Home Prescribing Gaussian and Geodesic Curvature on Disks
Article Open Access

Prescribing Gaussian and Geodesic Curvature on Disks

  • Sergio Cruz-Blázquez and David Ruiz EMAIL logo
Published/Copyright: July 20, 2018

Abstract

In this paper, we consider the problem of prescribing the Gaussian and geodesic curvature on a disk and its boundary, respectively, via a conformal change of the metric. This leads us to a Liouville-type equation with a non-linear Neumann boundary condition. We address the question of existence by setting the problem in a variational framework which seems to be completely new in the literature. We are able to find minimizers under symmetry assumptions.

MSC 2010: 35J20; 35R01; 53A30

1 Introduction

The problem of prescribing the Gaussian curvature on a compact surface Σ under a conformal change of the metric is a classical one, and dates back to [4, 21, 22]. Let us denote by g the original metric, by g~ the new one and by eu the conformal factor (that is, g~=eug). This problem reduces to solving the problem

- Δ g u + 2 K g = 2 K g ~ e u ,

where Kg and Kg~ denote the curvatures with respect to g and g~, respectively. The solvability of this equation has been studied for a long time, and it is not possible to give here a comprehensive list of references.

If Σ has a boundary, then boundary conditions are in order. Homogeneous Dirichlet and Neumann boundary conditions have already been considered in the literature. In this paper, our aim is to prescribe not only the Gaussian curvature in Σ, but also the geodesic curvature on Σ. In this case, we are led with the boundary value problem

(1.1) { - Δ g u + 2 K g = 2 K g ~ e u in  Σ , u n + 2 h g = 2 h g ~ e u / 2 on  Σ ,

where hg and hg~ are the geodesic curvatures of Σ relative to g and g~, respectively.

Some versions of this problem have been studied in the literature. The case hg~=0 has been treated by Chang and Yang in [7]. Moreover, the case Kg~=0 has been treated in [6, 23, 25]. There is also some progress in the blow-up analysis, see [3, 10], although a complete description of the phenomenon is still missing.

The case of constants Kg~, hg~ has also been considered. For instance, Brendle [5] uses a parabolic flow to show that this problem admits always a solution for some constant curvatures. By using complex analysis techniques, explicit expressions for the solutions and the exact values of the constants are determined if Σ is a disk or an annulus; see [19, 20]. The case of the half-plane has also been studied; see [24, 15, 33]. However, the case in which both curvatures are not constant has not been much considered. In [9], some partial existence results are given, but they include a Lagrange multiplier which is out of control. Moreover, a Kazdan–Warner type of obstruction to existence has been found in [16]. In a forthcoming work, the case of K<0 in domains different from the disk is treated, and also a blow-up analysis is performed; see [26]. At present, as far as we know, those are the only works considering non-constant curvatures.

The higher-dimensional analogue of this question (that is, prescribing scalar curvature of a manifold and mean curvature of the boundary) has been studied more. The case of zero scalar curvature and constant mean curvature is known as the Escobar problem, in strong analogy with the Yamabe problem. In this regard, see [1, 11, 12, 13, 14, 17, 18, 28] and the references therein.

By integrating (1.1) and applying the Gauss–Bonnet theorem, one obtains

(1.2) Σ K g ~ e u + Σ h g ~ e u / 2 = 2 π χ ( Σ ) .

In this paper, we shall consider the case in which χ(Σ)=1. By the Uniformization Theorem, we can pass via a conformal map to a disk, obtaining Kg~=0, hg~=1. Taking this into account, we can consider the problem

(1.3) { - Δ u = 2 K e u in  𝔻 2 , u η + 2 = 2 h e u / 2 on  𝕊 1 ,

where now K, h are the curvatures to be prescribed.

Generally speaking, the case of a disk is especially challenging because of the non-compact action of the group of conformal maps of the disk, as happens in the Nirenberg problem for Σ=𝕊2. This issue has been only treated in [6] for K=0 (see also [10]). A blow-up analysis in this case for non-constant K, h is yet to be done, and will be the target of further research. In this paper, as a first step in the understanding of the problem, we shall impose symmetry conditions on K, h in order to rule out this phenomenon. This idea goes back to Moser [30] for the Nirenberg problem.

Being more specific, we fix a symmetry group as follows:

  1. We denote by G one of the following groups of symmetries of the disk:

    1. The dihedric group 𝔻k with k3.

    2. The group of rotations with minimal angle 2πk, k2.

    3. The whole group of symmetries O(2).

Notice that none of the groups above has fixed points on 𝕊1, that is, for each x𝕊1 there exists gG such that g(x)x. We say that a function f is G-symmetric if f(x)=f(g(x)) for all gG and for all x in the domain of f.

Our main results are the following theorems.

Theorem 1.1.

Let G be as in a, and let K:D2R, h:S1R be G-symmetric, Hölder continuous and nonnegative functions, not both of them identically equal to 0. Then problem (1.3) admits a solution.

We can also deal with changing sign curvatures K, h as long as their negative part is small.

Theorem 1.2.

Let G be as in a, and let K0:D2R, h0:S1R be G-symmetric, Hölder continuous and nonnegative functions, none of them identically equal to 0. Then there exists ε>0 such that problem (1.3) admits a solution for any Hölder continuous and G-symmetric functions K, h with K-K0L+h-h0L<ε.

One of the main goals of this paper is to find an original variational setting to this problem, which we think is natural and could be of use in future research on the topic. Let us be more specific. We define the parameter

ρ := 𝔻 2 K e u = 2 π - 𝕊 1 h e u / 2 .

In order to fix the ideas, let us assume that both K, h are nonnegative functions; by (1.2), 0<ρ<2π.

We shall show that (1.3) is equivalent to

(1.4) { - Δ u = 2 ρ K e u 𝔻 2 K e u in  𝔻 2 , u η + 2 = 2 ( 2 π - ρ ) h e u / 2 𝕊 1 h e u / 2 on  𝕊 1 , ( 2 π - ρ ) 2 ρ = ( 𝕊 1 h e u / 2 ) 2 𝔻 2 K e u for  0 < ρ < 2 π .

