Home Spacelike Graphs of Prescribed Mean Curvature in the Steady State Space
Article Open Access

Spacelike Graphs of Prescribed Mean Curvature in the Steady State Space

  • Rafael López EMAIL logo
Published/Copyright: September 27, 2016

Abstract

We study the Dirichlet problem of the constant mean curvature equation in the steady state space over a bounded domain of a slice. Under suitable conditions on the convexity of the domain, we prove the existence of a spacelike graph with constant mean curvature H, with -1H<0, and constant boundary values.

MSC 2010: 35J60; 53C50; 53C42

1 Introduction and Statement of Results

The steady state space n+1 is a model for the universe proposed by Bondi and Gold [4], and Hoyle [12] which is homogeneous and isotropic, that is, it looks the same not only at all points and in all directions, but also at all times [10, Section 5.2]. If 1n+2=(n+2, ,=dx12++dxn+12-dxn+22) denotes the n+2-dimensional Minkowski space, the steady state space n+1 is the half of the de Sitter space 𝕊1n+1={p1n+2:p,p=1}, given by n+1={p𝕊1n+1:p,a>0}, where a1n+1 is a non-zero null vector in the past half of the null cone. This space is a non-complete manifold foliated by a uniparametric family of totally umbilical hypersurfaces {Lτ:τ(0,)}, called slices, where Lτ={pn+1:p,a=τ}, and all have constant mean curvature H=-1. The boundary of n+1 is the null hypersurface L0={p𝕊1n+1:p,a=0} that represents the past infinity and L={p𝕊1n+1:p,a=} is the limit boundary that stands as the future infinity.

An equivalent model of n+1 is (+n+1,g), where +n+1=n×+, is the upper half-space of vector space n+1 endowed with the metric

(1.1) g = 1 x n + 1 2 ( d x 1 2 + + d x n 2 - d x n + 1 2 ) ,

where x=(x1,,xn+1) are the canonical coordinates of n+1. By the expression of the metric g in (1.1), the steady state space n+1 is the Lorentzian analogue to the upper half-space model of the hyperbolic space viewed as a subset of the (n+1)-dimensional Lorentz–Minkowski space 1n+1. In this model a slice is a horizontal hyperplane L(h)={xn+1:xn+1=h} for h>0.

In this paper we study the Dirichlet problem for the prescribed mean curvature equation on a domain of a slice: If Ωn is a smooth bounded domain, given φC0(Ω) and H, find a solution uC2(Ω)C0(Ω¯) of

(1.2) { div ( D u 1 - | D u | 2 ) = n u ( H + 1 1 - | D u | 2 ) in Ω , | D u | < 1 in Ω ¯ , u = φ along Ω .

Here div and D are the Euclidean divergence and gradient operators of n, respectively. If u is a solution of (1.2), its graph Σu={(x1,,xn,u(x1,,xn)):(x1,,xn)Ω} is a spacelike hypersurface in n+1 of constant mean curvature H with respect to the upwards orientation, and its boundary is the graph of φ. Spacelike hypersurfaces with constant mean curvature are of interest in a spacetime, as for example n+1, because they are used as convenient initial hypersurfaces for the Cauchy problem corresponding to the Einstein equations [15]. Also it is interesting to have foliations of the spacetime by hypersurfaces with constant mean curvature because all points of each leaf of the foliation are instantaneous observers (or normal observers) and the (timelike) unit normal vector of the hypersurfaces measures how the observers get away with respect to the next ones. Thus, the graph of a solution of (1.2) can be viewed as a local result (on the domain Ω) of prescribing the behavior of normal observers. In the simplest spacetime, the Lorentz–Minkowski space, the first remarkable result on the Dirichlet problem is due to Bartnik and Simon [3] who proved the solvability for almost any domain Ω and boundary values φ .

In the literature, the first existence result of the Dirichlet problem (1.2) appears in [16]. More recently, the existence of radially symmetric spacelike hypersurfaces in n+1 with constant mean curvature was shown in [7]. The methods employed therein were the classical Schauder fixed point theorem, and the results were extended to a wider class of spacetimes, namely, the Robertson–Walker space. However, our approach for solving (1.2) is the method of continuity following [3, 16]. Proving the existence of graphs with prescribed mean curvature and boundary requires establishing a priori C1 estimates, and these estimates are accomplished by the use of hyperbolic planes as barriers. For the question of uniqueness of solutions of (1.2), standard arguments of elliptic PDEs do not provide a complete answer (see Remark 2.2). On the other hand, the Dirichlet problem of the Gauss curvature equation was studied in [17]. Recently, the author has proved that planar discs and hyperbolic caps are the only compact spacelike surfaces with constant mean curvature in 3 spanning a circle [13].

