ0\psi\equiv 0 on the boundary and the mean value of ψ\psi on α×[h/2,h/2]\alpha\times[-h/2,h/2] is zero.

Lemma 3.1.

There holds

αK1+ω2ξ12𝑑s=0,\int_{\alpha}K\sqrt{1+\omega^{2}\xi_{1}^{2}}\>ds=0\>,

and hence

α×[h/2,h/2]K𝑑Σ=0.\int_{\alpha\times[-h/2,h/2]}K\>d\Sigma=0\>.
Proof.

From calculations found in [4], one finds

K=ω2(1+κξ2)(1+ω2ξ12)2=ω2(ξ1)s(1+ω2ξ12)2K=\frac{-\omega^{2}(1+\kappa\xi_{2})}{(1+\omega^{2}\xi_{1}^{2})^{2}}=\frac{-\omega^{2}(\xi_{1})_{s}}{(1+\omega^{2}\xi_{1}^{2})^{2}}
αK1+ω2ξ12𝑑s=αω2(ξ1)s(1+ω2ξ12)3/2𝑑s=0,\int_{\alpha}K\sqrt{1+\omega^{2}\xi_{1}^{2}}\>ds=\int_{\alpha}\frac{-\omega^{2}(\xi_{1})_{s}}{(1+\omega^{2}\xi_{1}^{2})^{3/2}}\>ds=0\>,

since the last integrand is the ss derivative of a function of ξ1\xi_{1}. ∎

Proposition 3.2.

A necessary condition for the stability of α×[h/2,h/2]\alpha\times[-h/2,h/2] for the fixed boundary problem is that

(3.6) 2π2h2α11+ω2ξ12𝑑sα4H21+ω2ξ12𝑑s+a𝒜\frac{2\pi^{2}}{h^{2}}\int_{\alpha}\frac{1}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}\>ds\geq\int_{\alpha}4H^{2}\sqrt{1+\omega^{2}\xi_{1}^{2}}\>ds+a{\mathcal{A}}\>

holds. Equivalently, this can be expressed as

(3.7) 2π2h2α11+ω2ξ12𝑑sα×[h/2,h/2]4H2𝑑Σ+a𝒱(α×[h/2,h/2]).\frac{2\pi^{2}}{h^{2}}\int_{\alpha}\frac{1}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}\>ds\geq\int_{\alpha\times[-h/2,h/2]}4H^{2}\>d\Sigma+a{\mathcal{V}}(\alpha\times[-h/2,h/2])\>.
Proof.

We choose ψ=sin((2πt)/h)\psi=\sin((2\pi t)/h) in the second variation formula. For this choice of ψ\psi, we have ψ0\psi\equiv 0 on the boundary and the mean value of ψ\psi on α×[h/2,h/2]\alpha\times[-h/2,h/2] is zero. The result then follows directly from (3) and the previous lemma. ∎

The bound (3.6) gives a condition on the maximum height of a stable helicoidal surface in terms of the geometry of the generating curve.

There is no possible way to obtain a positive lower bound for the right hand side of (3.6). For a round cylinder of radius RR, the equation 2ξ21+ω2ξ12+Λ0R2aR44=c\frac{2\xi_{2}}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}+\Lambda_{0}R^{2}-\frac{aR^{4}}{4}=c becomes 2R+Λ0R2aR44=c2R+\Lambda_{0}R^{2}-\frac{aR^{4}}{4}=c, so for arbitrary aa, we can simply define cc by this equation so and hence the cylinder will be an equilibrium surface. For a cylinder, the potential in the second variation formula is

4H22K+aQ^=1R2+aR,4H^{2}-2K+a{\hat{Q}}=\frac{1}{R^{2}}+aR\>,

so for a<<0a<<0 the potential is non positive and the cylinder is stable for arbitrary heights.

We will now give an upper bound for the height of a stable helicoidal equilibrium surface which is valid for any such surface which is not a cylinder over a planar curve. This upper bound will only depend on the generating curve. In [7], this estimate is modified so that it applies to non circular cylindrical equilibrium surfaces as well.

Theorem 3.3.