Observe that problem (1.4) is now invariant under addition of constants to u, and ρ is unknown here. This formulation may seem rather artificial but it has the advantage of being related to the critical points of the energy functional

I ( u , ρ ) = 1 2 𝔻 2 | u | 2 - 2 ρ log 𝔻 2 K e u + 2 𝕊 1 u - 4 ( 2 π - ρ ) log 𝕊 1 h e u / 2
(1.5) + 4 ( 2 π - ρ ) log ( 2 π - ρ ) + 2 ρ + 2 ρ log ρ .

We highlight the fact that the functional above depends on the couple (u,ρ), where uH1(𝔻2) and ρ(0,2π). The form of this energy functional seems to be completely new in the related literature.

If we freeze the variable ρ, the form of this functional is adequate for the use of Moser–Trudinger-type inequalities (or Onofri-type inequalities) which are already available also for boundary terms. Indeed, by interpolating these inequalities we will show that I is bounded from below. We will gain coercivity in the u variable by imposing symmetry, as done first by Moser in [30]. Finally, we will need to exclude the possibility of obtaining minima at the endpoints ρ=0 or ρ=2π. Those limit cases correspond to the problem in which K=0 or h=0, respectively, so some study of these cases is needed. By energy estimates we can assure that the minimum is attained at ρ(0,2π), concluding the proof.

If either K or h changes sign, the above approach fails. We shall also give in Theorem 3.4 a more general result; as a corollary, and making use of a compactness result for minima of the functional I, we will obtain the perturbation result stated in Theorem 1.2.

The rest of the paper is organized as follows: In Section 2, we set the notation and the variational formulation of the problem. After that, an analysis of the properties of the energy functional is performed by means of Moser–Trudinger-type inequalities. Section 3 is devoted to the proof of Theorem 1.1, for which we first need to address the limiting cases ρ=0 and ρ=2π. A more general version is also given. Finally, the proof of Theorem 1.2 is completed in Section 4.

2 Variational Setting

2.1 Notation

Let us first set some notation. Given a set AX in a metric space, we set

( A ) r = { x X : dist ( x , A ) < r } .

Regarding the integrals, in this paper we shall consider only the Lebesgue measure and we drop the element of area or length, that is, we shall only write 𝔻2Keu or 𝕊1heu/2. We also use the symbol f to denote the mean value of f, that is,

Σ f = 1 | Σ | Σ f .

In our estimates, we sometimes write C to denote a positive constant, independent of the variables considered, that may change from line to line.

2.2 Variational Formulation

As commented in the introduction, we will consider the functional I given by (1.5) and defined on the space

𝕏 × ( 0 , 2 π ) = { u H 1 ( 𝔻 2 ) : 𝔻 2 K e u > 0 , 𝕊 1 h e u / 2 > 0 } × ( 0 , 2 π ) .

With the purpose of clarifying the notation, for a fixed ρ(0,2π) we denote by Iρ the functional uI(u,ρ) defined for every u𝕏. We should notice that the functionals Iρ are invariant under the addition of constants.

Lemma 2.1.

The space X is nonempty if and only if K and h are positive somewhere.

Proof.

We reduce ourselves to prove that if K and h are positive somewhere, then 𝕏 is nonempty as the reciprocal is immediate. As K is continuous and there exists x0Int(𝔻2) such that K(x0)>0, there exists r>0 such that ({x0})r𝕊1= and K(x)>0 for all x({x0})r.

Moreover, we know that there exists x1𝕊1 satisfying h(x1)>0, and again by continuity we get s>0 such that h(x)>0 for all x({x1})s𝕊1. It is not restrictive to assume ({x0})r({x1})s=. We define Ω0r:=({x0})r and Ω1s:=({x1})s and consider a cutoff function φH1(𝔻2) satisfying

φ ( x ) = { a if  x Ω 0 r / 2 , b if  x Ω 1 s / 2 , 0 if  𝔻 2 ( Ω 0 r Ω 1 s ) ,

where a and b are real constants to determine. We see that

𝕊 1 h e φ / 2 = Ω 1 s / 2 𝔻 2 h e φ / 2 + ( Ω 1 s Ω 1 s / 2 ) 𝔻 2 h e φ / 2 + 𝔻 2 Ω 1 s h e φ / 2
e b / 2 Ω 1 s / 2 𝔻 2 h + 𝔻 2 Ω 1 s / 2 h
= C 1 e b / 2 + C ,

where C1>0 and C. We can choose b large enough so that

𝕊 1 h e φ / 2 > 0 .

Furthermore,

𝔻 2 K e φ = Ω 0 r / 2 K e φ + Ω 1 s / 2 K e φ + 𝔻 2 ( Ω 0 r Ω 1 r ) K e φ + Ω 1 s Ω 1 s / 2 K e φ + Ω 0 r Ω 0 r / 2 K e φ
e a Ω 0 r / 2 K - e b C 2 K + C
= C 1 e a - C 2 e b + C .

So we can also set a big enough so that

𝔻 2 K e φ > 0 .

Let us point out that the Euler–Lagrange equation of I is given by (1.4), which is a reformulation of (1.3) in view of the next lemma.

Lemma 2.2.

Problems (1.3) and (1.4) are equivalent.

Proof.

In order to check that every solution of (1.3) is a solution of (1.4) we just need to take

ρ = 𝔻 2 K e u = 2 π - 𝕊 1 h e u / 2 > 0 .

Reciprocally, if u𝕏 solves (1.4), applying the invariance under addition of constants of that problem, we have, for any C,

- Δ ( u + C ) = 2 ρ K e u + C e C 𝔻 2 K e u ,
( u + C ) η + 2 = 2 ( 2 π - ρ ) h e u / 2 e C 2 𝕊 1 h e u / 2 .

If we want u+C to solve (1.3), we need C such that

e C = ρ 𝔻 2 K e u , e C 2 = ( 2 π - ρ ) 𝕊 1 h e u / 2 .

The third equation of (1.4) tells us that both conditions are actually the same. Thus, it is enough to choose

C = log ρ - log 𝔻 2 K e u .