The solvability of the Dirichlet problem (1.2) depends on whether H is less than or greater than -1, which is just the value of the mean curvature of slices. For example, for φ=h, the graph Σu lies in one side of the slice L(h), namely, Σu is above L(h) (or u>h) if H<-1 and below L(h) (or u<h) if H>-1. Notice that the space is not symmetric with respect to L(h) and one can not expect similar results if H<-1 or H>-1; see also the difference about the uniqueness in Remark 2.2. In the mentioned paper [16], Montiel showed the existence of solutions of (1.2) for φ=h and H<-1, assuming that Ω is strictly mean convex. This result allowed to prove the existence of non-compact spacelike hypersurfaces with boundary in the infinity. By the maximum principle, the value of H for these hypersurfaces must be less than -1 and, after letting h0 together with a suitable control of the C1 estimates, he established successfully the existence of complete spacelike hypersurfaces with boundary at the past infinity. The goal of the present paper is to consider the Dirichlet problem for H>-1, where getting a priori estimates is not a consequence of the case H<-1, and the technical difficulties are hard. Indeed, the existence result that we will prove holds when -1H<0 and Ω satisfies stronger convexity assumptions. Here we recall that if κ>0, a domain ΩL(h) is said to be κ-convex if the principal curvatures κ1,,κn-1 of Ω with respect to the inward normal vector satisfy κiκ. The main result that we prove is the following.

Theorem 1.1

Let -1H<0. Let ΩL(h) be a κ-convex domain strictly contained in a ball of L(h) radius 1. If

(1.3) κ 1 - H 2 ,

then there exists a solution of the Dirichlet problem (1.2) for φ=h.

Let us point out that some kind of smallness assumption on the domain Ω is needed because the following result was proved by the author in [13].

Theorem 1.2

Let -1<H<0 and let ΩL(h) be a bounded domain. If Ω contains a ball of radius

(1.4) r 0 = 1 - H 1 + H ,

then there does not exist a spacelike graph on Ω of constant mean curvature H and with boundary Ω.

The proof of this result uses an argument of comparison of the graph of a prospective solution with a uniparametric family of hyperbolic planes that have the role of barrier hypersurfaces. The value of r0 in (1.4) implies that r0>1 and this is the reason that we suppose, in Theorem 1.1, that the domain Ω is contained in a ball of radius 1. On the other hand, as L(h) is isometric to the Euclidean hyperplane of equation xn=h, then the κ-convexity of Ω means that Ω is also κ/h-convex as a submanifold of the Euclidean hyperplane xn=h.

This paper is organized as follows. In Section 2 we recall some basics of the steady state space and we obtain the prescribed mean curvature equation. Section 3 is devoted to giving the notion of a graph on a slice. The height and gradient estimates needed in the continuity method are obtained in Section 4. Finally, in Section 5 we prove Theorem 1.1. As a consequence of the methods employed in the previous sections, in Section 6 we obtain a priori gradient estimates for a solution of the Dirichlet problem for other values of H and boundary conditions φ.

2 Preliminaries

Let Σ be a connected hypersurface. A smooth immersion ψ:Σn+1 is said to be a spacelike hypersurface if the induced metric via ψ is a Riemannian metric on Σ. A spacelike hypersurface is always orientable because the causal character of the ambient space allows to choose a unique unit timelike normal vector field N, globally defined on Σ, which is future-directed. As the metric of n+1 is conformal to the Minkowski metric, the causal character of n+1 is the same to that of the upper half-space viewed as an open set of the Minkowski space 1n+1. Recall that there exists an explicit isometry Φ:n+1𝕊1n+1n+1+n+1 between both models which inverts the orientation [16]. We will also use the terminology horizontal and vertical in the Euclidean sense considering the target ambient space +n+1. In this model of n+1, the isometries are the conformal transformations of the Minkowski space 1n+1 that preserve the upper half-space +n+1. For example, the rotations about a vertical straight line are isometries of n+1 as well as the horizontal translations or homotheties from any point of n-1×{0}.

If stands for the Levi-Civita connection on Σ, the mean curvature H is defined as

(2.1) H = - trace ( W ) n = - κ 1 + + κ n n ,

where W is the shape operator and κi are the principal curvatures of ψ. When H is constant we say that Σ is a CMC hypersurface or a H-hypersurface if we want to emphasize the value of the mean curvature. We point out that the choice of the minus sign in (2.1) follows [1, 5] and it is of opposite sign to the one adopted in [16]. According to (2.1), a slice Lτ has constant mean curvature H=-1 with respect to the future unit normal vector.

In view of the metric (1.1), we can consider on Σ the Minkowski metric , and the metric g. If N is the Gauss map of ψ:Σn+1, then N=N/xn is the Gauss map of the immersion ψ:Σ1n+1. Since g and , are conformal metrics by (1.1), we have κi=xnκi+(xnN), where κi and κi are the principal curvatures of (Σ,g) and (Σ,,), respectively. From the definition of H in (2.1), we conclude

(2.2) H = x n H - ( x n N ) ,

where H and H are the mean curvatures of Σn+1 and Σ1n+1, respectively.

Convention.

In what follows, the orientation N of a spacelike hypersurface Σ in both models of n+1 will be future directed. If Σ𝕊1n+1, then, as the vector a lies in the past half of the null cone, we have N,a>0. If Σ1n+1, then N lies in the same time-orientation with 𝐞n+1=(0,,0,1), that is, g(N,𝐞n+1)<0 or, equivalently, N,𝐞n+1<0. According to this choice of future directed orientations, if H is the mean curvature of Σ𝕊1n+1, then -H is the mean curvature of Σ(+n+1,g).

