Longtime Dynamics of Irrotational Spherical Water Drops: Initial Notes
In this note, we propose several unsolved problems concerning the irrotational oscillation of a water droplet under zero gravity. We will derive the governing equation of this physical model, and convert it to a quasilinear dispersive partial differential equation defined on the sphere, which formally resembles the capillary water waves equation but describes oscillation defined on curved manifold instead. Three types of unsolved mathematical problems related to this model will be discussed in observation of hydrodynamical experiments under zero gravity111There are numbers of visual materials on such experiments conducted by astronauts. See for example https://www.youtube.com/watch?v=H_qPWZbxFl8&t or https://www.youtube.com/watch?v=e6Faq1AmISI&t.: (1) Strichartz type inequalities for the linearized problem (2) existence of periodic solutons (3) normal form reduction and generic lifespan estimate. It is pointed out that all of these problems are closely related to certain Diophantine equations, especially the third one.
1. Capillary Spherical Water Waves Equation: Derivation
1.1. Water Waves Equation for a Bounded Water Drop
Comparing to gravity water waves problems, the governing equation for a spherical droplet of water under zero gravity takes a very different form. At a first glance it looks similar to the water waves systems as mentioned above, but some crucial differences do arise after careful analysis. To the author’s knowledge, besides those dealing with generic free-boundary Euler equation ([CS07], [SZ08]), the only reference on this problem is Beyer-Günther [BG98], in which the local well-posedness of the equation is proved using a Nash-Moser type implicit function theorem. We will briefly discribe known results for gravity water waves problems in the next subsection.
To start with, let us pose the following assumptions on the fluid motion that we try to describe:
-
•
(A1) The perfect, irrotational fluid of constant density occupies a smooth, compact region in .
-
•
(A2) There is no gravity or any other external force in presence.
-
•
(A3) The air-fluid interface is governed by the Young-Laplace law, and the effect of air flow is neglected.
We assume that the boundary of the fluid region has the topological type of a smooth compact orientable surface , and is described by a time-dependent embedding . We will denote a point on by , the image of under by , and the region enclosed by by . The outer normal will be denoted by . We also write for the flat connection on .
Adopting assumption (A3), we have the Young-Laplace equation:
where is the (scalar) mean curvature of the embedding, is the surface tension coefficient (which is assumed to be a constant), and are respectively the inner and exterior air pressure at the boundary; they are scalar functions on the boundary and we assume that is a constant. Under assumptions (A1) and (A2), we obtain Bernoulli’s equation, sometimes referred as the pressure balance condition, on the evolving surface:
(1.1) |
where is the velocity potential of the velocity field of the air. Note that is determined up to a function in , so we shall leave the constant around for convenience reason that will be explained shortly. According to assumption (A1), the function is a harmonic function within the region , so it is uniquely determined by its boundary value, and the velocity field within is . The kinematic equation on the free boundary is naturally obtained as
(1.2) |
Finally, we would like to discuss the conservation laws for (1.1)-(1.2). The preservation of volume is a consequence of incompressibility. The system describes an Eulerian flow without any external force, so the center of mass moves at a uniform speed along a fixed direction, i.e.
(1.3) |
with Vol being the Lebesgue measure, marking points in , and being the velocity and starting position of center of mass respectively. Furthermore, the total momentum is conserved, and since the flow is a potential incompressible one, the conservation of total momentum is expressed as
(1.4) |
Most importantly, it is not surprising that (1.1)-(1.2) is a Hamilton system (the Zakharov formulation for water waves; see Zakharov [Zak68]), with Hamiltonian
(1.5) |
i.e. potential proportional to surface area plus kinetic energy of the fluid.
1.2. Converting to a Differential System
It is not hard to verify that the system (1.1)-(1.2) is invariant if is composed with a diffeomorphism of ; we may thus regard it as a geometric flow. If we are only interested in perturbation near a given configuration, we may reduce system (1.1)-(1.2) to a non-degenerate dispersive differential system concerning two scalar functions defined on , just as Beyer and Günther did in [BG98]. In fact, during a short time of evolution, the interface can be represented as the graph of a function defined on the initial surface: if is a fixed embedding close to the initial embedding , we may assume that , where is a scalar “height” function defined on and is the outer normal vector field of .

With this observation, we shall transform the system (1.1)&(1.2) into a non-local system of two real scalar functions defined on , where is the “height” function described as above, and is the boundary value of the velocity potential, pulled back to the underlying manifold .
The operator
maps the pulled-back Dirichlet boundary value to the boundary value of the gradient of . We shall write
for its normal part, where is the Dirichlet-Neumann operator corresponding to the region enclosed by the image of . Thus
We also need to calculate the restriction of on in terms of and . By the chain rule,
We thus arrive at the following nonlinear system:
(EQ(M)) |
where .
Remark 1.
We may obtain an explicit expression of in terms of (together with the connection on the fixed embedding ), just as standard references did for Euclidean or periodic water waves, but that is not necessary for our discussion at the moment. It is important to keep in mind that the preservation of volume and conservation of total momentum (1.3)-(1.4) convert to integral equalities of . These additional restrictions are not obvious from the differential equations (EQ(M)), though they can be deduced from (EQ(M)) since they are just rephrase of the original physical laws (1.1)-(1.2).
