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

Sobolev Inequalities in Spacelike Submanifolds of Minkowski Space

Liang Xu School of Mathematical Sciences, Zhejiang University, Hangzhou 310058, China [email protected]
Abstract.

We follow the method of ABP estimate in [Bre21] and apply it to spacelike submanifolds in n,1\mathbb{R}^{n,1}. We then obtain Michael-Simon type inequalities. Surprisingly, our investigation leads to a Sobolev inequality without a mean curvature term, provided the hypersurface is mean convex.

1. Introduction

In this paper we are mainly concerned with a specific type of Sobolev inequality for spacelike submanifolds in Minkowski space. Associated to a spacelike submanifold Σmn,1\Sigma^{m}\hookrightarrow\mathbb{R}^{n,1}, we define the maximal slope by

τ(Σ)=sup{|ν0(x)|:xΣ,ν(x) is a unit normal to Σ at x}.\tau(\Sigma)=\sup\{|\nu_{0}(x)|:x\in\Sigma,\nu(x)\text{ is a unit normal to }\Sigma\text{ at }x\}.

See 2.1 for details. The main results are as follows.

Theorem 1.1.

Suppose Σnn,1\Sigma^{n}\subseteq\mathbb{R}^{n,1} is a smooth, compact and spacelike hypersurface. Assume that Σ\Sigma is mean convex and that ff is any smooth and positive function defined on Σ\Sigma. Then

(1.1) Σ|f|+ΣfCn,τ(Σfnn1)n1n,\int_{\Sigma}|\nabla f|+\int_{\partial\Sigma}f\geq C_{n,\tau}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}},

where the constant Cn,τ=nωn1n(n+1)1nτ1n(τ+τ21)1C_{n,\tau}=n\omega_{n}^{\frac{1}{n}}(n+1)^{-\frac{1}{n}}\tau^{-\frac{1}{n}}(\tau+\sqrt{\tau^{2}-1})^{-1}.

Under the condition of mean convexity, there is no mean curvature term involved, which, to our knowledge, is new, Without the assumption of mean convexity, similar result holds with a curvature term involved.

Theorem 1.2.

Suppose Σnn,1\Sigma^{n}\subseteq\mathbb{R}^{n,1} is a smooth, compact and spacelike hypersurface. Assume that ff is any smooth and positive function defined on Σ\Sigma. Then

(1.2) Σ|f|2+f2H2+ΣfCn,τ(Σfnn1)n1n,\int_{\Sigma}\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}}+\int_{\partial\Sigma}f\geq C_{n,\tau}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}},

with Cn,τ=21nωn1n(n+1)1nτ1n(τ+τ21)1C_{n,\tau}=2^{-1}n\omega_{n}^{\frac{1}{n}}(n+1)^{-\frac{1}{n}}\tau^{-\frac{1}{n}}(\tau+\sqrt{\tau^{2}-1})^{-1}.

Here H2=|H|20\|H\|^{2}=-|H|^{2}\geq 0; see Section 2 for details. Next we establish the same Sobolev inequality for submanifold Σmn,1\Sigma^{m}\subseteq\mathbb{R}^{n,1} of higher codimension nm+1n-m+1, with 0<m<n0<m<n. Let ν\nu be a normal vector field of Σ\Sigma with |ν|2=1|\nu|^{2}=-1. We then write TxΣ=Tx,1ΣTx,2ΣT_{x}^{\bot}\Sigma=T_{x}^{\bot,1}\Sigma\oplus T_{x}^{\bot,2}\Sigma, where Tx,2Σ=Span{ν(x)}T_{x}^{\bot,2}\Sigma=\operatorname{Span}\{\nu(x)\}. Accordingly, the mean curvature vector is decomposed into H,1+H,2H^{\bot,1}+H^{\bot,2}.

Theorem 1.3.