The first examples of CMC hypersurfaces of n+1 are the totally umbilical hypersurfaces and among them, slices and upper hyperbolic planes stand out. A sliceL(h)={xn+1:xn+1=h} (h>0) has mean curvature H=-1 and, by the isometry Φ, we have Φ(Lτ)=L(h), where h=1/τ.

An (upper) hyperbolic plane of radius r>0 and center c=(c1,,cn+1)1n+1 is defined by

n ( r ; c ) = { p 1 n + 1 : p - c , p - c = - r 2 , p - c , 𝐞 n + 1 < 0 } .

This hypersurface has constant mean curvature H=cn+1/r for N(p)=(p-c)/r. In what follows, if the center of a hyperbolic plane is (0,,0,t), we only write n(r;t).

We finish this section by obtaining the expression of H in local coordinates. A spacelike hypersurface of n+1 (so of 1n+1) is locally a graph Σ of a function u defined in the (x1,,xn)-plane with |Du|<1. The mean curvature H of Σ, as a hypersurface of the Minkowski space 1n+1, satisfies

(2.3) div ( D u 1 - | D u | 2 ) = n H ,

with respect to the future directed unit normal vector field

(2.4) N = 1 1 - | D u | 2 ( u x 1 , , u x n , 1 ) .

For the Gauss map N=uN, the expression (2.2) combined with equation (2.3) gives the following partial differential equation for H of Σn+1:

(2.5) Q H [ u ] := div ( D u 1 - | D u | 2 ) - n u ( H + 1 1 - | D u | 2 ) = 0 .

Equation (2.5) is of quasilinear elliptic type with the property that the difference function of two solutions satisfies a linear equation. Then the strong maximum principle of Hopf [9, Theorem 3.5] applies, and this allows extending the usual tangency principle of CMC hypersurfaces in Euclidean space.

Proposition 2.1

Proposition 2.1 (The Tangency Principle)

Let Σ1 and Σ2 be two spacelike hypersurfaces in Hn+1 with an interior or boundary tangent point p and with both hypersurfaces having the same constant mean curvature. If Σ1 lies in one side of Σ2 around p, then Σ1 coincides with Σ2 in an open set around p.

Applying successively the tangency principle where Σ1 and Σ2 coincide, we find that Σ1Σ2 is an open set in Σi for i=1,2.

Remark 2.2

The uniqueness of the Dirichlet problem associated to (1.2) is not assured. If we write equation (2.5) as

i , j = 1 n a i j ( x , u , D u ) 2 u x i x j + b ( x , u , D u ) = 0 , a i j = a j i ,

then the term

b ( x , u , D u ) = - n u ( H + 1 1 - | D u | 2 )

is increasing in the variable u and we cannot apply the standard theory [9, Theorem 10.1]. However, when the domain Ω is star-shaped, there is uniqueness of solutions (see [13] and [16, Corollary 12]).

3 Graphs in the Steady State Space

The notion of a graph for the Dirichlet problem (1.2) coincides with the graph on a domain of a slice of n+1. We need to link this concept in both models to obtain, in Section 4, the height and gradient estimates needed for solving (1.2). Let h>0. Given a domain Ωn, take Ω×{h}L(h). If f is a function on Ω, we associate to each qΩ×{h} the point of the geodesic of n+1 passing by q and orthogonal to L(h) at distance f(q) from L(h). In the upper half-space model, this type of geodesic is a vertical straight line, so a graph on Ω×{h} is an Euclidean graph xn+1=u(x1,,xn) on Ω, where u=hef. In the model of n+1 as a subset of 𝕊1n+1, the orthogonal geodesic to ΩLτ at qΩ is

γ ( t ) = cosh ( t ) q + sinh ( t ) ( - q + a τ ) .

If f is a function on Ω, the graph Σf of f is

Σ f = { X ( q ) = cosh ( f ( q ) ) q + sinh ( f ( q ) ) ( - q + a τ ) : q Ω } .

Notice that

(3.1) X , a = τ e - f .

The tangent space of Σ at X(q) is

T X ( q ) Σ f = { - f , v e - f q + e - f v + cosh ( f ) f , v τ a : v T q L τ } ,

where is the gradient operator in Lτ. The unit normal vector field N pointing to the future is

N ( X ( f ) ) = 1 e - 2 f - | f | 2 ( e - 2 f q - 1 τ e - f cosh ( f ) a - f ) .

A computation on 𝕊1n+1 then gives

N , a = τ e - f 1 - e 2 f | f | 2 .

By the isometry Φ, let Σu denote the corresponding graph on the domain Φ(Ω)ΩL(h), where we relate the functions f and u. If we let p stand for a point in 𝕊1n+1 and Φ(p)=x+n+1, then from (3.1) it follows that

(3.2) p , a = 1 u = τ e - f ,

so h=1/τ. We also express N, defined in Σ𝕊1n+1, with the Gauss map N given in (2.4). By the isometry Φ between both models, we find N,a=(xn+1N)/xn+1, and thus

(3.3) N , a = 1 u 1 - | D u | 2 .