For , the case that we shall discuss in detail, we refer to the system as the capillary spherical water waves equation. To simplify our discussion, we shall be working under the center of mass frame, and require the eigenmode to vanish for all . This could be easily accomplished by absorbing the eigenmode into since the equation is invariant by a shift of . In a word, from now on, we will be focusing on the non-dimensional capillary spherical water waves equation
(EQ) |
where , and . We assume that the total volume of the fluid is , so that the preservation of volume is expressed as
(1.6) |
where is the standard measure on . The inertial movement of center of mass (1.3) and conservation of total momentum (1.4) under our center of mass frame are expressed respectively as
(1.7) |
where is the induced surface measure. Further, the Hamiltonian of the system is
(1.8) |
and for a solution there holds .
Up to this point, we are still working within the realm of well-established frameworks. We already know that the general free-boundary Euler equation is locally well-posed due to the work of [CS07] or [SZ08], and due to the curl equation the curl free condition persists during the evolution. On the other hand, the Cauchy problem of system (1.1) and (1.2) is known to be locally well-posed, due to Beyer and Günther in [BG98]. They used an iteration argument very similar to a Nash-Moser type argument in the sense that it involves multiple scales of Banach spaces and “tame” maps in a certain sense. Finally, it is not hard to transplant the potential-theoretic argument of Wu [Wu99] to prove the local well-posedness.
To sum up, we already know that the system (EQ(M)) for a compact orientable surface (hence (EQ) specifically) is locally well-posed. But this is all we can assert for the motion of a water droplet under zero gravity. In the following part of this note, we will propose several questions and conjectures concerning the long-time behaviour of water droplets under zero gravity.
1.3. Previous Works on Water Waves
There has already been several different approaches to describe the motion of perfect fluid with free boundary. One is to consider the motion of perfect fluid occupying an arbitrary domain in or , either with or without surface tension. The motion is described by a free boundary value problem of Euler equation. This generic approach was employed by Countand-Shkoller [CS07] and Shatah-Zeng [SZ08]. Both groups proved the local-wellposedness of the problem. This approach has the advantage of being very general, applicable to all geometric shapes of the fluid.
On the other hand, when coming to potential flows of a perfect fluid, the curl free property results in dispersive nature of the problem. The motion of a curl free perfect fluid under gravity and a free boundary value condition is usually referred to as the gravity water waves problem. The first breakthroughs in understanding local well-posedness were works of Wu [Wu97] [Wu99], who proved local well-posedness of the gravity water waves equation without any smallness assumption. Lannes [Lan05] extended this to more generic bottom shapes. Taking surface tension into account, the problem becomes gravity-capillary water waves. Schweizer [Sch05] proved local well-posedness with small Cauchy data of the gravity-capillary water waves problem, and Ming-Zhang [MZ09] proved local well-posedness without smallness assumption. Alazard-Metevier [AM09] and Alazard-Burq-Zuily [ABZ11] used para-differential calculus to obtain the optimal regularity for local well-posedness of the water waves equation, either with or without surface tension.
For discussion of long time behavior, it is important to take into account the dispersive nature of the problem. For linear dispersive properties, there has been work of Christianson-Hur-Staffilani [CHS10]. For gravity water waves living in , works on lifespan estimate include Wu [Wu09] and Hunter-Ifrim-Tataru [HIT16] (almost global result), Ionescu-Pusateri [IP15] and Alazard-Delort [AD13] and Ifrim-Tataru [IT14] (global result). For gravity water waves living in , there are works of Germain-Masmoudi-Shatah [GMS12] and Wu [Wu11] (no surface tension), Germain-Masmoudi-Shatah [GMS15] (no gravity), Deng-Ionescu-Pausader-Pusateri [DIPP17] (gravity-capillary water waves) and Wang [Wan20] (gravity-capillary water waves with finite depth). These results all employed different forms of decay estimates derived from dispersive properties.
As for long time behavior of periodic water waves, Berti-Delort [BD+18] considered gravity-capillary water waves defined on , Berti-Feola-Pusateri [BFP22] considered gravity water waves defined on , Ionescu-Pusateri [IP19] considered gravity-capillary water waves defined on , and obtained an estimate on the lifespan beyond standard energy method. All of the three groups used para-differential calculus and suitable normal form reduction; the results of Berti-Delort and Ionescu-Pusateri were proved for physical data of full Lebesgue measure.
To sum up, all results on the gravity water waves problem listed above are concerned with an equation for two scalar functions defined on a fixed flat manifold, being one of the following: , , , , sometimes called the “bottom” of the fluid. These two functions represents the geometry of the liquid-gas interface and the boundary value of the velocity potential, respectively. The manifold itself is considered as the bottom of the container in which all dynamics are performed. We observe that the differential equation (EQ(M)) is, mathematically, fundamentally different from the water waves equations that have been well-studied.
2. Initial Notes on Unsolved Problems
In this section, we propose unsolved problems related to the spherical capillary water waves system (EQ(M)) with . Not surprisingly, these problems all have deep backgrounds in number theory.
2.1. Linearization Around the Static Solution
We are mostly interested in the stability of the static solution of (EQ(M)). A static solution should be a fluid region whose shape stays still, with motion being a mere shift within the space. In this case, we have since the reference is relatively static with respect to the air. Moreover, the velocity field and “potential of acceleration” must both be spatially uniform, so that the left-hand-side of the pressure balance condition (1.1) is a function of alone. It follows that is always a compact embedded surface of constant mean curvature, hence in fact always an Euclidean sphere by the Alexandrov sphere theorem (see [MP19]), and we may just take . Moreover, since should enclose a constant volume by incompressibility, the radius of that sphere does not change. After suitable scaling, we may assume that the radius is always 1, and , , to make (1.1)-(1.2) non-dimensional. Finally, by choosing the center of mass frame, we may simply assume that the spatial shift is always zero, so that the velocity potential , a real constant. It is harmless to fix it to be zero.