For a helicoidal surface which is not a round cylinder, a necessary condition for the stability of the part the surface between horizontal planes separated by a distance hh is that

(3.8) 4π2e4h2ω2α(1+ω2R2(1+κξ2)2(1+ω2ξ12)7/2𝑑sα11+ω2ξ12𝑑s(ω2),\frac{4\pi^{2}e^{4}}{h^{2}}\geq\frac{\omega^{2}\oint_{\alpha}\frac{(1+\omega^{2}R^{2}(1+\kappa\xi_{2})^{2}}{(1+\omega^{2}\xi_{1}^{2})^{7/2}}\>ds}{\oint_{\alpha}\frac{1}{1+\omega^{2}\xi_{1}^{2}}\>ds}(\geq\omega^{2}),

holds. The result also holds true if a=0a=0, i.e. if the surface has constant mean curvature.

Remark. In [7] a similar estimate is given for cylindrical surfaces which are not round cylinders.

Proof.

To begin, note that the the third component of the normal ν3\nu_{3} satisfies L[ν3]=0L[\nu_{3}]=0 since vertical translation is a symmetry of the normal. Also, this function will vanish identically if and only if the surface is a cylinder.

The function ν3\nu_{3} can be written [4]

ν3=ωξ11+ω2ξ12,\nu_{3}=\frac{\omega\xi_{1}}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}\>,

so ν3\nu_{3} is a function of ss only. Using local coordinate expressions found in [4], we can write

0\displaystyle 0 =\displaystyle= L[ν3]\displaystyle L[\nu_{3}]
=\displaystyle= 1g(gg11(ν3)s)s+(|dν|2+aξ2)ν3\displaystyle\frac{1}{\sqrt{g}}(\sqrt{g}g^{11}(\nu_{3})_{s})_{s}+(|d\nu|^{2}+a\xi_{2})\nu_{3}
=\displaystyle= 11+ω2ξ12[1+ω2R21+ω2ξ12(ν3)s]s+(|dν|2+aξ2)ν3\displaystyle\frac{1}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}\bigl{[}\frac{1+\omega^{2}R^{2}}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}(\nu_{3})_{s}\bigr{]}_{s}+(|d\nu|^{2}+a\xi_{2})\nu_{3}
=:\displaystyle=: [ν3].\displaystyle{\mathcal{L}}[\nu_{3}]\>.

Note that (|dν|2+aξ2)(|d\nu|^{2}+a\xi_{2}) only depends on ss. For any smooth function u=u(s)u=u(s), there holds

[eu]=eu([u]+g11us2)=eu([u]+(1+ωR2)us2).{\mathcal{L}}[e^{u}]=e^{u}({\mathcal{L}}[u]+g^{11}u_{s}^{2})=e^{u}({\mathcal{L}}[u]+(1+\omega R^{2})u_{s}^{2})\>.

If we now take ψ=eν3(s)sin((2πt)/h)\psi=e^{\nu_{3}(s)}\sin((2\pi t)/h), then (3.4) holds and we can obtain from (3),

δ2a,Λ0=2π2hαe2ν31+ω2ξ12𝑑sh2αe2ν3(1+ω2R2)1+ω2ξ12((ν3)s)2𝑑s\delta^{2}{\mathcal{E}}_{a,\Lambda_{0}}=\frac{2\pi^{2}}{h}\oint_{\alpha}\frac{e^{2\nu_{3}}}{1+\omega^{2}\xi_{1}^{2}}\>ds-\frac{h}{2}\oint_{\alpha}\frac{e^{2\nu_{3}}(1+\omega^{2}R^{2})}{\sqrt{1+\omega^{2}\xi_{1}^{2}}}((\nu_{3})_{s})^{2}\>ds