From the expression of N in (2.4), we have

N , 𝐞 n + 1 = - 1 1 - | D u | 2 .

Two examples of graphs are slices and hyperbolic planes. For a slice L(h), we have u=h, N=𝐞n+1 and N,𝐞n+1=-1. For a hyperbolic plane n(r;t), t, we have

u ( x 1 , , x n ) = t + x 1 2 + + x n 2 - r 2

and

N , 𝐞 n + 1 = - 1 r x 1 2 + + x n 2 - r 2 .

The control of the functions p,a and N,a given in (3.2) and (3.3), respectively, will be decisive in the proof of Theorem 1.1 when we establish a priori C1 estimates of a solution u. These estimates will be based on the expressions of the Laplacians of p,a and N,a. For the function p,a, we are identifying p with its image ψ(p) by the immersion ψ:Σ𝕊1n+1. With respect to the induced metric on Σ via ψ, the Laplacian of p,a (or ψ(p),a) is

(3.4) Δ p , a = - n p , a + n H N , a .

If we now suppose that the mean curvature is constant, then the Laplacian of N,a is

(3.5) Δ N , a = | σ | 2 N , a - n H p , a ,

where σ is the second fundamental form of ψ, see [16].

4 The Method of Continuity

The technique used in the proof of Theorem 1.1 is the method of continuity [9], in the context of the prescribed mean curvature equation, see, e.g., [2, 6, 14]. After an isometry of n+1, we assume that h=1, that is, Ω is included in the slice L(1). In this slice, the curvature Ω coincides with the Euclidean one as subset of n. The method of continuity considers the family of Dirichlet problems

(4.1) { Q H ( t ) [ u ] = 0 in Ω , u = 1 along Ω ,

where H(t)=t(1+H)-1 and t[0,1]. We show that the subset of [0,1] defined by

A = { t [ 0 , 1 ] : there exists u t C 2 , α ( Ω ) such that Q H ( t ) ( u t ) = 0 and u t | Ω = 1 }

is non-empty, closed and open in [0,1]. In such case, 1A, proving the existence of a solution uC2,α(Ω¯). As H is constant and Ω is smooth, any C2,α solution will be smooth on Ω¯ (see [9, Theorem 6.17]), proving Theorem 1.1.

Let us observe that if H=-1 then the solution of (1.2) for φ=1 is the constant function u=1. Moreover, this solution is unique as a consequence of the tangency principle comparing Σu with the slices L(c) for c>0.

The proof that A is closed follows once we establish a priori C1 estimates of the prospective solutions of (1.2), that is, height and gradient estimates for every solution u of (1.2), see [9]. The convexity condition (1.3) is required first to obtain an estimate of the C0 norm of the solution comparing the graph with hyperbolic planes of type n(r;c) and, in addition, these hyperbolic planes will provide the boundary gradient estimates. Establishing gradient estimates, |Du| goes through to prevent that |Du|1 at a point of Ω¯ or, in the terminology of Marsden and Tipler, that the hypersurface cannot ‘go null’ [15, p. 124]. This is the hard part of the proof of Theorem 1.1, and it will be proved in the next subsections.

4.1 Height Estimates

We introduce the next notation. For each t0, the light cone with apex ξ=(0,,0,t) is defined by 𝒞t={x+n+1:x-ξ,x-ξ=0}. The apex ξ separates the cone in two half-cones. Let 𝒞t+ denote the upper half-cone and let int(𝒞t+) be the convex domain of +n+1 bounded by 𝒞t+.

Let -1<H<0 and let u be a solution of (1.2) for u=1 along Ω. By the tangency principle, comparing with slices L(h) for values of h going from h= to h=1, we have u<1 in Ω. For any t[0,1), let 𝒞t+ be the upper light cone. Denote by Ωρn the ball of radius ρ>0 centered at the origin O and suppose ΩΩ1. Since the inclusion ΩΩ1 is strict, let ρ0<1 be a number such that Ω¯Ωρ0. As u=1 on Ω, the spacelike condition |Du|<1 and the convexity of Ω imply that |u(q)-u(q)|<|q-q| whenever q,qΩ for qq. This inequality together with the fact that ΣuΩρ0×{1} imply Σuint(𝒞1-ρ0+), see Figure 1. Because the apex of 𝒞1-ρ0+ is (1-ρ0)p0, where p0=(0,0,1), we find

(4.2) 1 - ρ 0 < x n + 1 ( p ) 1

for all pΣu, obtaining the uniform height estimate for u.

Figure 1 
						Proof of the boundary gradient estimates in Theorem 1.1.
Figure 1

Proof of the boundary gradient estimates in Theorem 1.1.

4.2 Boundary Gradient Estimates

We establish an a priori gradient estimate along the boundary Ω.

Proposition 4.1

Under the assumptions of Theorem 1.1, we have

(4.3) sup Ω | D u | 1 - H 2

for every solution u of the Dirichlet problem (1.2).

Proof.

If H=-1, then we know that u=1 and |Du|=0, hence (4.3) is trivial. Let -1<H<0. Consider the uniparametric family of hyperbolic planes

{ n ( r ; H r ) : r ( 0 , r 1 ] } , where r 1 = H H 2 - 1 .