Thus, under our convention, a static solution of (1.1)-(1.2) takes the form
(2.1) |
where is a constant vector, and is the standard embedding of as . Equivalently, this means that a static solution of (EQ(M)) under our convention must be . Note here that the Gauss map of coincides with itself.
We can now start our perturbation analysis around a static solution at the linear level. Let be the space of spherical harmonics of order , normalized according to the standard surface measure on . In particular, is spanned by three components of . Let be the orthogonal projection on onto , be the orthogonal projection on onto , be the orthogonal projection on onto . For , the linearization of around the sphere is , where is the Laplacian on the sphere ; cf. the standard formula for the second variation of area in [BC12]. Then acts on as the multiplier . Note that even if we consider the dimensional form of (EQ), there will only be an additional scaling factor , where is the radius of the sphere. On the other hand, the following solution formula for the Dirichlet problem on is well-known: if , then the harmonic function in with Dirichlet boundary value is determined by
where is the spherical coordinate in . Thus the Dirichlet-Neumann operator acts on as the multiplier . Note again that even if we consider the dimensional form (EQ), there will only be an additional scaling factor .
Thus, setting
we find that the linearization of (EQ) around the static solution (2.1) is a linear dispersive equation
(2.2) |
where the -order elliptic operator is given by a multiplier
Note that is completely determined by . At the linear level, there must hold because the first variation of volume must be zero; and because of the conservation laws (1.7).
Let us also re-write the original nonlinear system (EQ) into a form that better illustrates its perturbative nature. For simplicity, we use to abbreviate a quantity that can be controlled by -linear expressions in , and disregard its continuity properties for the moment. For example, is an expression of order when .
Since the operator acts degenerately on , we should be more careful about the eigenmodes and . The volume preservation equation (1.6) implies . Projecting (EQ) to , which is spanned by the components of , we obtain since the conservation law (1.7) implies ; and since annihilates . We can thus formally re-write the nonlinear system (EQ) as the following:
(2.3) |
with vanishing quadratically as . Note that we are disregarding all regularity problems at the moment.
2.2. Question at Linear Level
At the linear level, our first unanswered question is
Question 1.
Does the solution of the linear capillary spherical water waves equation (2.2) satisfy a Strichartz type estimate of the form
where , and the admissible indices and should be determined?
Answer to Question 1 should be important in understanding the dispersive nature of linear capillary spherical water waves. For Schrödinger equation on a compact manifold, a widly cited result was obtained by Burq-Gérard-Tzvetkov [BGT04]:
Theorem 2.1.
On a general compact Riemannian manifold , there holds
where
The authors used a time-localization argument for the parametrix of to prove this result. For the sphere , this inequality is not optimal. The authors further used a Bourgain space argument to obtain the optimal Strichartz inequality:
Theorem 2.2.
Let be the standard n-dimensional sphere. For a function , there holds the Strichartz inequality
where
Furthermore these inequalities are optimal in the sense that the Sobolev index cannot be less than or equal to .
The proof is the consequence of two propositions. The first one is the “decoupling inequality on compact manifolds”, in particular the following result proved by Sogge [Sog88]:
Proposition 2.1.
Let be the spectral projection to eigenspaces with eigenvalues in on . Then there holds
where
These estimates are sharp in the following sense: if is a zonal spherical harmonic function of degree on , then as ,
The second one is a Bourgain space embedding result:
Proposition 2.2.
For a function , define the Bourgain space norm
Then for and , there holds
The key ingredient for proving this proposition is the following number-theoretic result:
As for optimality of the Strichartz inequality, the authors of [BGT04] implemented standard results of Gauss sums.
The parametrix and Bourgain space argument can be repeated without essential change for the linear capillary spherical water waves equation (2.2), but this time the Bourgain space argument would be more complicated: the Bourgain space norm is now
and the embedding result becomes
for and all , , where is the infimum of all exponents such that when , the number
Some basic analytic number theory implies , and thus the range of is . Surprisingly this index is not better than that predicted by the parametrix method. It remains unknown whether this index could be further optimized.
To close this subsection, we note that the capillary spherical water wave lives on a compact region, so the dispersion does not take away energy from a locality to infinity. This is a crucial difference between waves on compact regions and waves in Euclidean spaces. In particular, we do not expect decay estimate for . For the nonlinear problem (EQ), techniques like vector field method (Klainerman-Sobolev type inequalities) do not apply.
2.3. Rotationally Symmetric Solutions: Bifurcation Analysis
Illuminated by observations in hydrodynamical experiments under zero gravity, and suggested by the existence of standing gravity capillary water waves due to Alazard-Baldi [AB15], we propose the following conjecture:
Conjecture 2.1.
There is a Cantor family of small amplitude periodic solutions to the spherical capillary water waves system (EQ).
Let us conduct the bifurcation analysis that suggests why this conjecture should be true. By time rescaling, we aim to find solution of the following system that is -peiodic in :
(2.4) |
together with the conservation laws (1.6)-(1.8). Here we refer as the fundamental frequency. The linearization of this system at the equilibrium is
(2.5) |
We restrict to rotationally symmetric solutions of the system: that is, water droplets which are always rotationally symmetric with a fixed axis. In addition, we require to be even in and to be odd in . The solution thus should take the form
where is the ’th zonal spherical harmonic, i.e. the (unique) normalized spherical harmonic of degree that is axially symmetric. In spherical coordinates this means that , where is the ’th Legendre polynomial. Since are irrelevant we fix them to be 0. Then
In order that , at the level , we must have for all . At the level , we have , and for there holds and , so for all . Hence can be nonzero only for and .