2.3 Moser–Trudinger Inequalities

The Moser–Trudinger inequality (see [7, 29, 30, 32]) and their variations are useful tools to deal with the non-linear terms of exponential type which appear in our functional. In particular, we are interested in weaker versions of these inequalities, also called Onofri-type inequalities.

Theorem 2.3.

Let Σ be a compact surface with C1 boundary. Then there exists a constant CR, depending only on Σ, such that

(2.1) log Σ e u 1 16 π Σ | u | 2 + C for all  u H 0 1 ( Σ ) ,

and

(2.2) log Σ e u 1 8 π Σ | u | 2 + Σ u + C for all  u H 1 ( Σ ) .

The first inequality is classical, whereas the second is given in [7, Proposition 2.3 and the subsequent corollary]. In both cases the constant is optimal.

In order to address the non-linear boundary terms of the functional I, we will use an analogous version of Theorem 2.3 for the boundary of a compact surface that can be found for instance in [23].

Proposition 2.4.

Let Σ be a compact surface with C1 boundary. Then there exists a constant C>0, depending only on Σ, such that

log Σ e u 1 4 π Σ | u | 2 + Σ u + C for all  u H 1 ( Σ ) .

In the case of the disk, the above inequality is the so-called Lebedev–Milin inequality (with C=0, see for instance [31, (4’)]).

By interpolating the previous inequalities, we will obtain a lower bound for the functional I. First, we notice that inequality (2.2) can be manipulated so that the mean value of u in Σ replaces the mean in Σ.

Corollary 2.5.

Let Σ be a compact surface with C1 boundary. There exists a constant CR, depending only on Σ, such that

log Σ e v 1 8 π Σ | v | 2 + Σ v + C for all  v H 1 ( Σ ) .

Proof.

We consider the problem

(2.3) { - Δ w = - 4 π | Σ | in  Σ , w η = 4 π | Σ | on  Σ .

Let us point out that (2.3) is solvable in H1(Σ) because

Σ 4 π | Σ | = - Σ 4 π | Σ | = 4 π .

We fix a solution w of (2.3) and apply (2.2) to v+w, obtaining

log Σ e v 1 8 π Σ | v | 2 + 1 4 π Σ w η v - 1 4 π Σ ( Δ w ) v + Σ v + C .

Finally, we use that w solves (2.3):

log Σ e v 1 8 π Σ | v | 2 + Σ v - Σ v + Σ v + C = 1 8 π Σ | v | 2 + Σ v + C .

In a similar way, one can obtain a modified version of Proposition 2.4 in which the mean value of u on Σ substitutes the mean on Σ.

Corollary 2.6.

Let Σ be a compact surface with C1 boundary. There exists CR, depending only on Σ, such that

log Σ e u 1 4 π Σ | u | 2 + Σ u + C for all  u H 1 ( Σ ) .

The combined use of inequality (2.3) and Corollary 2.5 allows us to prove that I is bounded from below in H1(𝔻2).

Proposition 2.7.

There exists a constant CR such that Iρ(u)C for every uX and every ρ[0,2π].

Proof.

Let us define f:(0,2π) as the correction term in (1.5), that is,

f ( ρ ) = 4 ( 2 π - ρ ) log ( 2 π - ρ ) + 2 ρ + 2 ρ log ρ .

It is clear that

lim ρ 0 f ( ρ ) = 8 π log ( 2 π ) , lim ρ 2 π f ( ρ ) = 4 π + 4 π log ( 2 π ) .

Then f can be continuously extended to the compact interval [0,2π]. Thus, there exists a constant M>0 such that |f(ρ)|M for all ρ[0,2π]. Moreover, since K and h are continuous, there exist M1,M2 such that

log 𝔻 2 K e u log 𝔻 2 e u + C , log 𝕊 1 h e u / 2 𝕊 1 e u / 2 + C .

Then for every a,b,

I ρ ( u ) 1 2 𝔻 2 | u | 2 - 2 ρ log 𝔻 2 e u - 4 ( 2 π - ρ ) log 𝕊 1 e u / 2 + 2 𝕊 1 u + C
= 8 π - 2 a - b 16 π 𝔻 2 | u | 2 + a 8 π 𝔻 2 | u | 2 + b 16 π 𝔻 2 | u | 2 - 2 ρ log 𝔻 2 e u - 4 ( 2 π - ρ ) log 𝕊 1 e u / 2 + 2 𝕊 1 u + C .

As the functional I is invariant under the addition of constants, we can assume that Σu=0 and apply Corollary 2.5 and Proposition 2.4, taking a=2ρ and b=4(2π-ρ), and thus obtaining:

I ρ ( u ) - 2 ρ 𝕊 1 u - 2 ( 2 π - ρ ) 𝕊 1 u + 2 𝕊 1 u + C = C .

We highlight that the constant C does not depend on ρ. ∎

Proposition 2.7 states that the functional I is bounded from above, but we do not have coercivity. The reason for this is the non-compact action of the conformal group of the disk. This effect appears, for instance, also in the Nirenberg problem in the sphere and makes the problem rather difficult.

We will show now that we can gain coercivity by restricting ourselves to spaces of symmetric functions. In order to do that, we introduce local versions of the inequalities above. This idea dates back to [2], and these inequalities are known as Chen–Li-type inequalities (see [8] for more details).

Proposition 2.8.

Assume Σ to be a compact surface with C1 boundary, and let Σ1Σ and δ>0 be such that (Σ1)δΣ=. Then for every ε>0 there exists a constant CR depending on ε and δ such that

16 π log Σ 1 e u ( Σ 1 ) δ | u | 2 + ε Σ | u | 2 + C for all  u H 1 ( Σ ) with  Σ u = 0 .

The details of the proof of this precise statement can be found, for instance, in [27, Proposition 2.2], but the idea dates back to [8]. Roughly speaking, one applies (2.1) to the function u multiplied by a cut-off function in Σ1.

If the function u has mass in several separated regions satisfying the hypothesis of the propositions above, the obtained bounds improve by a factor of the number of such regions. This information is collected in the following corollary (see for instance [27, Lemma 2.4] for the case l=2; the case of general l is analogous).

Corollary 2.9.