These hyperbolic planes satisfy the following properties:

  1. We have n(r;Hr)L(1)=ΩR(r)×{1}, where R(r)=(1-Hr)2-r2. In the interval (0,r1], we have R(r)>1 and the function R(r) is increasing in r, taking all values between 1 and 1/1-H2.

  2. The vertex of n(r;Hr) is V(r)=(H+1)rp0. Then V(r) goes from O at r=0 until V(r1)=p0/2.

  3. The intersection n(r;Hr)𝒞0 is a sphere of radius

    z ( r ) = r ( H 2 - 1 ) 2 H ,

    included in the slice L(z(r)). The part of n(r;Hr) inside the open set int(𝒞0+) lies below L(z(r)). The function z(r) is increasing on r, with z(0)=0.

We now proceed to prove estimate (4.3). Let r0>0 be sufficiently close to r=0 such that z(r0)<1-ρ0, where ρ0 is the number obtained in (4.2). Because 𝒞1-ρ0+int(𝒞0+), property (iii) implies that n(r0;Hr0) does not intersect Σu, and that n(r0;Hr0) lies below Σu (with respect to the height coordinate xn+1). Let us observe that n(r0;Hr0)L(1)=ΩR(r0)×{1} with R(r0)>1. We increase r from the value r=r0 to r=r1. By property (i) and as the radius R(r) of the ball ΩR(r) increases in r, the intersection n(r;Hr)L(1) does not meet Σu because the assumption on Σu asserts that ΣuΩ1×{1}, see Figure 1. Since the mean curvature of every hyperbolic plane n(r;Hr) is H, the tangency principle assures that there are no touching points between Σu and n(r;Hr) for all r0rr1. In particular, and once we arrive at r=r1, the graph Σu lies above the hyperbolic plane n(r1;Hr1), finding a new lower height estimate for Σu, namely,

H H - 1 < x n + 1 ( p ) , p Σ u .

Let ρ1=1/1-H2. We recall by property (i) that the intersection of n(r1;Hr1) with the slice L(1) is an Euclidean sphere of radius ρ1. The assumption on the k-convexity of Ω says κ1/ρ1, so the domain Ω has the following Blaschke outer rolling sphere property: for every point in the boundary Ω, there exists a ball in L(1) of radius ρ1 touching that point and the interior of the ball includes Ω, see, e.g., [11]. Thus, for each qΩ, it is possible to horizontally move Ωρ1×{1} in the hyperplane L(1), so in the new position, we have qΩρ1Ω and ΩΩρ1. Consequently, for each qΩ, we take the initial hyperbolic plane n(r1;Hr1) and move it horizontally until it touches Ω×{1} at the point q*=(q,1), as it was described previously. Then there exists a neighborhood of q in Ω such that the graph Σu of u lies locally sandwiched between the hyperbolic plane n(r1;Hr1) and the slice L(1). This implies that at the point q*, the value N(q*),𝐞n+1 is bounded from above and from below by the value at q* of the scalar product of 𝐞n+1 with the Gauss maps of L(1) and n(r1;Hr1), respectively. For the slice L(1), this value is -1 and for n(r1;Hr1), it is -R(r1)2-r12/r1=1/H. Then we have the inequality 1/H<N(q*),𝐞n+1<-1. From (3.3), it follows that |Du|(q)<1-H2, and thus

(4.4) sup Ω | D u | < 1 - H 2 .

In particular, the gradient estimate for |Du| does not depend on q. ∎

4.3 Gradient Estimates

We will obtain a global a priori estimate for |Du| on the domain Ω based on the estimates of |Du| along Ω. This is accomplished by studying the Jacobi equation that satisfies the Gauss map of a spacelike CMC hypersurface. The next result holds for H0 and guarantees that obtaining global estimates of the gradient reduces to getting boundary gradient estimates. Here we consider a general case of boundary data in the Dirichlet problem assuming u=φ along Ω, where φC0(Ω). Let us introduce the next notation:

φ m = inf Ω φ , φ M = sup Ω φ .

Proposition 4.2

Let H0. Let also ΩRn be a bounded domain and u a bounded solution of the Dirichlet problem (1.2). Denote

u m = inf Ω u , u M = sup Ω u .

If there exists a positive constant C1<1 such that

sup Ω | D u | C 1 ,

then there is a constant C=C(H,φm,um,uM,C1)<1 such that

(4.5) sup Ω | D u | C .

Proof.

Denote by Σu the graph of u in both models of n+1. Since the isometry Φ inverts the orientations, H0 when N,a>0 in the model of n+1, a subset of 𝕊1n+1. Then (3.3) implies

sup Σ u N , a = sup Ω 1 u 1 - | D u | 2 1 φ m 1 - C 1 2 := C 2 ,

where C2=C2(φm,C1). On the other hand, from identities (3.4) and (3.5), we have

Δ ( - H p , a + N , a ) = ( | σ | 2 - n H 2 ) N , a 0 ,

because |σ|2nH2. Then -Hp,a+N,a is a subharmonic function and the maximum principle yields

(4.6) - H p , a + N , a sup Σ u ( - H p , a + N , a ) sup Σ u N , a C 2 .