Consequently, has a one-dimensional kernel if and only if the Diophantine equation
(2.6) |
has exactly one solution . If has this property, then at the linear level, the lowest frequency of oscillation is
We look into this Diophantine equation. Obviously has to be a rational number. The equation is closely related to a family of elliptic curves over :
If we set (irreducible fraction), then integral solutions of (2.6) are in 1-1 correspondence with integral points with natural number coordinates on the elliptic curve , under the following map:
Thus we just need to find natural numbers such that there is a unique (up to negation) integral point , where is divided by , and is divided by . We seek for as small as possible with such property, which gives lowest frequency of oscillation as small as possible. For , we find that if , then the integral points on elliptic curve (up to negation of Mordel-Weil group) are
The only point with and is (90,900), which gives , and the lowest frequency of oscillation is .
There are other choices of . We list down the value of below 50, the corresponding and the lowest frequency :
15 | 6 | |
17 | 49 | |
22 | 9 | |
26 | 50 | |
42 | 7 | |
46 | 576 | |
50 | 25 |
See Appendix A for the MAGMA code used to find these values. This list suggests that might be the smallest order that meets the requirement, but this remains unproved. We summarize these into the following number-theoretic question:
Question 2.
For the family of elliptic curves
how many choices of are there such that, there is exactly one integral point with and , ? For such and integral point , is the minimal value of exactly 6, or is it smaller?
Of course, a complete answer of Question 2 should imply very clear understanding of periodic solutions of the spherical capillary water waves equation constructed using bifurcation analysis. But at this moment we are satisfied with existence, so we may pick any and that meets the requirement, for example the simplest case , and any of the following choices of and :
We thus refine our conjecture as follows:
Conjecture 2.2.
Let be a pair of natural numbers with , , and set . Suppose that the only natural number solution of the Diophantine equation
is . Then there is a Cantor set with positive measure of parameters , clustered near , such that the spherical capillary water waves equation (2.4) admits small amplitude periodic solution with frequency .
The counterpart of Conjecture 2.2 for gravity capillary standing water waves was proved by Alazard-Baldi [AB15] using a Nash-Moser type theorem. The key technique in their proof was to find a conjugation of the linearized operator of the gravity capillary water waves system on to an operator of the form
where is an elliptic Fourier multiplier of order , and are real constants. The frequency lives in a Cantor type set that clusters around a given frequency so that the kernel of the linearized operator is 1-dimensional. With this conjugation, they were able to find periodic solutions of linearized problems required by Nash-Moser iteration.
It is expected that this technique could be transplanted to the equation (2.4), since our analysis for (2.6) suggests that the 1-dimensional kernel requirement for bifurcation analysis is met. It seems that the greatest technical issue is to find a suitable conjugation that takes the linearized operator of (2.4) to an operator of the form
where each is a real Fourier multiplier acting on spherical harmonics. The difficulty is that, since we are working with pseudo-differential operators on , the formulas of symbolic calculus are not as neat as those on flat spaces. It seems necessary to implement some global harmonic analysis for compact homogeneous spaces, e.g. extension of results collected in Ruzhansky-Turunen [RT09]. Unfortunately, those results do not include pseudo-differential operators with “rough coefficients” and para-differential operators, so it seems necessary to re-write the whole theory.
2.4. Number-Theoretic Obstruction with Normal Form Reduction
As pointed out in Section 1, the system (2.3) is locally well-posed due to a result in [BG98]. General well-posedness results [CS07], [SZ08] for free boundary value problem of Euler equation also apply. As for lifespan estimate for initial data -close to the static solution (2.1), it should not be hard to conclude that the lifespan should be bounded below by . The result relates to the fact that the sphere is a stable critical point of the area functional, cf. [BC12]. This is nothing new: a suitable energy inequality should imply it. However, although the clue is clear, the implementation is far from standard since we are working on a compact manifold. A rigorous proof still calls for hard technicalities.
Now we will be looking into the nonlinear equation (2.3) for its longer time behavior. Although appearing similar to the well-studied water waves equation in e.g. [IP19], [Lan05], [Wu97], [Wu99], there is a crucial difference between the dispersive relation in (2.3) and the well-studied water waves equations: the dispersive relation exhibits a strong rigidity property, i.e. the arbitrary physical constants enter into the dispersive relation only as scaling factors. For the gravity-capillary water waves, the linear dispersive relation reads
where is the gravitational constant and is the surface tension coefficient. For the Klein-Gordon equation on a Riemannian manifold, the linear dispersive relation reads
where is the mass. In [IP19], Ionescu and Pusateri referred such dispersive relations as having non-degenerate dependence on physical parameter, while in our context it is appropriate to refer to the dependence as degenerate. We will see that this crucial difference brings about severe obstructions for the long-time well-posedness of the system.
Following the idea of Delort and Szeftel [DS04], we look for a normal form reduction of (2.3) and explain why the rigidity property could cause obstructions. Not surprisingly, the obstruction is due to resonances, and strongly relates to the solvablity of a Diophantine equation. Delort and Szeftel cast a normal form reduction to the small-initial-data problem of quasilinear Klein-Gordon equation on the sphere and obtained an estimate on the lifespan longer than the one provided by standard energy method. After their work, normal form reduction has been used by mathematicians to understand water waves on flat tori, for example [BD+18], [BFP22] and [IP19]. The idea was inspired by the normal form reduction method introduced by Shatah [Sha85]: for a quadratic perturbation of a linear dispersive equation
using a new variable with a suitably chosen quadratic addendum can possibly eliminate the quadratic part of , thus extending the lifespan estimate beyond the standard .