Suppose Σmn,1\Sigma^{m}\subseteq\mathbb{R}^{n,1} is a smooth, compact and spacelike submanifold. Let ν\nu be any normal vector field of Σ\Sigma with |ν|2=1|\nu|^{2}=-1 and ff any smooth and positive function defined on Σ\Sigma. Then

(1.3) Σ|f|2+f2(|H,1|2+H,22)+ΣfCn,m,τ(Σfmm1)m1m,\int_{\Sigma}\sqrt{|\nabla f|^{2}+f^{2}(|H^{\bot,1}|^{2}+\|H^{\bot,2}\|^{2})}+\int_{\partial\Sigma}f\geq C_{n,m,\tau}\left(\int_{\Sigma}f^{\frac{m}{m-1}}\right)^{\frac{m-1}{m}},

where the constant Cn,m,τ=m2nm(n+1)1mωn1mτ1m(τ+τ21)nmC_{n,m,\tau}=m2^{-\frac{n}{m}}(n+1)^{-\frac{1}{m}}\omega_{n}^{\frac{1}{m}}\tau^{-\frac{1}{m}}(\tau+\sqrt{\tau^{2}-1})^{-\frac{n}{m}}.

At the end of this work, we discovered that in [TW22] similar results were established. Nevertheless, compared to the results therein, our construction yields a Sobolev inequality without curvature terms for a mean convex hypersurface. The history of geometric inequalities probably dates back to ancient Greece. The classical isoperimetric inequality asserts that for a domain Σn\Sigma\subseteq\mathbb{R}^{n} with sufficiently well-behaved boundary, there holds

(1.4) |Σ|nωn1n|Σ|n1n.|\partial\Sigma|\geq n\omega_{n}^{\frac{1}{n}}|\Sigma|^{\frac{n-1}{n}}.

It is known that such an isoperimetric inequality is essentially equivalent to the W1,1W^{1,1} Sobolev inequality for domains in Euclidean space:

(1.5) Σ|f|+Σfnωn1n(Σfnn1)n1n.\int_{\Sigma}|\nabla f|+\int_{\partial\Sigma}f\geq n\omega_{n}^{\frac{1}{n}}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}}.

Intensive work has been established to extend Eq. 1.4 or Eq. 1.5 to more general settings. It has long been conjectured that the same inequality Eq. 1.4 holds in Cartan-Hadamard manifolds [Aub76]. Partial results include [Wei26, BR33, Cro84, Kle92]. See also [GS21] for a recent attempt to resolve the conjecture. It is also possible to replace areas and volumes in Eq. 1.4 by more general quermassintegrals. The resulting isoperimetric inequality is proved for hypersurfaces with certain convexity in Euclidean space. See [Gua] for details.

For a domain Σ\Sigma in a two dimensional space form of constant curvature KK, there is a neat result which states that

(1.6) 4π|Σ||Σ|2+K|Σ|2.4\pi|\Sigma|\leq|\partial\Sigma|^{2}+K|\Sigma|^{2}.

Please see [Cho05] and references therein. The same inequality Eq. 1.6 is proved by Choe and Gulliver [CG92a] for minimal surfaces Σ2\Sigma^{2} with certain topological constraints in hyperbolic space n\mathbb{H}^{n}. Yau [Yau75], Choe and Gulliver [CG92] showed that if Σ\Sigma is a domain in n\mathbb{H}^{n} or a nn-dimensional minimal submanifold in n+m\mathbb{H}^{n+m}, then it satisfies the linear isoperimetric inequality

(n1)|Σ||Σ|.(n-1)|\Sigma|\leq|\partial\Sigma|.

Another linear inequality for proper minimal submanifolds in n\mathbb{H}^{n} is obtained in [MS14] using Poincaré model.

It is a longstanding conjecture that Eq. 1.4 holds true for minimal hypersurfaces in n+1\mathbb{R}^{n+1}. Using the method of sliding, Brendle fully settled the problem in a recent work [Bre21], and later extended his results to Riemannian manifolds with nonnegative Ricci curvature [Bre20]. The most classical application of the sliding method is perhaps Aleksandrov’s maximum principle. In [Cab08] Cabré first employed the sliding method and gave a simple and elegant proof of Eq. 1.4.

We follow Brendle’s method and apply it to submanifolds in the Minkowski space, and obtain some Michael-Simon-Sobolev type inequalities.

Acknowledgements. The author would like to thank Research Professor Qi-Rui Li for his instructions and many helpful discussions.

2. Notations and Preliminaries

Let n,1\mathbb{R}^{n,1} be the Minkowski space endowed with metric

g¯=dx02+dx12++dxn2.\bar{g}=-dx_{0}^{2}+dx_{1}^{2}+\cdots+dx_{n}^{2}.