Let Σ be a compact surface with C1 boundary, let lN and let Σ1,,ΣlΣ for which there exists a δ>0 such that (Σi)δ(Σj)δ= if ij. Assume that there exists γ(0,1l) such that

Σ i e u Σ e u γ for all  i = 1 , , l .

Then for every ε>0 there exists a constant CR depending on ε, δ and γ such that

8 l π log Σ e u Σ | u | 2 + ε Σ | u | 2 + C for all  u H 1 ( Σ ) with  Σ u = 0 .

Using the same techniques, we can give a localized version of Proposition 2.4.

Proposition 2.10.

Let Σ be a compact surface with C1 boundary, and let Γ1Σ. Then for every ε,δ>0 there exists a constant CR depending on ε and δ such that

4 π log Γ 1 e u ( Γ 1 ) δ | u | 2 + ε Σ | u | 2 + C for all  u H 1 ( Σ ) with  Σ u = 0

Proof.

Following [8], we consider a cutoff function gδ:Σ[0,1] satisfying

g δ = { 1 if  x Γ 1 , 0 if  x Σ ( Γ 1 ) δ / 2 .

We have gδuH1(Σ), hence we can apply Corollary 2.6:

(2.4) 4 π log Γ 1 e u = 4 π log Γ 1 e g δ u 4 π log Σ e g δ u Σ | ( g δ u ) | 2 + 4 π Σ g δ u + C .

Then

Σ | ( g δ u ) | 2 = Σ u 2 | g δ | 2 + 2 Σ g δ u u , g δ + Σ ( g δ ) 2 | u | 2
(2.5) C δ Σ u 2 + 2 Σ g δ u | u | | g δ | + ( Γ 1 ) δ | u | 2 .

The central term can be bounded using Cauchy’s inequality, and thus obtaining

(2.6) Σ g δ u | u | | g δ | C δ Σ u | u | C ε , δ Σ u 2 + ε Σ | u | 2 .

Combining (2.5) and (2.6), we obtain

(2.7) Σ | ( g δ u ) | 2 ( Γ 1 ) δ | u | 2 + ε Σ | u | 2 + C ε , δ Σ u 2 .

Also, we have the following bound for the mean value of gδu on Σ:

(2.8) Σ g δ u Σ 1 2 ( ( g δ ) 2 + u 2 ) 1 2 Σ ( g δ ) 2 + 1 2 | Σ | Σ u 2 C δ + C Σ u 2 .

Now, apply both inequalities (2.7) and (2.8) to (2.4) to get

(2.9) 4 π log Γ 1 e u ( Γ 1 ) δ | u | 2 + ε Σ | u | 2 + C ε , δ Σ u 2 + C .

Finally, we address the term Σu2.

Let a, η=|{xΣ:u(x)a}| and (u-a)+=max{0,u-a}. Clearly, u(u-a)++a. We now apply formula (2.9) to the function (u-a)+:

4 π log Γ 1 e u 4 π log ( e a Γ 1 e ( u - a ) + )
4 π a + log Γ 1 e ( u - a ) +
4 π a + ( Γ 1 ) δ | ( u - a ) + | 2 + ε Σ | ( u - a ) + | 2 + C ε , δ Σ ( ( u - a ) + ) 2
(2.10) 4 π a + ( Γ 1 ) δ | u | 2 + ε Σ | u | 2 + C ε , δ Σ ( ( u - a ) + ) 2 .

By means of the Sobolev, Hölder and Poincaré–Wirtinger inequalities,

Σ ( ( u - a ) + ) 2 = { x Σ : u ( x ) a } ( ( u - a ) + ) 2
η 1 / 2 ( Σ ( ( u - a ) + ) 4 ) 1 / 2
η 1 / 2 ( u - a ) + H 1 ( Σ ) 2
(2.11) C η 1 / 2 Σ | u | 2 .

Again by the Poincaré–Wirtinger inequality,

(2.12) a η { x Σ : u ( x ) a } u Σ | u | C ( Σ | u | 2 ) 1 / 2 C ( Σ | u | 2 ) 1 / 2 .

From (2.12), using Cauchy’s inequality, we obtain

(2.13) a θ Σ | u | 2 + C 2 η 2 θ for all  θ > 0 .

Mixing (2.10), (2.11) and (2.13), we obtain

4 π log Γ 1 e u 4 π θ Σ | u | 2 + ( Γ 1 ) δ | u | 2 + ε Σ | u | 2 + C ε , δ η 1 / 2 Σ | u | 2 + C ,

and it is enough to take θ=14π and η1/2εCε,δ to conclude the proof. ∎

Corollary 2.11.

Let Σ be a compact surface with C1 boundary, let lN and let Γ1,,ΓlΣ for which there exists a δ>0 such that (Γi)δ(Γj)δ= if ij. Moreover, assume that there exists γ(0,1l) such that

(2.14) Γ i e u Σ e u γ for all  i = 1 , , l .

Then for every ε>0 there exists a constant CR depending on ε,δ and γ such that

4 l π log Σ e u Σ | u | 2 + ε Σ | u | 2 + C for all  u H 1 ( Σ ) with  Σ u = 0 .

Proof.

First, we apply to each Γi the previous result, obtaining

4 π log Γ i e u ( Γ i ) δ | u | 2 + ε Σ | u | 2 + C .

Using (2.14), we get

4 π log Γ i e u 4 π log Σ e u + C .

Then

4 π log Σ e u ( Γ i ) δ | u | 2 + ε Σ | u | 2 + C .

Finally, summing over i{1,,l}, we obtain

4 l π log Σ e u i ( Γ i ) δ | u | 2 + ε l Σ | u | 2 + C Σ | u | 2 + ε l Σ | u | 2 + C .

We have just seen that the more regions the mass of a function is separated in, the better bounds we obtain using the local versions of the Moser–Trudinger inequalities. If an H1(𝔻2) function is concentrated in an interior point of the disk, Proposition 2.8 gives us a lower bound which is sufficient to achieve coercivity, but that is not the case when a function concentrates around a boundary point. To avoid this we will restrict ourselves to consider functions satisfying a symmetry condition guaranteeing that a function cannot concentrate around a single point of the boundary. Hence we will obtain coercivity by interpolating Corollaries 2.9 and 2.11 with l=2.