So we shall write the quadraticr part of as
where are complex bi-linear operators, following the argument of Section 4 in [DS04]. They are independent of since the right-hand-side of the equation does not depend on explicitly. Let’s look for a diffeomorphism
in the function space , where is a bilinear operator, so that the equation (2.3) with quadratic nonlinearity reduces to an equation with cubic nonlinearity. The is supposed to take the form , with
where the ’s are complex numbers to be determined. Implementing (2.3), we find
We aim to eliminate most of the second order portions of . The coefficients are fixed as follows:
(2.7) | ||||
then a large portion of the second order part of will be eliminated. In fact, for , if , then the term
is cancelled out. On the other hand, by the volume preservation equality (1.6) and conservation law (1.7), there holds , , so the low-low interaction
is automatically .
Thus the existence and continuity of the normal form depends on the property of the 3-way resonance equation
(2.8) |
which is equivalent to the Diophantine equation
(2.9) |
where .
If the tuple is non-resonant, i.e. it is such that , then some elementary number theoretic argument will give a lower bound on in terms of a negative power (can be fixed as ) of . This is usually referred as small divisor estimate.
To study the distribution of resonant frequencies, we propose the following unsolved question:
Question 3.
Does the Diophantine equation (2.9) have finitely many solutions?
However, the Diophantine equation (2.9) does admit non-trivial solutions and . In other words, the second order terms e.g.
in the quadratic part of cannot be eliminated by normal form reduction. On the other hand, it seems to be very hard to determine whether (2.9) still admits any other solution. We have the following proposition (the author would like to thank Professor Bjorn Poonen for the proof):
Proposition 2.3.
The Diophantine equation (2.9) has no solution with other than and .
The proof of this proposition is computer-aided. The key point is to use the so-called Runge’s method to show that if is a solution, then there must hold . For a given , this reduces the proof to numerical verification for finitely many possibilities. The algorithm can of course be further optimized, but due to some algebraic geometric considerations, it is reasonable to conjecture that the solution of (2.9) should be very rare. In fact, there are two ways of viewing the problem. We observe that if is a solution, then must be a square, and the square free part of must be the same. Further reduction turns the problem into finding integral points on a family of elliptic curves
which is of course difficult, but since Siegel’s theorem asserts that there are only finitely many integer points on an elliptic curve over , it is reasonable to conjecture that there are not “too many” solutions to (2.9). We may also view the problem as finding integral (rational) points on a given algebraic surface. The complex projective surface corresponding to (2.9) is given by
where is the homogeneous coordinate on . With the aid of computer, we obtain
Proposition 2.4.
The complex projective surface has Kodaira dimension 2 (i.e. it is of general type under Kodaira-Enriquez classification), and its first Betti number is 0.
See Appendix A for the code.
The first part of the proposition suggests that the rational points of should be localized on finitely many algebraic curves laying on ; nevertheless, this seemingly simple suggestion is indeed a special case of the Bomberi-Lang conjecture, a hard problem in number theory (its planar case is known as the celebrated Faltings’s theorem). The second part suggests that the rational points of should be rare since its Albanese variety is a single point. But these are just heuristics that solutions to (2.9) should be rare. In general, determining the solvability of a given Diophantine equation is very difficult 222For example, the seemingly simple Diophantine equation is in fact a puzzle of more than 60 years, and its first solution was found recently by Booker-Sutherland [BS21]. It is of extremely large magnitude: . Another example is the equation of same type . Beyond the easily found solutions (1, 1, 1), (4, 4, -5), (4, -5, 4), (-5, 4, 4), the next solution reads ., as number theorists and arithmetic geometers generally believe.
The reason that such issues do not occur for water waves in the flat setting or nonlinear Klein-Gordon equations is twofold. First of all, the resonance equation is easily understood even in the degenerate case in the flat setting. For example, the capillary water waves without gravity on has dispersive relation , and the 3-way resonance equation is
with the additional requirement . We already know that the resonance equation has no non-trivial solution at all, cf. [BFP22]. But even without using , we would be able to conclude that there are at most finitely many non-trivial solutions from the celebrated Faltings’s theorem on rational points on high-genus algebraic projective curves (although this is like using a sledge hammer to crack a nut). Secondly, for the non-degenrate case, for example the gravity-capillary waves, the dispersive relation reads , so if the ratio is a transcedental number then the 3-way resonance equation has no solution. Furthermore, using some elementary calculus and a measure-theoretic argument, it can be shown, not without technicalities, that the resonances admit certain small-divisor estimates for almost all parameters. This is exactly the argument employed by Delort-Szeftle [DS04], Berti-Delort [BD+18] and Ionescu-Pusateri [IP19], so that their results were stated for almost all parameters. These parameters are, roughly speaking, badly approximated by algebraic numbers.
However, the resonance equation (2.8) is inhomogeneous and allows no arbitrary physical parameter at all. Furthermore, since product of spherical harmonics are no longer spherical harmonics in general, Fourier series techniques employed by [BFP22] [BD+18] [IP19] that works for the torus are never valid for ; for example, we cannot simply assume in (2.8), as already illustrated by the solutions (5,5,8) (10,10,16). These are the crucial differences between the capillary spherical water waves and all known results for water waves in the flat setting.
2.5. Heuristics for Lifespan Estimate
To summarize, almost global lifespan estimate of (2.3) depends on the difficult number theoretic question 3. Before it is fully resolved, we can only expect partial results regarding the normal form transformation.