The usual Euclidean metric of n+1\mathbb{R}^{n+1} is denoted by δ\delta. The volume element of n,1\mathbb{R}^{n,1}, dμg¯=detg¯dXd\mu_{\bar{g}}=\sqrt{-\det\bar{g}}dX, is just the usual Lebesgue measure. A vector YY is called unit if |Y|2=σ(Y)|Y|^{2}=\sigma(Y), where σ(Y)\sigma(Y) is the signature of YY, i.e. σ(X)=1,1,0\sigma(X)=1,-1,0 if XX is spacelike, timelike, lightlike, respectively. Finally, we define Y=σ(Y)|Y|2\|Y\|=\sqrt{\sigma(Y)|Y|^{2}}.

Let (Σn,g)(n,1,g¯)(\Sigma^{n},g)\hookrightarrow(\mathbb{R}^{n,1},\bar{g}) be a spacelike hypersuface. We denote by xx a point in Σ\Sigma, and by X(x)X(x) the corresponding position vector in n,1\mathbb{R}^{n,1}. Anything with a ‘bar’ is a quantity of the ambient space. Then the second fundamental form is defined by

¯YZ=YZh(Y,Z),Y,Z𝒳(Σ).\bar{\nabla}_{Y}Z=\nabla_{Y}Z-h(Y,Z),\quad\forall Y,Z\in\mathscr{X}(\Sigma).

Let ν(x)\nu(x) be a normal to Σ\Sigma at the point xx. Clearly ν(x)\nu(x) is also a vector in n,1\mathbb{R}^{n,1}. Denote by να(x)\nu_{\alpha}(x) the α\alpha-th coordinate of ν(x)\nu(x) in n,1\mathbb{R}^{n,1}, so that

|ν|2=ν02+ν12++νn2.|\nu|^{2}=-\nu_{0}^{2}+\nu_{1}^{2}+\cdots+\nu_{n}^{2}.
Definition 2.1.

Associated to a spacelike submanifold Σmn,1\Sigma^{m}\hookrightarrow\mathbb{R}^{n,1}, the maximal slope is defined by τ(Σ)=sup{|ν0(x)|:xΣ,ν(x)TxΣ,|ν(x)|2=1}.\tau(\Sigma)=\sup\{|\nu_{0}(x)|:x\in\Sigma,\nu(x)\in T_{x}^{\bot}\Sigma,|\nu(x)|^{2}=-1\}.

The quantity τ\tau characterizes how ‘lightlike’ Σ\Sigma is; for instance, τ=1\tau=1 if Σnn\Sigma^{n}\Subset\mathbb{R}^{n}; and τ\tau is uniquely determined by the diameter if Σnn\Sigma^{n}\Subset\mathbb{H}^{n}.

Lemma 2.2.

For any function ww on the ambient space, 2w=¯2wh,¯wg¯\nabla^{2}w=\bar{\nabla}^{2}w-\langle h,\bar{\nabla}w\rangle_{\bar{g}}.

Proof.

We assume that Txn,1T_{x}\mathbb{R}^{n,1} is spanned by orthonormal basis {εα:0αn}\{\varepsilon_{\alpha}:0\leq\alpha\leq n\}, and TxΣT_{x}\Sigma by {εi:1in}\{\varepsilon_{i}:1\leq i\leq n\}, with |εi|2=1,|ε0|2=1|\varepsilon_{i}|^{2}=1,|\varepsilon_{0}|^{2}=-1. We then compute

¯ij2w\displaystyle\bar{\nabla}^{2}_{ij}w =ij2w¯εiεj,¯w\displaystyle=\partial^{2}_{ij}w-\langle\bar{\nabla}_{\varepsilon_{i}}\varepsilon_{j},\bar{\nabla}w\rangle
=ij2wΓ¯ijkεk+Γ¯ij2ε0,wε+w0ε0\displaystyle=\partial^{2}_{ij}w-\langle\bar{\Gamma}_{ij}^{k}\varepsilon_{k}+\bar{\Gamma}_{ij}^{2}\varepsilon_{0},w^{\ell}\varepsilon_{\ell}+w^{0}\varepsilon_{0}\rangle
=ij2wΓ¯ijkkw+Γ¯ij00w.\displaystyle=\partial^{2}_{ij}w-\bar{\Gamma}_{ij}^{k}\partial_{k}w+\bar{\Gamma}_{ij}^{0}\partial_{0}w.