Figure 1 
            If a symmetric function concentrates around x∈𝕊1{x\in{\mathbb{S}^{1}}},then it also concentrates around g⁢(x){g(x)} for all g∈G{g\in G}.
Figure 1

If a symmetric function concentrates around x𝕊1,then it also concentrates around g(x) for all gG.

We let G be a subgroup of the orthogonal transformation group of 𝔻2 such that the set of fixed points on 𝕊1 under the action of G is empty; in other words,

{ x 𝕊 1 : g ( x ) = x  for all  g G } = .

For instance, we can take G as the group of rotations generated by g(z)=e(2πi)/kz as well as the dihedral groups 𝔻k (k, k>1).

In the sequel, K and h will be assumed to be G-symmetric functions, and we set

H G 1 ( 𝔻 2 ) = { u H 1 ( 𝔻 2 ) : u g = u  for all  u G }

and

𝕏 G = { u 𝕏 : u g = u  for all  g G } .

As in Lemma 2.1, we observe that if K and h are G-symmetric functions somewhere positive, then 𝕏G is not empty.

Proposition 2.12.

Given ρ[0,2π], the functional Iρ is coercive on XG, that is,

I ρ ( u ) + ( u H 1 ( 𝔻 2 ) + , u 𝕏 G ) .

Proof.

Take a sequence (un) in 𝕏G. We know that Iρ is invariant under the addition of constants, so we can assume that 𝔻2un=0 for every n. We have

I ρ ( u n ) 1 2 𝔻 2 | u n | 2 - 2 ρ log 𝔻 2 e u n - 4 ( 2 π - ρ ) log 𝕊 1 e u n / 2 + 2 𝕊 1 u n + C .

Then for any a,b one has

I ρ ( u n ) 16 π - 2 a - b 32 π 𝔻 2 | u n | 2 + a 16 π 𝔻 2 | u n | 2 + b 32 π 𝔻 2 | u n | 2 + 2 𝕊 1 u n
- 2 ρ log 𝔻 2 e u n - 4 ( 2 π - ρ ) log 𝕊 1 e u n / 2 .

We can now apply Corollaries 2.9 and 2.11 with l=2 (see Figure 1):

I ρ ( u n ) 16 π - 2 a - b 32 π 𝔻 2 | u n | 2 + a log 𝔻 2 e u n - a ε 𝔻 2 | u n | 2 + b log 𝕊 1 e u n / 2
- b ε 𝔻 2 | u n | 2 - 2 ρ log 𝔻 2 e u n - 4 ( 2 π - ρ ) log 𝕊 1 e u n / 2 + 2 𝕊 1 u n + C .

Choosing a=2ρ and b=4(2π-ρ) and applying the trace inequality, we obtain

I ρ ( u n ) ( 1 4 - ε ) 𝔻 2 | u n | 2 - 2 C 2 u n H 1 ( 𝔻 2 ) + C , C 2 > 0 .

Finally, taking ε small enough and using the Poincaré–Wirtinger inequality, we obtain

I ρ ( u n ) C 1 u n H 1 ( 𝔻 2 ) 2 - C 2 u n H 1 ( 𝔻 2 ) + C , C 1 , C 2 > 0 .

Again, we remark that the constant C1 is independent of ρ. ∎

3 Proof of Theorem 1.1 and Its Generalization

We begin this section by considering the limiting cases ρ=0 and ρ=2π. These cases have their own interest, as will be shown, but their study will be useful also for the proofs of Theorems 1.1, 1.2 and 3.4.

Observe that

I ( u , 0 ) = 1 2 𝔻 2 | u | 2 + 2 𝕊 1 u - 8 π log 𝕊 1 h e u / 2 + 8 π log ( 2 π ) ,

and, as K does not play any role, it can be defined on the bigger space

𝕏 1 = { u H 1 ( 𝔻 2 ) : 𝕊 1 h e u / 2 > 0 } 𝕏 .

The critical points of I0 on 𝕏1 are weak solutions of the problem

{ - Δ u = 0 in  𝔻 2 , u η + 2 = 4 π h e u / 2 𝕊 1 h e u / 2 on  𝕊 1 ,

which is clearly equivalent to the problem of prescribing Gaussian curvature K=0 and geodesic curvature h, that is,

(3.1) { - Δ u = 0 in  𝔻 2 , u η + 2 = 2 h e u / 2 on  𝕊 1 .

Under the hypothesis that h is G-symmetric, we can seek a minimizer of I0 on the space of symmetric functions

𝕏 G 1 = { u 𝕏 1 : u g = u  for all  g G } .

Theorem 3.1.

Let G be as in a, and let h:S1R be a G-symmetric, Hölder continuous and somewhere positive function. Then problem (3.1) admits a solution as a minimum of I0 on XG1.

Proof.

The functional is bounded from below as seen in Proposition 2.7, so there exists

α = inf u 𝕏 G 1 I 0 ( u ) .

Let (un) be a minimizing sequence in 𝕏G1, that is, I0(un)α. By Proposition 2.12, we know that I0 is coercive, so un is bounded in the H1(𝔻2) norm and we can assume that there exists u0 in H1(𝔻2) such that, up to a subsequence, unu0. Then we also have

𝕊 1 u n 𝕊 1 u 0 , 𝕊 1 h e u n / 2 𝕊 1 h e u 0 / 2 .

Combining this information with the fact that the function u𝔻2|u|2 is weakly lower semicontinuous, we have I0(u0)α. It is easy to check that

𝕊 1 h e u 0 2 > 0

because if we had

𝕊 1 h e u n / 2 0 ,

then I0(un)+, which contradicts that un is minimizing. Also, notice that weak convergence respects symmetry, so u0 is a G-symmetric function. ∎

Analogously, we can consider the functional related to the limiting case ρ=2π:

I ( u , 2 π ) = 1 2 Σ | u | 2 + 2 𝕊 1 u - 4 π log 𝔻 2 K e u + 4 π + 4 π log ( 2 π )

defined on the space

𝕏 G 2 = { u H G 1 ( 𝔻 2 ) : 𝔻 2 K e u > 0 } 𝕏 G .