If there are only finitely many solutions to the Diophantine equation (2.9), then under the normal form reduction with coefficients given by (2.7), the equation (2.3) is transformed into the following system:
where is the orthogonal projection to , with being either 0 or 1, or exhausting the third component of all nontrivial solutions of (2.9).
There is no reasonable assertion to be made if the conjecture fails. However, if the conjecture does hold true, then we can expect that the lifespan estimate for -Cauchy data of (EQ) goes beyond , as what we expect for gravity water waves in the periodic setting, e.g. in [IP19]:
Conjecture 2.3.
Let’s explain the heuristic as follows. The argument we aim to implement is the standard “continuous induction method”, i.e. for some suitably large and and suitable , assuming and , we try to prove a better bound . Here is the standard metric on . It is intuitive to expect such a result for the cubic equation . As for the quadratic equation , it is crucial to implement the conservation of energy for a solution. We summarize it as
Proposition 2.5.
Fix . Let be a smooth solution of (2.3) and be as above. Suppose for some suitably large and , there holds
with sufficiently small. Then there is in fact a better bound for the low frequency part :
Proof.
We consider the “approximate” energy functional
where is the standard area measure on and is the standard connection on . This is nothing but the quadratic approximation to in (1.8) at , so there holds . Using volume preservation (1.6), we obtain , so summarizing we have
(2.10) |
Note that we used the conservation law . By spectral calculus on , we have
and by volume preservation and conservation of momentum (1.7) we find
(2.11) |
We point out that the above proof is independent from the magnitude of the lifespan , so it is always applicable as long as the cubic equation is well-understood. There are two crucial points in the proof of Proposition 2.5: the conservation of energy, and that the projection is of finite rank, so that a Bernstein type inequality holds. The last fact holds only if there are finitely many 3-way resonances, i.e. there are only finitely many solutions to the Diophantine equation (2.8).
Finally, we propose an even more ambitious conjecture concerning global dynamical properties of spherical water droplets, which is again illuminated by observation in hydrodynamical experiments under zero gravity, and also suggested by the results of Berti-Montalto [BM20]:
Appendix A MAGMA Code
MAGMA is a large, well-supported software package designed for computations in algebra, number theory, algebraic geometry, and algebraic combinatorics. In this appendix, we give the MAGMA code used to conduct computations on Diophantine equations related to the spherical capillary water waves system.
A.1. Integral Points on Elliptic Curve
We can find all integral points on a given elliptic curve over using MAGMA. For a monic cubic polynomial , the function EllipticCurve(f)
creates the elliptic curve
and the function IntegralPoints(E)
returns a sequence containing all the integral points on under the homogeneous coordinate of , modulo negation. We use this to find out all integral points on the elliptic curve
for natural number . The MAGMA code is listed below, which excludes all the ’s such that there are only trivial integral points on .
¿ Qx¡x¿ := PolynomialRing(Rationals()); ¿ for c in [1..50] do ¿ E := EllipticCurve(x^3+c*x^2-2*c^2*x); ¿ S, reps := IntegralPoints(E); ¿ if # S gt 3 then ¿ print c, E; ¿ print S; ¿ end if; ¿ end for; 2 Elliptic Curve defined by y^2 = x^3 + 2*x^2 - 8*x over Rational Field [ (-4 : 0 : 1), (-2 : 4 : 1), (-1 : -3 : 1), (0 : 0 : 1), (2 : 0 : 1), (4 : 8 : 1), (8 : -24 : 1), (50 : 360 : 1) ] 8 Elliptic Curve defined by y^2 = x^3 + 8*x^2 - 128*x over Rational Field [ (-16 : 0 : 1), (-8 : 32 : 1), (-4 : -24 : 1), (0 : 0 : 1), (8 : 0 : 1), (9 : -15 : 1), (16 : 64 : 1), (32 : -192 : 1), (200 : 2880 : 1) ] 13 Elliptic Curve defined by y^2 = x^3 + 13*x^2 - 338*x over Rational Field [ (-26 : 0 : 1), (0 : 0 : 1), (13 : 0 : 1), (121 : 1386 : 1) ] 15 Elliptic Curve defined by y^2 = x^3 + 15*x^2 - 450*x over Rational Field [ (-30 : 0 : 1), (-5 : -50 : 1), (0 : 0 : 1), (15 : 0 : 1), (24 : 108 : 1), (90 : -900 : 1) ] 17 Elliptic Curve defined by y^2 = x^3 + 17*x^2 - 578*x over Rational Field [ (-34 : 0 : 1), (-32 : -56 : 1), (0 : 0 : 1), (17 : 0 : 1), (833 : 24276 : 1) ] 18 Elliptic Curve defined by y^2 = x^3 + 18*x^2 - 648*x over