By Gauss formula ¯εiεj=εiεjhijε0\bar{\nabla}_{\varepsilon_{i}}\varepsilon_{j}=\nabla_{\varepsilon_{i}}\varepsilon_{j}-h_{ij}\varepsilon_{0} we see that Γ¯ijk=Γijk\bar{\Gamma}_{ij}^{k}=\Gamma_{ij}^{k} and Γ¯ij0=hij\bar{\Gamma}_{ij}^{0}=-h_{ij}. Hence

¯ij2w=ij2whij0w=ij2w+h(εi,εj),¯w\bar{\nabla}^{2}_{ij}w=\nabla^{2}_{ij}w-h_{ij}\partial_{0}w=\nabla^{2}_{ij}w+\langle h(\varepsilon_{i},\varepsilon_{j}),\bar{\nabla}w\rangle

By tensorality, the same formula holds in any coordinate systems. ∎

3. Proof of 1.1

Since Eq. 1.1 is homogeneous in ff, by normalization we may assume that

(3.1) Σf=nΣfnn1Σ|f|.\int_{\partial\Sigma}f=n\int_{\Sigma}f^{\frac{n}{n-1}}-\int_{\Sigma}|\nabla f|.

Let η\eta be the outward unit normal of Σ\partial\Sigma. Consider the following PDE

(3.2) {div(fu)=nfnn1|f|, in ΣΣ,u,ηg¯=1, along Σ.\begin{cases}\operatorname{div}(f\nabla u)=nf^{\frac{n}{n-1}}-|\nabla f|,&\text{ in }\Sigma\setminus\partial\Sigma,\\ \langle\nabla u,\eta\rangle_{\bar{g}}=1,&\text{ along }\partial\Sigma.\end{cases}

By our normalization, the equation has a solution uC2,αu\in C^{2,\alpha}. Since Σ\Sigma is mean convex, we may assume that the mean curvature vector HH is either zero or timelike pointing to the past. Let ν\nu be the unit normal and timelike vector field pointing to the past. Now fix r>0r>0. For any xΣx\in\Sigma we define 𝒜r=xΣΛr(x)\mathcal{A}_{r}=\cap_{x\in\Sigma}\Lambda_{r}(x), where

(3.3) Λr(x)={pn,1:|(pX(x))|2<r2,rpX,H1H(x)0}.\displaystyle\Lambda_{r}(x)=\left\{p\in\mathbb{R}^{n,1}:|(p-X(x))^{\top}|^{2}<r^{2},-r\leq\langle p-X,\|H\|^{-1}H(x)\rangle\leq 0\right\}.

Note that when H=0\|H\|=0, the notation H1H(x)\|H\|^{-1}H(x) simply means the normal ν(x)\nu(x).

Refer to caption
Figure 1. An illustration of Λr(x)\Lambda_{r}(x).

We write pX=sξ+tνp-X=s\xi+t\nu, where ξ\xi is a unit tangent vector and ν\nu is a unit normal vector pointing to the past. The the definition of Λr(x)\Lambda_{r}(x) implies that r<s<r,0tr-r<s<r,0\leq t\leq r. Therefore Λr(x)\Lambda_{r}(x) is a parallelogram illustrated as in Fig. 1. We continue and define

𝒰={(x,y):xΣΣ,yTxΣ,|u(x)|<1,1y,H1H(x)0},\displaystyle\mathcal{U}=\left\{(x,y):x\in\Sigma\setminus\partial\Sigma,y\in T^{\bot}_{x}\Sigma,|\nabla u(x)|<1,-1\leq\langle y,\|H\|^{-1}H(x)\rangle\leq 0\right\},
r={(x,y)𝒰:r2u(x)+rh(x),y+g(x)0}.\displaystyle\mathcal{B}_{r}=\left\{(x,y)\in\mathcal{U}:r\nabla^{2}u(x)+r\langle h(x),y\rangle+g(x)\geq 0\right\}.