One can check that its variation with respect to u produces weak solutions of the problem

{ - Δ u = 4 π K e u 𝔻 2 K e u in  𝔻 2 , u η + 2 = 0 on  𝕊 1 ,

which is equivalent to the problem of prescribing geodesic curvature h=0 and Gaussian curvature K:

{ - Δ u = 2 K e u in  𝔻 2 , u η + 2 = 0 on  𝕊 1 .

A trivial adaptation of the proof of Theorem 3.1 gives the following result.

Theorem 3.2.

Let G be as in a, and let K:D2R be a G-symmetric, Hölder continuous and somewhere positive function. Then problem (3.1) admits a solution as a minimum of I2π on XG2.

Remark 3.3.

The existence result of Theorem 3.1 is known; see for instance [25]. We have not found an explicit statement of the existence result of Theorem 3.2, but we guess that it must be also known. However, in this section we have reinterpreted those solutions as minimizers of I0 and I2π, respectively. This will be of use in what follows.

Let us now conclude the proof of Theorem 1.1.

Proof of Theorem 1.1.

If K=0 or h=0, then we are under the assumptions of Theorems 3.1 or 3.2. Then we can assume that both K and h are positive in some point and nonnegative. In this case,

𝕏 G = 𝕏 G 1 = 𝕏 G 2 = H G 1 ( 𝔻 2 ) .

By Proposition 2.12, there exists a minimizer (u^,ρ^)HG1(𝔻2)×[0,2π] for I. We conclude the proof if we exclude the possibilities ρ^=0 or ρ^=2π.

Assume that ρ^=0. Observe that in this case u^ is a minimizer for I(,0). Then

I ( u ^ , 0 ) I ( u ^ , ρ )
= I ( u ^ , 0 ) - 2 ρ log ( 𝔻 2 K e u ^ ) + 4 ρ log ( 𝕊 1 h e u ^ / 2 )
+ 8 π log ( 2 π - ρ 2 π ) - 4 ρ log ( 2 π - ρ ) + 2 ρ + 2 ρ log ρ .

But observe that, as ρ0, the main term above is 2ρlogρ, which is negative. This gives a contradiction that excludes the case ρ^=0. One can exclude the case ρ^=2π in an analogous way. ∎

The proof of Theorem 1.1 can be adapted to a more general setting as follows.

Theorem 3.4.

Let G be as in a, and let K, h be G-symmetric, Hölder continuous functions that are positive somewhere. We define

S 0 = { u 𝕏 G 1 : I 0 ( u ) = min 𝕏 G 1 I 0 } , S 2 π = { u 𝕏 G 2 : I 2 π ( u ) = min 𝕏 G 2 I 2 π } .

If S0XG and S2πXG are nonempty, then (1.3) admits a solution.

Clearly, Theorem 1.1 is an immediate consequence of Theorem 3.4. Notice also that the sets S0 and S2π of the hypotheses are nonempty because of Theorems 3.1 and 3.2.

Proof.

The proof follows from the same energy comparison argument as above, but a couple of details are worth writing down. First, the existence of a minimizer is not clear a priori. Let (un,ρn)𝕏G×(0,2π) be a minimizing sequence, that is, I(un,ρn)infI. Clearly, un is bounded in HG1(𝔻2) by Proposition 2.12, but its weak limit u^ could fall outside 𝕏G.

If ρnρ^(0,2π), from the fact that I(un,ρn) is bounded we obtain

0 < ε < 𝔻 2 K e u n < C , 0 < ε < 𝕊 1 h e u n / 2 < C

for some ε>0 and C>0. As a consequence, unu^𝕏G and we are done.

Assume now that ρn0. If n is sufficiently large, we have the estimate

I ( u n , ρ n ) - 2 ρ n log ( 𝔻 2 K e u n ) - 4 ( 2 π - ρ n ) log ( 𝕊 1 h e u n / 2 ) + C .

Notice that

lim inf n - 2 ρ n log ( 𝔻 2 K e u n ) 0 .

Thus, -log(𝕊1heun/2) must be bounded from above, which means that

0 < ε < 𝕊 1 h e u n / 2 .

Now, we write

I ( u n , ρ n ) = I ( u n , 0 ) - 2 ρ n log ( 𝔻 2 K e u n ) + 4 ρ n log ( 𝕊 1 h e u n / 2 )
+ 8 π log ( 2 π - ρ n 2 π ) - 4 ρ n log ( 2 π - ρ n ) + 2 ρ n + 2 ρ n log ρ n .

From this we deduce that

inf I = lim n I ( u n , ρ n ) lim inf n I ( u n , 0 ) I ( u 0 , 0 ) ,

where u0S0𝕏G. But, as in the proof of Theorem 1.1,

I ( u 0 , 0 ) > I ( u 0 , ρ )

for small values of ρ. This contradiction shows that ρn cannot converge to 0. In an analogous way, we can exclude its convergence to 2π. ∎

4 A Perturbation Result

In this section, it is necessary to specify the dependence of I on the curvature functions K and h, so we are writing I(u,ρ)=I[K,h](u,ρ). We begin with a compactness result:

Lemma 4.1.

Let (Kn) and (hn) be sequences of Hölder continuous G-symmetric functions, defined on D2 and S1, respectively, such that

K n K uniformly in  𝔻 2 and  K C 0 , α ( 𝔻 2 ) ,
h n h uniformly on  𝕊 1 and  h C 0 , α ( 𝕊 1 ) .

Let us consider a sequence (un), where each un is a solution of the problem

(4.1) { - Δ u = 2 K n e u in  𝔻 2 , u n + 2 = 2 h n e u / 2 on  𝕊 1 ,

satisfying

(4.2) ρ n = 𝔻 2 K n e u n > 0 , 𝕊 1 h n e u n / 2 > 0    for all  n .

Assume that I[Kn,hn](un,ρn) is uniformly bounded from above. Then unu on H1(D2), with u being a solution of the problem

(4.3) { - Δ u = 2 K e u in  𝔻 2 , u n + 2 = 2 h e u / 2 on  𝕊 1 .

Proof.