Hence, (3.2) implies

N , a H p , a + C 2 H u m + C 2 := C 3 ,

where C3=C3(H,um,C2). We substitute this inequality into (3.3) obtaining

1 u 1 - | D u | 2 C 3 .

Thus,

| D u | 1 - 1 C 3 2 u 2 1 - 1 C 3 2 u M 2 := C ,

and this proves the result. ∎

The interior gradient estimate needed for Theorem 1.1 is now a consequence of Propositions 4.1 and 4.2. However, we can give explicitly an expression of this estimate. From (4.4) and (4.6) (now 0<H<1 and h=1), we have

(4.7) - H p , a + N , a sup Σ u ( - H + 1 1 - | D u | 2 ) < 1 - H 2 H .

Hence, we obtain N,a, and using the value of p,a in (3.2), we have

N , a < H u + 1 - H 2 H .

Using (3.3) again, we conclude

(4.8) sup Ω | D u | < 1 - H 2 .

5 Proof of Theorem 1.1

Now that the height and gradient estimates are obtained, for a solution of the Dirichlet problem, we prove definitively Theorem 1.1. Again suppose h=1 on the boundary values in (1.2). If H=-1, then u=1 is a solution of (1.2). We suppose -1<H<0 and use the notation of Section 4. In the continuity method, let ut be a solution of the Dirichlet problem (4.1). First, we note that A is non-empty because u=1 is a solution, so 0A.

In order to prove that A is a closed subset of [0,1], we have to find C1 estimates of the solutions ut for tA. The height estimates (4.2) hold for all values of the mean curvature between -1 and 0. From the global gradient estimate (4.8), we have

| D u t | < 1 - H ( t ) 2 1 - H ( 1 ) 2 = 1 - H 2 .

The proof that A is an open subset of [0,1] is a consequence of the implicit function theorem. The linearization of the mean curvature is the Jacobi operator, corresponding to the second variation of the area functional, and it is given by L=Δ-|σ|2+n, see [16]. In this context, if we show that the kernel of L is trivial, then L is a self-adjoint Fredholm operator of index 0, hence L is invertible. From the implicit function theorem for Banach spaces, it follows that if the Dirichlet problem (1.2) can be solved for the value H=H0, then it can also be solved in an open interval around H0. However, we point out that we cannot apply the standard theory [9, Theorem 10.1].

We prove the openness of A by considering the following eigenvalue problem for L:

{ L [ f ] + λ f = 0 on Σ , f = 0 on Σ .

Associated to L, there is a quadratic form Q acting on the subspace C0(Σ) of smooth functions on Σ satisfying the condition f=0 on Σ. Here Q is

Q ( f ) = - Σ f L [ f ] 𝑑 Σ .

We say that Σ is strongly stable if Q(f)0 for all fC0(Σ) which, in terms of the spectrum of L, is equivalent to say that the first eigenvalue λ1(L) of L is non-negative, so the kernel of L is zero. The next result closely follows [8, Theorem 1].

Lemma 5.1

Let Σ be a compact CMC spacelike hypersurface in Hn+1. Assume that there exists a function g on Σ such that g>0 on Σ and L[g]0. Then Σ is strongly stable.

Proof.

Let fC(Σ) with f=0 on Σ. Define h=log(g). Since L[g]0, it follows that Δh|σ|2-n-|h|2. Multiplying by f2 and integrating on Σ, we have

(5.1) Σ ( | σ | 2 f 2 - n f 2 ) 𝑑 Σ - Σ f 2 | h | 2 𝑑 Σ Σ f 2 Δ h 𝑑 Σ .

As div(f2h)=f2Δh+2ff,h, the divergence theorem yields

- Σ f 2 Δ h 𝑑 Σ = 2 Σ f f , h 𝑑 Σ
2 Σ | f | | h | | f | d Σ
Σ f 2 | h | 2 d Σ + Σ | f | 2 d Σ .

Combining this inequality with (5.1), we get Q(f)0. In fact, if f0, then Q(f)>0, because in case of Q(f)=0, we find that f is proportional to h, contradicting that h0 along M. ∎

Lemma 5.2

The subset A is open in [0,1].

Proof.

Let t0A and set H0=H(t0). Denote by Σ the graph of ut0. Using Lemma 5.1, the kernel of the Jacobi operator L is zero provided that we find a function g>0 on Σ such that L[g]0. Consider the model of n+1 as a subset of 𝕊1n+1, where we now have 0<H01 and N,a>0. Define

g = p , a - H 0 N , a .

By making use of (3.4) and (3.5), we obtain

L [ g ] = ( n H 0 2 - | σ | 2 ) p , a 0 ,

since |σ|2nH02. It remains to prove that g>0 on Σ. From inequality (4.7), we have

H 0 N , a < H 0 2 p , a + 1 - H 0 2 .

Hence, we deduce

g = p , a - H 0 N , a > ( 1 - H 0 2 ) ( p , a - 1 ) > 0 ,

where we have used the fact that u1 and the relation between u and p,a given in (3.2). ∎

Finally, this lemma concludes the proof of Theorem 1.1.