Finally, we take the map Φr:TΣn,1\Phi_{r}:T^{\bot}\Sigma\to\mathbb{R}^{n,1},

(3.4) Φr(x,y)=X(x)+r(u(x)+y).\Phi_{r}(x,y)=X(x)+r(\nabla u(x)+y).
Lemma 3.1.

We have asymptotic behavior

(3.5) lim infrrn1|𝒜r|C~n,τ,\liminf_{r\to\infty}r^{-n-1}{|\mathcal{A}_{r}|}\geq\tilde{C}_{n,\tau},

where |𝒜r||\mathcal{A}_{r}| is the usual Lebesgue measure in n+1\mathbb{R}^{n+1} and C~n,τ=ωn(n+1)τ(τ+τ21)n\tilde{C}_{n,\tau}=\frac{\omega_{n}}{(n+1)\tau(\tau+\sqrt{\tau^{2}-1})^{n}}.

Proof.

We blow down 𝒜r\mathcal{A}_{r} by factor rr. As rr\to\infty, the bounded domain Σ\Sigma collapses to a single point: the origin, and each Λr(x)\Lambda_{r}(x) converges to a Λ~(x)\tilde{\Lambda}(x), specified by

Λ~(x)={p~n,1:|p~|2<1,1<p~,ν(x)<0},\tilde{\Lambda}(x)=\{\tilde{p}\in\mathbb{R}^{n,1}:|\tilde{p}^{\top}|^{2}<1,-1<\langle\tilde{p},\nu(x)\rangle<0\},

where ν(x)n\nu(x)\in-\mathbb{H}^{n} is the unit normal to Σ\Sigma at xx pointing to the past. See the left of Fig. 2 for an illustration. Clearly Λ~(x)\tilde{\Lambda}(x) contains the cone 𝒞~(x)={p~n,1:|p~|2<0,1<p~,ν(x)<0}\tilde{\mathcal{C}}(x)=\{\tilde{p}\in\mathbb{R}^{n,1}:|\tilde{p}|^{2}<0,-1<\langle\tilde{p},\nu(x)\rangle<0\}.

Refer to caption
Figure 2. Left: blowing down Λ(x)\Lambda(x) to get Λ~(x)\tilde{\Lambda}(x), which contains 𝒞~(x)\tilde{\mathcal{C}}(x), the shaded area. Right: 𝒜~\tilde{\mathcal{A}} contains at least 𝒞~\tilde{\mathcal{C}}, the shaded area.

Therefore 𝒜~=xΣΛ~(x)\tilde{\mathcal{A}}=\cap_{x\in\Sigma}\tilde{\Lambda}(x) contains at least 𝒞~=xΣ𝒞~(x)\tilde{\mathcal{C}}=\cap_{x\in\Sigma}\tilde{\mathcal{C}}(x). Since by assumption ν(x)\nu(x) has minimal height τ-\tau, we may assume that 𝒞~\tilde{\mathcal{C}} is the union of two cones, as illustrated by the right of Fig. 2, with the zeroth coordinate of point aa being τ-\tau.

We now proceed by computing the volume of 𝒞~\tilde{\mathcal{C}}. Without loss of generality, we assume that the points a,b,ca,b,c lie in the plane Ox0x1Ox_{0}x_{1}. Then a=(τ,τ21,0,,0)a=(-\tau,-\sqrt{\tau^{2}-1},0,\cdots,0), and the tengential ac\vec{ac} is parallel to ξ=(τ21,τ,0,,0)\xi=(\sqrt{\tau^{2}-1},\tau,0,\cdots,0). From this we readily obtain b=(τ1,0,,0)b=(-\tau^{-1},0,\cdots,0) and c=(τ+τ21,ττ21,0,,0)c=(-\tau+\sqrt{\tau^{2}-1},\tau-\sqrt{\tau^{2}-1},0,\cdots,0). Consequently

|𝒞~|=ωn(ττ21)nτ1(n+1)1=C~n,τ|\tilde{\mathcal{C}}|=\omega_{n}(\tau-\sqrt{\tau^{2}-1})^{n}\cdot\tau^{-1}\cdot(n+1)^{-1}=\tilde{C}_{n,\tau}

Hence lim infrrn1|𝒜r||𝒞~|=C~n,τ\liminf_{r\to\infty}r^{-n-1}|\mathcal{A}_{r}|\geq|\tilde{\mathcal{C}}|=\tilde{C}_{n,\tau}. ∎

Lemma 3.2.