First, we notice that Kn-K0 and hn-h0 imply that for every ε>0 there exists n0 such that, for nn0,

K n < K + ε , h n < h + ε .

Hypothesis (4.2) gives us 0<ρn<2π for all n. Then for nn0 we have the following bound:

I [ K n , h n ] ( u n , ρ n ) I ( K + ε , h + ε ) ( u n , ρ n ) .

And then, by Proposition 2.12, there exist constants C1,C2>0 independent of n such that

I [ K n , h n ] ( u n , ρ n ) C 1 u n H 1 2 - C 2 u n H 1 + C ε .

Taking into account the hypothesis that I[Kn,hn](un,ρn) is uniformly bounded from above, we have immediately that un is bounded in the H1(𝔻2) norm. Hence, up to a subsequence we can assume that there exists uH1(𝔻2) such that unu.

Then it is known that

2 K n e u n 2 K e u and 2 h n e u n / 2 2 h e u 2

on Lp for 1p<+, and that un,wu,w for all wH1(𝔻2). In particular,

u n | 𝕊 1 u | 𝕊 1 in  L 2 ( 𝕊 1 ) .

We now pass to the limit in the weak formulation of (4.1):

𝔻 2 u n , v - 2 𝔻 2 K n e u n v + 2 𝕊 1 v - 𝕊 1 h n e u n / 2 v = 0

for all vH1(𝔻2). As a consequence, u is a weak solution of (4.3). By standard regularity estimates, u is indeed a classical solution. ∎

The next step is to check that, when considering a sequence of minimum-type solutions, the hypotheses of Lemma 4.1 are automatically satisfied. Note that under our hypotheses we have I[Kn,hn](,)I[K,h](,) pointwise in 𝕏×(0,2π).

If (un,ρn) is a sequence of minimum-type solutions of (4.1), then

lim sup n + I [ K n , h n ] ( u n , ρ n ) = lim sup n + min 𝕏 × ( 0 , 2 π ) I [ K n , h n ] ( , ) min 𝕏 × ( 0 , 2 π ) I .

The previous inequality is due to the fact that (fn) converging pointwise to f implies

lim n + inf f n ( y ) inf f ( y ) .

Proof of Theorem 1.2.

We apply Theorem 3.4 to the problem

{ - Δ u = 2 K e u in  𝔻 2 , u η + 2 = 2 h e u / 2 on  𝕊 1 ,

for which we need that the limiting problems

(P1K) { - Δ u = 2 K e u in  𝔻 2 , u n + 2 = 0 on  𝕊 1 ,
(P2h) { - Δ u = 0 in  𝔻 2 , u η + 2 = 2 h e u / 2 on  𝕊 1 ,

admit minimum-type solutions u1 and u2, respectively, verifying

𝔻 2 K e u 2 > 0 , 𝕊 1 h e u 1 / 2 > 0 .

By contradiction, take Hölder continuous functions Kn and hn converging uniformly to K0 and h0. We can assume that n is large enough so that Kn and hn are somewhere positive, so that solutions for the limiting problems in the form of minimizers can be found via Theorems 3.1 and 3.2. Now, take a sequence of minimum-type solutions (u~n) of problems (PKn1) and a sequence of minimum-type solutions (u^n) of problems (Phn2) such that

(4.4) either 𝔻 2 K n e u ^ n 0 or 𝕊 1 h n e u ~ n / 2 0 for all  n .

By Lemma 4.1, we know that u^nu^ and u~nu~ are solutions for the limiting problems (PK01) and (Ph02), respectively. Taking the limit when n+ in (4.4), we obtain

either 𝔻 2 K 0 e u ~ 0 or 𝕊 1 h 0 e u ^ / 2 0 ,

which is a contradiction since both K0 and h0 are nonnegative functions somewhere positive. ∎


Communicated by Antonio Ambrosetti


Award Identifier / Grant number: MTM2015-68210-P

Funding source: Junta de Andalucía

Award Identifier / Grant number: FQM116

Award Identifier / Grant number: 713485

Funding statement: D. Ruiz has been supported by the Feder–Mineco Grant MTM2015-68210-P and by J. Andalucia (FQM116). S. Cruz-Blázquez is supported by the Istituto Nazionale di Alta Matematica “F. Severi”, and by Marie Skłodowska-Curie Actions through grant number 713485.

References

[1] A. Ambrosetti, Y. Li and A. Malchiodi, On the Yamabe problem and the scalar curvature problems under boundary conditions, Math. Ann. 322 (2002), no. 4, 667–699. 10.1007/s002080100267Search in Google Scholar

[2] T. Aubin, Meilleures constantes dans le théorème d’inclusion de Sobolev et un théorème de Fredholm non linéaire pour la transformation conforme de la courbure scalaire, J. Funct. Anal. 32 (1979), no. 2, 148–174. 10.1016/0022-1236(79)90052-1Search in Google Scholar

[3] J. Bao, L. Wang and C. Zhou, Blow-up analysis for solutions to Neumann boundary value problem, J. Math. Anal. Appl. 418 (2014), no. 1, 142–162. 10.1016/j.jmaa.2014.03.088Search in Google Scholar

[4] M. S. Berger, Riemannian structures of prescribed Gaussian curvature for compact 2-manifolds, J. Differential Geom. 5 (1971), 325–332. 10.4310/jdg/1214429996Search in Google Scholar

[5] S. Brendle, A family of curvature flows on surfaces with boundary, Math. Z. 241 (2002), no. 4, 829–869. 10.1007/s00209-002-0439-1Search in Google Scholar

[6] K. C. Chang and J. Q. Liu, A prescribing geodesic curvature problem, Math. Z. 223 (1996), no. 2, 343–365. 10.1007/BF02621603Search in Google Scholar

[7] S.-Y. A. Chang and P. C. Yang, Conformal deformation of metrics on S2, J. Differential Geom. 27 (1988), no. 2, 259–296. 10.4310/jdg/1214441783Search in Google Scholar

[8] W. X. Chen and C. Li, Prescribing Gaussian curvatures on surfaces with conical singularities, J. Geom. Anal. 1 (1991), no. 4, 359–372. 10.1007/BF02921311Search in Google Scholar