Using 1ν31-1\leq\nu_{3}\leq 1 and using

(ν3)s=(ωξ1(1+ω2ξ12)1/2)s=(ξ1)sω(1+ω2ξ12)3/2=ω(1+κξ2)(1+ω2ξ12)3/2,(\nu_{3})_{s}=(\omega\xi_{1}(1+\omega^{2}\xi_{1}^{2})^{-1/2})_{s}=(\xi_{1})_{s}\omega(1+\omega^{2}\xi_{1}^{2})^{-3/2}=\omega(1+\kappa\xi_{2})(1+\omega^{2}\xi_{1}^{2})^{-3/2},

yields the result. ∎

4. Appendix

We assume that Σ\Sigma is contained in a three dimensional region Ω\Omega and that Σ\partial\Sigma is contained in a supporting surface SS which is part of Ω\partial\Omega. We assume that there is a (not necessarily connected) domain S1SS_{1}\subset S such that ΣS1\Sigma\cup S_{1} bounds a subregion Ω1Ω\Omega_{1}\subset\Omega. The volume of Ω1\Omega_{1} will be denoted by 𝒱{\mathcal{V}}.

Let ϕ\phi be a solution of Δϕ=1\Delta^{\prime}\phi=1 in Ω\Omega with ϕN=0\nabla^{\prime}\phi\cdot N=0 on SS where NN is the outward pointing normal to SS. This boundary value problem is undertermined and is solvable provided SS is not closed.

We subject the surface to a variation which keeps Σ\partial\Sigma on SS. Then δX=:T+ψνN\delta X=:T+\psi\nu\perp N along Σ\partial\Sigma and

𝒱=ΣϕνdΣ.{\mathcal{V}}=\int_{\Sigma}\nabla^{\prime}\phi\cdot\nu\>d\Sigma\>.

We have

δ𝒱\displaystyle\delta{\mathcal{V}} =\displaystyle= ΣT+ψνϕν+ϕδνdΣ+Σϕν(T2Hψ)𝑑Σ\displaystyle\int_{\Sigma}\nabla^{\prime}_{T+\psi\nu}\nabla^{\prime}\phi\cdot\nu+\nabla^{\prime}\phi\cdot\delta\nu\>d\Sigma+\int_{\Sigma}\nabla^{\prime}\phi\cdot\nu(\nabla\cdot T-2H\psi)\>d\Sigma
=\displaystyle= Σψνϕν2HψϕνϕψdΣ+Σ(ϕν)Tn𝑑s\displaystyle\int_{\Sigma}\psi\nabla^{\prime}_{\nu}\nabla^{\prime}\phi\cdot\nu-2H\psi\nabla^{\prime}\phi\cdot\nu-\nabla\phi\cdot\nabla\psi\>d\Sigma+\oint_{\partial\Sigma}(\nabla^{\prime}\phi\cdot\nu)T\cdot n\>ds
=\displaystyle= Σψνϕν2Hψϕν+ψΔϕdΣ\displaystyle\int_{\Sigma}\psi\nabla^{\prime}_{\nu}\nabla^{\prime}\phi\cdot\nu-2H\psi\nabla^{\prime}\phi\cdot\nu+\psi\Delta\phi\>d\Sigma
+Σ((ϕν)Tψϕ)n𝑑s.\displaystyle+\oint_{\partial\Sigma}((\nabla^{\prime}\phi\cdot\nu)T-\psi\nabla\phi)\cdot n\>ds\>.

A well known formula relating the Laplacian on a submanifold to the Laplacian on the ambient space gives Δϕ=νϕν2Hϕ+Δϕ\Delta^{\prime}\phi=\nabla^{\prime}_{\nu}\nabla^{\prime}\phi\cdot\nu-2H\nabla^{\prime}\phi+\Delta\phi. Therefore we obtain

δ𝒱=Σψ𝑑Σ+𝑑X×ϕδX.\delta{\mathcal{V}}=\int_{\Sigma}\psi\>d\Sigma+\oint dX\times\nabla^{\prime}\phi\cdot\delta X\>.

However, all of dXdX,ϕ\nabla^{\prime}\phi and δX\delta X are perpendicular to NN on Σ\partial\Sigma, so the line integral above vanishes.