There holds Φr(r)𝒜r\Phi_{r}(\mathcal{B}_{r})\supseteq\mathcal{A}_{r}.

Proof.

For any given p𝒜rp\in\mathcal{A}_{r}, consider the function

F(x)=ru(x)+12|pX(x)|2,xΣ.F(x)=ru(x)+\frac{1}{2}|p-X(x)|^{2},\quad x\in\Sigma.

By compactness FF attains its minimum at some x¯Σ\bar{x}\in\Sigma. We claim that x¯Σ\bar{x}\notin\partial\Sigma. For if otherwise x¯Σ\bar{x}\in\partial\Sigma, then at this point 0F(x¯),η=rpX(x¯),η.0\geq\langle\nabla F(\bar{x}),\eta\rangle=r-\langle p-X(\bar{x}),\eta\rangle. We have by definition of 𝒜r\mathcal{A}_{r} that pX(x¯),η=pX(x¯),η<r\langle p-X(\bar{x}),\eta\rangle=\langle p-X(\bar{x})^{\top},\eta\rangle<r, a contradiction. Hence x¯ΣΣ\bar{x}\in\Sigma\setminus\partial\Sigma and at which F(x¯)=ru(x¯)(pX(x¯))=0\nabla F(\bar{x})=r\nabla u(\bar{x})-(p-X(\bar{x}))^{\top}=0. We then find y¯Tx¯Σ\bar{y}\in T_{\bar{x}}^{\bot}\Sigma such that ry¯=(pX(x¯))r\bar{y}=(p-X(\bar{x}))^{\bot}. Obviously r|u(x¯)|=|(pX(x¯))|<rr|\nabla u(\bar{x})|=|(p-X(\bar{x}))^{\top}|<r and rHry¯,H(x¯)=pX(x¯),H(x¯)0-r\|H\|\leq r\langle\bar{y},H(\bar{x})\rangle=\langle p-X(\bar{x}),H(\bar{x})\rangle\leq 0. Finally, by 2.2,

0\displaystyle 0 2F(x¯)=r2u(x¯)+¯2(12|pX(x¯)|2)h,¯(12|pX(x¯)|2)\displaystyle\leq\nabla^{2}F(\bar{x})=r\nabla^{2}u(\bar{x})+\bar{\nabla}^{2}\left(\frac{1}{2}|p-X(\bar{x})|^{2}\right)-\left\langle h,\bar{\nabla}\left(\frac{1}{2}|p-X(\bar{x})|^{2}\right)\right\rangle
=r2u(x¯)+g(x¯)h,X(x¯)p\displaystyle=r\nabla^{2}u(\bar{x})+g(\bar{x})-\langle h,X(\bar{x})-p\rangle
=r2u(x¯)+g(x¯)+ry¯,h,\displaystyle=r\nabla^{2}u(\bar{x})+g(\bar{x})+r\langle\bar{y},h\rangle,

completing the proof. ∎

In the Riemannian or Lorentzian setting, in order for the area formula to be true, the Jacobian of a map should be modified with volume elements; that is,

J(Φr)=|det(DΦr)|detg¯detg,J(\Phi_{r})=|\det(D\Phi_{r})|\cdot\frac{\sqrt{-\det\bar{g}}}{\sqrt{\det g}},

where DΦrD\Phi_{r} is the usual tangent map of the coordinate map.

Lemma 3.3.

The invariant Jacobian of Φr\Phi_{r} is given by

J(Φr)=rn+1det(iju(x)+y,hij+δij/r).J(\Phi_{r})=r^{n+1}\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right).
Proof.