Rational Field [ (-36 : 0 : 1), (-32 : -80 : 1), (-18 : 108 : 1), (-9 : -81 : 1), (0 : 0 : 1), (18 : 0 : 1), (36 : 216 : 1), (72 : -648 : 1), (450 : 9720 : 1) ] 22 Elliptic Curve defined by y^2 = x^3 + 22*x^2 - 968*x over Rational Field [ (-44 : 0 : 1), (-32 : -144 : 1), (0 : 0 : 1), (22 : 0 : 1), (198 : 2904 : 1) ] 23 Elliptic Curve defined by y^2 = x^3 + 23*x^2 - 1058*x over Rational Field [ (-46 : 0 : 1), (0 : 0 : 1), (23 : 0 : 1), (50 : -360 : 1) ] 26 Elliptic Curve defined by y^2 = x^3 + 26*x^2 - 1352*x over Rational Field [ (-52 : 0 : 1), (-49 : -105 : 1), (0 : 0 : 1), (26 : 0 : 1), (1300 : 47320 : 1) ] 30 Elliptic Curve defined by y^2 = x^3 + 30*x^2 - 1800*x over Rational Field [ (-60 : 0 : 1), (-50 : 200 : 1), (-45 : -225 : 1), (-24 : -216 : 1), (-20 : 200 : 1), (-6 : 108 : 1), (0 : 0 : 1), (30 : 0 : 1), (36 : 144 : 1), (40 : -200 : 1), (75 : -675 : 1), (90 : 900 : 1), (300 : 5400 : 1), (324 : -6048 : 1), (480 : -10800 : 1), (7290 : 623700 : 1), (10830 : -1128600 : 1), (226875 : 108070875 : 1) ] 32 Elliptic Curve defined by y^2 = x^3 + 32*x^2 - 2048*x over Rational Field [ (-64 : 0 : 1), (-32 : 256 : 1), (-16 : -192 : 1), (0 : 0 : 1), (32 : 0 : 1), (36 : -120 : 1), (64 : 512 : 1), (128 : -1536 : 1), (800 : 23040 : 1) ] 33 Elliptic Curve defined by y^2 = x^3 + 33*x^2 - 2178*x over Rational Field [ (-66 : 0 : 1), (0 : 0 : 1), (33 : 0 : 1), (81 : -756 : 1) ] 35 Elliptic Curve defined by y^2 = x^3 + 35*x^2 - 2450*x over Rational Field [ (-70 : 0 : 1), (-49 : 294 : 1), (-45 : -300 : 1), (-40 : 300 : 1), (-14 : -196 : 1), (0 : 0 : 1), (35 : 0 : 1), (50 : 300 : 1), (175 : -2450 : 1), (224 : 3528 : 1), (280 : -4900 : 1), (4410 : -294000 : 1), (14450 : -1739100 : 1) ] 39 Elliptic Curve defined by y^2 = x^3 + 39*x^2 - 3042*x over Rational Field [ (-78 : 0 : 1), (0 : 0 : 1), (39 : 0 : 1), (147 : -1890 : 1) ] 42 Elliptic Curve defined by y^2 = x^3 + 42*x^2 - 3528*x over Rational Field [ (-84 : 0 : 1), (-56 : -392 : 1), (-12 : 216 : 1), (0 : 0 : 1), (42 : 0 : 1), (63 : -441 : 1), (294 : 5292 : 1) ] 43 Elliptic Curve defined by y^2 = x^3 + 43*x^2 - 3698*x over Rational Field [ (-86 : 0 : 1), (-32 : -360 : 1), (0 : 0 : 1), (43 : 0 : 1) ] 46 Elliptic Curve defined by y^2 = x^3 + 46*x^2 - 4232*x over Rational Field [ (-92 : 0 : 1), (0 : 0 : 1), (46 : 0 : 1), (26496 : 4316640 : 1) ] 50 Elliptic Curve defined by y^2 = x^3 + 50*x^2 - 5000*x over Rational Field [ (-100 : 0 : 1), (-50 : 500 : 1), (-25 : -375 : 1), (-4 : 144 : 1), (0 : 0 : 1), (50 : 0 : 1), (100 : 1000 : 1), (200 : -3000 : 1), (1250 : 45000 : 1) ] Running Magma V2.27-7. Seed: 821911319; Total time: 2.430 seconds; Total memory usage: 85.16MB.
A.2. Classification of Projective Surface
Some basic geometric parameters of the complex projective surface
in can be computed using MAGMA. The author would like to thank Professor Bjorn Poonen for introducing MAGMA and providing the code listed below.
> Q:=Rationals(); > P<x,y,z,w>:=ProjectiveSpace(Q,3); > fx:=x*(x-w)*(x+2*w); > fy:=y*(y-w)*(y+2*w); > fz:=z*(z-w)*(z+2*w); > V:=Surface(P,(fz-fx-fy)^2-4*fx*fy); > KodairaEnriquesType(V); 2 0 General type Running Magma V2.27-7. Seed: 1338492807; Total time: 0.670 seconds; Total memory usage: 32.09MB.
Note that the Kodaira dimension is invariant regardless of the choice of base field, so it is legitimate to choose the base field to be in the above code. The variable is used to homogenize the equation. The function KodairaEnriquesType(V)
returns three values for the given projective surface : the first is the Kodaira dimension, the second is irrelevant when the Kodaira dimension is not , 1 or 0, and the third is the Kodaira-Enriquez classification of the surface .
References
- [AB15] Thomas Alazard and Pietro Baldi. Gravity capillary standing water waves. Archive for Rational Mechanics and Analysis, 217:741–830, 2015.
- [ABZ11] Thomas Alazard, Nicolas Burq, and Claude Zuily. On the water-wave equations with surface tension. 2011.
- [AD13] Thomas Alazard and Jean-Marc Delort. Global solutions and asymptotic behavior for two dimensional gravity water waves. arXiv preprint arXiv:1305.4090, 2013.
- [AM09] Thomas Alazard and Guy Métivier. Paralinearization of the dirichlet to neumann operator, and regularity of three-dimensional water waves. Communications in Partial Differential Equations, 34(12):1632–1704, 2009.
- [BC12] Joao Lucas Barbosa and Manfredo do Carmo. Stability of hypersurfaces with constant mean curvature. In Manfredo P. do Carmo–Selected Papers, pages 221–235. Springer, 2012.