To obtain (1.2), we let WW be a vector field on Ω\Omega satisfying W=R2\nabla^{\prime}\cdot W=R^{2} and WN=0W\cdot N=0 along SS. This boundary value problem is undertermined and is solvable provided SS is not closed. Then by the Divergence Theorem

Ω1R2d3x=ΣWν𝑑Σ.\int_{\Omega_{1}}R^{2}\>d^{3}x=\int_{\Sigma}W\cdot\nu\>d\Sigma\>.
δΩ1R2d3x\displaystyle\delta\int_{\Omega_{1}}R^{2}\>d^{3}x =\displaystyle= ΣT+ψνWν+WδνdΣ+ΣWν(T2Hψ)𝑑Σ\displaystyle\int_{\Sigma}\nabla^{\prime}_{T+\psi\nu}W\cdot\nu+W\cdot\delta\nu\>d\Sigma+\int_{\Sigma}W\cdot\nu\>(\nabla\cdot T-2H\psi)\>d\Sigma
=\displaystyle= ΣψνWνWψ2HψWνdΣ+Σ(Wν)Tn𝑑s\displaystyle\int_{\Sigma}\psi\nabla^{\prime}_{\nu}W\cdot\nu-W\cdot\nabla\psi-2H\psi W\cdot\nu\>d\Sigma+\oint_{\partial\Sigma}(W\cdot\nu)T\cdot n\>ds
=\displaystyle= ΣψW𝑑Σ+Σ((Wν)TψW)n𝑑s\displaystyle\int_{\Sigma}\psi\nabla^{\prime}\cdot W\>d\Sigma+\oint_{\partial\Sigma}((W\cdot\nu)T-\psi W)\cdot n\>ds
=\displaystyle= ΣψR2𝑑Σ+Σ𝑑X×WδX.\displaystyle\int_{\Sigma}\psi R^{2}\>d\Sigma+\oint_{\partial\Sigma}dX\times W\cdot\delta X\>.

Again, all of dXdX, WW and δX\delta X are perpendicular to NN along Σ\partial\Sigma so the line integral will vanish.

If the pair (ϕ,W)(\phi,W) used above are replaced by another pail (ϕ¯,W¯)({\underline{\phi}},{\underline{W}}) satisfying the same equations: Δϕ¯=1\Delta^{\prime}{\underline{\phi}}=1 and W¯=R2\nabla^{\prime}{\underline{W}}=R^{2}, then by the Divergence Theorem

Σϕ¯νdΣ=𝒱+c1,\int_{\Sigma}\nabla^{\prime}{\underline{\phi}}\cdot\nu\>d\Sigma={\mathcal{V}}+c_{1}\>,
ΣWν𝑑Σ=Ω1R2d3x+c2,\int_{\Sigma}W\cdot\nu\>d\Sigma=\int_{\Omega_{1}}R^{2}\>d^{3}x+c_{2}\>,

for constants c1c_{1} and c2c_{2}. Thus these replacements will not affect the variational formulas for these integrals.

References

  • [1] Brown, R.A., Scriven, L.E., The Shape and Stability of Rotating Liquid Drops, Proc. R. Soc. Lond. A 30 June 1980 vol. 371 no. 1746 331-357
  • [2] Chandrasekhar, S. The Stability of a Rotating Liquid Drop Proc. R. Soc. Lond. A 25 May 1965 vol. 286 no. 1404 1-26.
  • [3] López, Rafael, Stationary rotating surfaces in Euclidean space, Calculus of Variations in Partial Differential Equations, 39, numbers 3-4, (2010), 333-359.
  • [4] Perdomo, O. A dynamical interpretation of cmc Twizzlers surfaces, Pacific J. Math. 258 (March 2012) No. 2, 459-585.
  • [5] Perdomo, O. Embedded constant mean curvature hypersurfaces of spheres, Asian J. Math. 14 (March 2010) No. 1, 73-108.
  • [6] Perdomo, O. Hellicoidal minimal surfaces in R3R^{3}, To appear in the Ilinois J. of Math.
  • [7] Palmer, B. and Perdomo, O. Equilibrium Shapes of Cylindrical Rotating Liquid Drops. Preprint.
  • [8] Solonikov, V., A., Instability of axially symmetric equilibrium figures of a rotating, viscous, incompressible liquid, Journal of Mathematical Sciences July 2006, Volume 136, Issue 2, pp 3812-3825.