Remark 5.3

One may use a similar argument to give an alternative proof that the analogous subset A in [16, p. 931] is open in [0,1]. Recall that it was proved (see [16, Corollary 8]) that a uniform gradient estimate holds for a solution u of the Dirichlet problem (1.2) when φ=h and H>1 in the de Sitter model for n+1, namely,

(5.2) sup Ω | D u | H 2 - 1 H .

Now Lemma 5.1 applies by letting g=N,a. So we have g>0, and taking into account (3.2), (3.3) and (3.5), we obtain

L [ g ] = - n H p , a + n N , a = n u ( - H + 1 1 - | D u | 2 ) n u ( - H + H ) = 0 ,

where, in the last inequality, we have used the gradient estimate (5.2).

Remark 5.4

The solutions obtained in Theorem 1.1, as well as the Montiel solutions in [16], are strongly stable.

Remark 5.5

We also have the following result: If H0 in the de Sitter space model of n+1, every spacelike H-graph on a bounded domain of a slice is strong stable. The proof follows taking the function g=p,a, which is positive, and observing that L[g]=nHN,a-|σ|2p,a0. Notice that this result holds for every domain Ω and boundary data φ along Ω, in contrast to the graphs that appear in [16] and in Theorem 1.1, where the boundary of the graph is a closed curve contained in a slice (φ=h).

6 Other Gradient Estimates Results

We have seen in Proposition 4.2 that boundary gradient estimates transfer into the interior of the domain. In this section we will extend these estimates for other assumptions on the value of H and boundary data φ. The following results have their own interest although they have not been used in this paper. The first one improves the estimate |Du| given in (4.5) when H-1.

Proposition 6.1

Let H-1. Let also ΩRn be a bounded domain and u a bounded solution of the Dirichlet problem (1.2). If there exists a positive constant C1<1 such that

sup Ω | D u | C 1 ,

then there is a constant C=C(H,φm,uM,C1)<1 such that

sup Ω | D u | C .

Proof.

We consider the model of n+1 as subset of 𝕊1n+1, where we know that H1. Using that |σ|2nH2, from (3.5), (3.2) and (3.3), we have

Δ N , a n H ( H N , a - p , a ) = n H u 1 - | D u | 2 ( H - 1 - | D u | 2 ) 0 .

The maximum principle yields N,asupΣuN,a. As uuM, from (3.3) we have that

1 u M 1 - | D u | 2 1 u 1 - | D u | 2 sup Ω 1 u 1 - | D u | 2 = 1 φ m 1 - C 1 2 ,

obtaining definitively

(6.1) | D u | 2 1 - φ m 2 u M 2 ( 1 - C 1 2 ) .

This completes the proof. ∎

Let us compare the estimates obtained in this result as well as Proposition 4.2 with the ones in [16]. Montiel gets directly the interior gradient estimates (5.2) from the hypothesis of mean convexity of the domain Ω. Our estimates are derived once we have boundary gradient estimates, and they hold for any bounded domain and any boundary condition φ. When the value of the mean curvature is non-negative, it is possible to obtain the same gradient estimate in the interior and in the boundary of Ω. In particular, we have the following result.

Proposition 6.2

Let H0. Let also ΩRn be a bounded domain and u a bounded solution of the Dirichlet problem (1.2). If there exists a positive constant C1<1 such that

sup Ω | D u | C 1 ,

then there is a constant C=C(H,φm,φM,C1)<1 such that

sup Ω | D u | C .

In the particular case where φ=h on Ω, we have

sup Ω | D u | = sup Ω | D u | .

Proof.

As H0, the tangency principle comparing Σu with the slices L(h) gives uφM. On the other hand, following the same steps as in the above proof, and as H0 in the model of 𝕊1n+1, we have ΔN,a0. The maximum principle gives the same inequality (6.1) but now uM=φM. The second part is immediate because φm=φM. ∎

7 Discussion and Conclusions

In this paper we have studied the Dirichlet problem for the prescribed mean curvature equation on a domain Ω of a slice in the steady state space n+1. This space can foliate by slices, which are spacelike hypersurfaces with constant mean curvature H=-1. The model for n+1, usually utilized in the literature, has been considered as a subset of the de Sitter space 𝕊1n+1. However, here we have worked with the upper half-space model of n+1, which is analogue to the hyperbolic space in the Lorentz–Minkowski space 1n+1. In this model, a slice L(h) is just a horizontal hyperplane of equation xn=h. Notice that the space is not symmetric with respect to a slice L(h), and the value H=-1 is critical for the geometrical behavior of a spacelike constant mean curvature hypersurface of n+1. In this sense, Montiel showed the existence of solutions when H<-1, assuming that Ω is strictly convex. In our work we studied the solvability of the Dirichlet problem for values of H with H>-1, a scenario that has not been previously considered. Now the graph of a solution lies in the halfspace of n+1 determined by L(h) pointing to the past and, in contrast to the case H<-1, there are restrictions on the size of the domain Ω, indeed, Ω can not be very large. We used the continuity method to solve the Dirichlet problem and we needed to get C1 a priori estimates of the prospective solutions. We utilized a rolling argument of comparison, with hyperbolic planes as barriers, to establish the boundary gradient estimates. Here we required that the domain is κ-convex and -1H<0. Once the boundary gradient estimates were obtained, we proved the existence of gradient estimates in the interior of Ω. This argument is typical in elliptic PDEs theory but it contrasts to the one utilized by Montiel when H<-1, where the mere mean convexity assumption of Ω gives directly the needed gradient estimates. The case H>-1 was harder and this was reflected, for instance, on the proof of the openness step in the continuity method, where the previous C1 estimates were suitably utilized. It was desirable that the κ-convexity assumption be replaced by a weaker hypothesis, as for example, that Ω was mean convex. This is expectable when H<-1 but it seems more difficult when H>-1. We discussed and initiated this process because we obtained interior gradient estimates once that, previously, we deduced gradient estimates along the boundary Ω.