- [BD+18] Massimiliano Berti, Jean-Marc Delort, et al. Almost global solutions of capillary-gravity water waves equations on the circle. Springer, 2018.
- [BFP22] Massimiliano Berti, Roberto Feola, and Fabio Pusateri. Birkhoff normal form and long time existence for periodic gravity water waves. Communications on Pure and Applied Mathematics, 2022.
- [BG98] Klaus Beyer and Matthias Günther. On the cauchy problem for a capillary drop. part i: irrotational motion. Mathematical methods in the applied sciences, 21(12):1149–1183, 1998.
- [BGT04] Nicolas Burq, Pierre Gérard, and Nikolay Tzvetkov. Strichartz inequalities and the nonlinear schrödinger equation on compact manifolds. American Journal of Mathematics, 126(3):569–605, 2004.
- [BM20] Massimiliano Berti and Riccardo Montalto. Quasi-periodic standing wave solutions of gravity-capillary water waves, volume 263. American mathematical society, 2020.
- [BS21] Andrew R Booker and Andrew V Sutherland. On a question of mordell. Proceedings of the National Academy of Sciences, 118(11):e2022377118, 2021.
- [CHS10] Hans Christianson, Vera Mikyoung Hur, and Gigliola Staffilani. Strichartz estimates for the water-wave problem with surface tension. Communications in Partial Differential Equations, 35(12):2195–2252, 2010.
- [CS07] Daniel Coutand and Steve Shkoller. Well-posedness of the free-surface incompressible euler equations with or without surface tension. Journal of the American Mathematical Society, 20(3):829–930, 2007.
- [DIPP17] Y Deng, A. D. Ionescu, B. Pausader, and F. Pusateri. Global solutions of the gravity-capillary water-wave system in three dimensions. Acta Mathematica, 219(2):213–402, 2017.
- [DS04] J-M Delort and Jeremie Szeftel. Long-time existence for small data nonlinear klein-gordon equations on tori and spheres. International Mathematics Research Notices, 2004(37):1897–1966, 2004.
- [GMS12] Pierre Germain, Nader Masmoudi, and Jalal Shatah. Global solutions for the gravity water waves equation in dimension 3. Annals of Mathematics, pages 691–754, 2012.
- [GMS15] Pierre Germain, Nader Masmoudi, and Jalal Shatah. Global solutions for capillary waves equation in dimension 3. Comm. Pure Appl. Math, 68(4):625–687, 2015.
- [HIT16] John K Hunter, Mihaela Ifrim, and Daniel Tataru. Two dimensional water waves in holomorphic coordinates. Communications in Mathematical Physics, 346(2):483–552, 2016.
- [IP15] Alexandru D Ionescu and Fabio Pusateri. Global solutions for the gravity water waves system in 2d. Inventiones mathematicae, 199:653–804, 2015.
- [IP19] AD Ionescu and F Pusateri. Long-time existence for multi-dimensional periodic water waves. Geometric and Functional Analysis, 29(3):811–870, 2019.
- [IT14] Mihaela Ifrim and Daniel Tataru. Two dimensional water waves in holomorphic coordinates ii: global solutions. arXiv preprint arXiv:1404.7583, 2014.
- [Lan05] David Lannes. Well-posedness of the water-waves equations. Journal of the American Mathematical Society, 18(3):605–654, 2005.
- [MP19] Rolando Magnanini and Giorgio Poggesi. On the stability for alexandrov’s soap bubble theorem. Journal d’Analyse Mathématique, 139(1):179–205, 2019.
- [MZ09] Mei Ming and Zhifei Zhang. Well-posedness of the water-wave problem with surface tension. Journal de mathématiques pures et appliquées, 92(5):429–455, 2009.
- [RT09] Michael Ruzhansky and Ville Turunen. Pseudo-differential operators and symmetries: background analysis and advanced topics, volume 2. Springer Science & Business Media, 2009.
- [Sch05] Ben Schweizer. On the three-dimensional euler equations with a free boundary subject to surface tension. In Annales de l’IHP Analyse non linéaire, volume 22, pages 753–781, 2005.
- [Sha85] Jalal Shatah. Normal forms and quadratic nonlinear klein-gordon equations. Communications on Pure and Applied Mathematics, 38(5):685–696, 1985.
- [Sog88] Christopher D Sogge. Concerning the lp norm of spectral clusters for second-order elliptic operators on compact manifolds. Journal of functional analysis, 77(1):123–138, 1988.
- [SZ08] Jalal Shatah and Chongchun Zeng. Geometry and a priori estimates for free boundary problems of the euler’s equation. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 61(5):698–744, 2008.
- [Wan20] X Wang. Global regularity for the 3d finite depth capillary water waves. Annales Scientifiques de l’École Normale Supérieure, 53(4):847–943, 2020.
- [Wu97] Sijue Wu. Well-posedness in sobolev spaces of the full water wave problem in 2-d. Inventiones mathematicae, 130(1):39–72, 1997.
- [Wu99] Sijue Wu. Well-posedness in sobolev spaces of the full water wave problem in 3-d. Journal of the American Mathematical Society, 12(2):445–495, 1999.
- [Wu09] Sijue Wu. Almost global wellposedness of the 2-d full water wave problem. Inventiones mathematicae, 177(1):45–135, 2009.
- [Wu11] Sijue Wu. Global wellposedness of the 3-d full water wave problem. Inventiones mathematicae, 184(1):125–220, 2011.
- [Zak68] Vladimir E Zakharov. Stability of periodic waves of finite amplitude on the surface of a deep fluid. Journal of Applied Mechanics and Technical Physics, 9(2):190–194, 1968.