At a fixed point (x,y)(x,y), we pick an orthonormal basis {ei,ν}\{e_{i},\nu\} that spans T(x,y)(TΣ)T_{(x,y)}(T^{\bot}\Sigma), and a normal coordinate system {xi,y}\{x_{i},y\} such that xi=ei,y=ν\frac{\partial}{\partial x_{i}}=e_{i},\frac{\partial}{\partial y}=\nu at (x,y)(x,y). Then Φr(x,y)=X+ru+ryν\Phi_{r}(x,y)=X+r\nabla u+ry\nu. We now compute

Φrxi,ej=ei+r¯eiu+ry¯eiν,ej=δij+ruijryhij,\displaystyle\langle\frac{\partial\Phi_{r}}{\partial x_{i}},e_{j}\rangle=\langle e_{i}+r\bar{\nabla}_{e_{i}}\nabla u+{ry\bar{\nabla}_{e_{i}}\nu},e_{j}\rangle=\delta_{ij}+ru_{ij}-ryh_{ij},
Φry,ej=rν,ej=0,\displaystyle\langle\frac{\partial\Phi_{r}}{\partial y},e_{j}\rangle=\langle r\nu,e_{j}\rangle=0,
Φy,ν=rν,ν=r.\displaystyle\langle\frac{\partial\Phi}{\partial y},\nu\rangle=\langle r\nu,\nu\rangle=-r.

Note that we have used the fact that ¯eiej,ν=hijν,ν=hij\langle\bar{\nabla}_{e_{i}}e_{j},\nu\rangle=\langle-h_{ij}\nu,\nu\rangle=h_{ij} and that ¯eiν,ej=¯eiej,ν=hij\langle\bar{\nabla}_{e_{i}}\nu,e_{j}\rangle=-\langle\bar{\nabla}_{e_{i}}e_{j},\nu\rangle=-h_{ij}. Thus det(DΦr)=rdet(ruijryhij+δij)\det(D\Phi_{r})=-r\det(ru_{ij}-ryh_{ij}+\delta_{ij}), whence

J(Φr)=rn+1det(iju+y,hij+δij/r).J(\Phi_{r})=r^{n+1}\det(\nabla_{i}\nabla^{j}u+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r).

By tensorality the same formula holds in any coordinate system. ∎

Lemma 3.4.

We have for any (x,y)r(x,y)\in\mathcal{B}_{r}

det(iju(x)+y,hij+δij/r)(f1n1+1/r)n.\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right)\leq\left(f^{\frac{1}{n-1}}+1/r\right)^{n}.
Proof.

By definition, 2u+y,h+g/r0\nabla^{2}u+\langle y,h\rangle+g/r\geq 0 for any (x,y)r(x,y)\in\mathcal{B}_{r}. Therefore

det(iju+y,hij+δij/r)(Δu+y,Hn+1r)n(Δun+1r)n.\det\left(\nabla_{i}\nabla^{j}u+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right)\leq\left(\frac{\Delta u+\langle y,H\rangle}{n}+\frac{1}{r}\right)^{n}\leq\left(\frac{\Delta u}{n}+\frac{1}{r}\right)^{n}.

On the other hand, from Eq. 3.2 and the fact that |u|<1|\nabla u|<1, we have

Δu=nf1n1f1(|f|+f,u)nf1n1,\Delta u=nf^{\frac{1}{n-1}}-f^{-1}\left(|\nabla f|+\langle\nabla f,\nabla u\rangle\right)\leq nf^{\frac{1}{n-1}},

completing the proof. ∎

Proof of 1.1.

By area/coarea formula and previous lemmas, we have

|𝒜r|rn+1r(f1n1+1/r)n𝑑y𝑑μgΣ(f1n1+1/r)n𝑑μg.\frac{|\mathcal{A}_{r}|}{r^{n+1}}\leq\int_{\mathcal{B}_{r}}\left(f^{\frac{1}{n-1}}+1/r\right)^{n}dyd\mu_{g}\leq\int_{\Sigma}\left(f^{\frac{1}{n-1}}+1/r\right)^{n}d\mu_{g}.

Sending rr\to\infty, we derive

(3.6) C~n,τΣfnn1𝑑μg.\tilde{C}_{n,\tau}\leq\int_{\Sigma}f^{\frac{n}{n-1}}d\mu_{g}.

Combining this and Eq. 3.1, we conclude that

Σf+Σ|f|=n(Σfnn1)1n(Σfnn1)n1nnC~n,τ1n(Σfnn1)n1n.\int_{\partial\Sigma}f+\int_{\Sigma}|\nabla f|=n\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{1}{n}}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}}\geq n\tilde{C}_{n,\tau}^{\frac{1}{n}}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}}.