Communicated by Ireneo Peral


Funding statement: The author has been partially supported by the grant number MTM2014-52368-P of the Spanish Ministerio de Economá y Competitividad and the European Regional Development Fund.

References

[1] Albujer A. and Alías L. J., Spacelike hypersurfaces with constant mean curvature in the steady state space, Proc. Amer. Math. Soc. 137 (2009), 711–721. 10.1090/S0002-9939-08-09546-4Search in Google Scholar

[2] Alías L. J., Dajczer M. and Rosenberg H., The Dirichlet problem for CMC surfaces in Heisenberg space, Calc. Var. Partial Differential Equations 30 (2007), 513–522. 10.1007/s00526-007-0101-1Search in Google Scholar

[3] Bartnik R. and Simon L., Spacelike hypersurfaces with prescribed boundary values and mean curvature, Comm. Math. Phys. 87 (1982), 131–152. 10.1007/BF01211061Search in Google Scholar

[4] Bondi H. and Gold T., On the generation of magnetism by fluid motion, Mon. Not. R. Astron. Soc. 110 (1950), 607–611. 10.1093/mnras/110.6.607Search in Google Scholar

[5] Colares A. G. and de Lima H. F., Spacelike hypersurfaces with constant mean curvature in the steady state space, Bull. Belg. Math. Soc. Simon Stevin 17 (2010), 287–302. 10.36045/bbms/1274896207Search in Google Scholar

[6] Dajczer M. and Ripoll J., An extension of a theorem of Serrin to graphs in warped products, J. Geom. Anal. 15 (2005), 193–205. 10.1007/BF02922192Search in Google Scholar

[7] de la Fuente D., Romero A. and Torres P. J., Radial solutions of the Dirichlet problem for the prescribed mean curvature equation in a Robertson–Walker spacetime, Adv. Nonlinear Stud. 15 (2015), 171–182. 10.1515/ans-2015-0109Search in Google Scholar

[8] Fischer-Colbrie D. and Schoen R., The structure of complete stable minimal surfaces in 3-manifolds of non-negative scalar curvature, Comm. Pure Appl. Math. 33 (1980), 199–211. 10.1002/cpa.3160330206Search in Google Scholar

[9] Gilbarg D. and Trudinger N. S., Elliptic Partial Differential Equations of Second Order, Classics Math., Springer, Berlin, 2001. 10.1007/978-3-642-61798-0Search in Google Scholar

[10] Hawking S. W. and Ellis G. F. R., The Large Scale Structure of Space-Time, Cambridge University Press, Cambridge, 1973. 10.1017/CBO9780511524646Search in Google Scholar

[11] Howard R., Blaschke’s rolling theorem for manifolds with boundary, Manuscripta Math. 99 (1999), 471–483. 10.1007/s002290050186Search in Google Scholar

[12] Hoyle F., A new model for the expanding universe, Mon. Not. R. Astron. Soc. 108 (1948), 372–382. 10.1093/mnras/108.5.372Search in Google Scholar

[13] López R., A characterization of hyperbolic caps in the steady state space, J. Geom. Phys. 98 (2015), 214–226. 10.1016/j.geomphys.2015.08.006Search in Google Scholar

[14] López R. and Montiel S., Existence of constant mean curvature graphs in hyperbolic space, Calc. Var. Partial Differential Equations 8 (1999), 177–190. 10.1007/s005260050122Search in Google Scholar

[15] Marsden J. E. and Tipler F. J., Maximal hypersurfaces and foliations of constant mean curvature in general relativity, Phys. Rep. 66 (1980), 109–139. 10.1016/0370-1573(80)90154-4Search in Google Scholar

[16] Montiel S., Complete non-compact spacelike hypersurfaces of constant mean curvature in de Sitter spaces, J. Math. Soc. Japan 55 (2003), 915–938. 10.2969/jmsj/1191418756Search in Google Scholar

[17] Spruck J. and Xiao L., Convex spacelike hypersurfaces of constant curvature in de Sitter space, Discrete Contin. Dyn. Syst. Ser. B 17 (2012), 2225–2242. 10.3934/dcdsb.2012.17.2225Search in Google Scholar

Received: 2016-07-01
Revised: 2016-08-26
Published Online: 2016-09-27
Published in Print: 2016-11-01

© 2016 by De Gruyter

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

Downloaded on 7.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2015-5056/html
Scroll to top button