[9] P. Cherrier, Problèmes de Neumann non linéaires sur les variétés riemanniennes, J. Funct. Anal. 57 (1984), no. 2, 154–206. 10.1016/0022-1236(84)90094-6Search in Google Scholar

[10] F. Da Lio, L. Martinazzi and T. Rivière, Blow-up analysis of a nonlocal Liouville-type equation, Anal. PDE 8 (2015), no. 7, 1757–1805. 10.2140/apde.2015.8.1757Search in Google Scholar

[11] Z. Djadli, A. Malchiodi and M. Ould Ahmedou, Prescribing scalar and boundary mean curvature on the three dimensional half sphere, J. Geom. Anal. 13 (2003), no. 2, 255–289. 10.1007/BF02930697Search in Google Scholar

[12] J. F. Escobar, Conformal deformation of a Riemannian metric to a scalar flat metric with constant mean curvature on the boundary, Ann. of Math. (2) 136 (1992), 1–50. 10.2307/2946545Search in Google Scholar

[13] J. F. Escobar, Conformal deformation of a Riemannian metric to a constant scalar curvature metric with constant mean curvature on the boundary, Indiana Univ. Math. J. 45 (1996), 917–943. 10.1512/iumj.1996.45.1344Search in Google Scholar

[14] V. Felli and M. Ould Ahmedou, Compactness results in conformal deformations of Riemannian metrics on manifolds with boundaries, Math. Z. 244 (2003), no. 1, 175–210. 10.1007/s00209-002-0486-7Search in Google Scholar

[15] J. A. Gálvez and P. Mira, The Liouville equation in a half-plane, J. Differential Equations 246 (2009), no. 11, 4173–4187. 10.1016/j.jde.2009.03.012Search in Google Scholar

[16] H. Hamza, Sur les transformations conformes des variétés riemanniennes à bord, J. Funct. Anal. 92 (1990), no. 2, 403–447. 10.1016/0022-1236(90)90057-RSearch in Google Scholar

[17] Z.-C. Han and Y. Li, The Yamabe problem on manifolds with boundary: Existence and compactness results, Duke Math. J. 99 (1999), no. 3, 489–542. 10.1215/S0012-7094-99-09916-7Search in Google Scholar

[18] Z.-C. Han and Y. Li, The existence of conformal metrics with constant scalar curvature and constant boundary mean curvature, Comm. Anal. Geom. 8 (2000), no. 4, 809–869. 10.4310/CAG.2000.v8.n4.a5Search in Google Scholar

[19] F. Hang and X. Wang, A new approach to some nonlinear geometric equations in dimension two, Calc. Var. Partial Differential Equations 26 (2006), no. 1, 119–135. 10.1007/s00526-005-0372-3Search in Google Scholar

[20] A. Jiménez, The Liouville equation in an annulus, Nonlinear Anal. 75 (2012), no. 4, 2090–2097. 10.1016/j.na.2011.10.009Search in Google Scholar

[21] J. L. Kazdan and F. W. Warner, Curvature functions for compact 2-manifolds, Ann. of Math. (2) 99 (1974), 14–47. 10.2307/1971012Search in Google Scholar

[22] J. L. Kazdan and F. W. Warner, Existence and conformal deformation of metrics with prescribed Gaussian and scalar curvatures, Ann. of Math. (2) 101 (1975), 317–331. 10.2307/1970993Search in Google Scholar

[23] Y. Li and P. Liu, A Moser–Trudinger inequality on the boundary of a compact Riemann surface, Math. Z. 250 (2005), no. 2, 363–386. 10.1007/s00209-004-0756-7Search in Google Scholar

[24] Y. Li and M. Zhu, Uniqueness theorems through the method of moving spheres, Duke Math. J. 80 (1995), no. 2, 383–417. 10.1215/S0012-7094-95-08016-8Search in Google Scholar

[25] P. Liu and W. Huang, On prescribing geodesic curvature on D2, Nonlinear Anal. 60 (2005), no. 3, 465–473. 10.1016/S0362-546X(04)00379-7Search in Google Scholar

[26] R. López-Soriano, A. Malchiodi and D. Ruiz, Conformal metrics with pescribed Gaussian and geodesic curvatures, preprint (2018), https://arxiv.org/abs/1806.11533. 10.29007/nl3fSearch in Google Scholar

[27] R. López-Soriano and D. Ruiz, Prescribing the Gaussian curvature in a subdomain of 𝕊2 with Neumann boundary condition, J. Geom. Anal. 26 (2016), no. 1, 630–644. 10.1007/s12220-015-9566-xSearch in Google Scholar

[28] F. C. Marques, Existence results for the Yamabe problem on manifolds with boundary, Indiana Univ. Math. J. 54 (2005), no. 6, 1599–1620. 10.1512/iumj.2005.54.2590Search in Google Scholar

[29] J. Moser, A sharp form of an inequality by N. Trudinger, Indiana Univ. Math. J. 20 (1970/71), 1077–1092. 10.1512/iumj.1971.20.20101Search in Google Scholar

[30] J. Moser, On a non-linear problem in differential geometry, Dynamical Systems, Academic Press, New York (1973), 273–280. 10.1016/B978-0-12-550350-1.50026-6Search in Google Scholar

[31] B. Osgood, R. Phillips and P. Sarnak, Extremals of determinants of Laplacians, J. Funct. Anal. 80 (1988), no. 1, 148–211. 10.1016/0022-1236(88)90070-5Search in Google Scholar

[32] N. S. Trudinger, Remarks concerning the conformal deformation of Riemannian structures on compact manifolds, Ann. Sc. Norm. Super. Pisa (3) 22 (1968), 265–274. Search in Google Scholar

[33] L. Zhang, Classification of conformal metrics on 𝐑+2 with constant Gauss curvature and geodesic curvature on the boundary under various integral finiteness assumptions, Calc. Var. Partial Differential Equations 16 (2003), no. 4, 405–430. 10.1007/s005260100155Search in Google Scholar

Received: 2018-03-23
Revised: 2018-05-29
Accepted: 2018-05-31
Published Online: 2018-07-20
Published in Print: 2018-08-01

© 2018 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 15.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2018-2021/html
Scroll to top button