We finally write Cn,τ=nC~n,τ1n=nωn1n(n+1)1nτ1n(τ+τ21)1.C_{n,\tau}=n\tilde{C}_{n,\tau}^{\frac{1}{n}}=n\omega_{n}^{\frac{1}{n}}(n+1)^{-\frac{1}{n}}\tau^{-\frac{1}{n}}(\tau+\sqrt{\tau^{2}-1})^{-1}.

4. Proof of 1.2

Without the ‘mean convexity’ assumption, the union 𝒜r=xΣΛr(x)\mathcal{A}_{r}=\cap_{x\in\Sigma}\Lambda_{r}(x), with Λr(x)\Lambda_{r}(x) defined by Eq. 3.3, might as well be empty. We therefore construct uu by

(4.1) {div(fu)=nfnn1|f|2+f2H2, in Σ,u,ηg¯=1, along Σ.\begin{cases}\operatorname{div}(f\nabla u)=nf^{\frac{n}{n-1}}-\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}},&\text{ in }\Sigma,\\ \langle\nabla u,\eta\rangle_{\bar{g}}=1,&\text{ along }\partial\Sigma.\end{cases}

Since Eq. 1.2 is homogeneous in ff, we may assume

(4.2) Σf=Σnfnn1Σ|f|2+f2H2,\int_{\partial\Sigma}f=\int_{\Sigma}nf^{\frac{n}{n-1}}-\int_{\Sigma}\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}},

so that Eq. 4.1 admits a solution. We then modify Eq. 3.3 by

Λr(x)={pn,1:|(pX(x))|2<r24,r2pX(x),H1H(x)r2}\Lambda_{r}(x)=\left\{p\in\mathbb{R}^{n,1}:|(p-X(x))^{\top}|^{2}<\frac{r^{2}}{4},-\frac{r}{2}\leq\langle p-X(x),\|H\|^{-1}H(x)\rangle\leq\frac{r}{2}\right\}

and 𝒜r=xΣΛr(x)\mathcal{A}_{r}=\cap_{x\in\Sigma}\Lambda_{r}(x). Accordingly,

𝒰={(x,y):xΣΣ,yTxΣ,|u(x)|<12,12y,H1H(x)12},\displaystyle\mathcal{U}=\left\{(x,y):x\in\Sigma\setminus\partial\Sigma,y\in T^{\bot}_{x}\Sigma,|\nabla u(x)|<\frac{1}{2},-\frac{1}{2}\leq\langle y,\|H\|^{-1}H(x)\rangle\leq\frac{1}{2}\right\},
r={(x,y)𝒰:r2u(x)+rh(x),y+g(x)0},\displaystyle\mathcal{B}_{r}=\left\{(x,y)\in\mathcal{U}:r\nabla^{2}u(x)+r\langle h(x),y\rangle+g(x)\geq 0\right\},

and Φr(x,y)=X(x)+r(u(x)+y)\Phi_{r}(x,y)=X(x)+r(\nabla u(x)+y). Note that when H=0\|H\|=0, the notation H1H\|H\|^{-1}H simply means the normal vector ν\nu. Similar as in Section 3, we still have

  • The inclusion Φr(r)𝒜r\Phi_{r}(\mathcal{B}_{r})\supseteq\mathcal{A}_{r}; and

  • The Jacobian J(Φr)=rn+1det(iju(x)+y,hij+δij/r)J(\Phi_{r})=r^{n+1}\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right).

The proofs are almost identical.

Lemma 4.1.

We have asymptotic behavior

(4.3) lim infrrn1|𝒜r|C~n,τ,\liminf_{r\to\infty}r^{-n-1}{|\mathcal{A}_{r}|}\geq\tilde{C}_{n,\tau},

where C~n,τ=2n(n+1)1ωnτ1(τ+τ21)n\tilde{C}_{n,\tau}=2^{-n}(n+1)^{-1}\omega_{n}\tau^{-1}(\tau+\sqrt{\tau^{2}-1})^{-n}.

Proof.
Refer to caption
Figure 3. Left: each Λ~(x)\tilde{\Lambda}(x) contains two cones, the shaded area; Right: Λ~\tilde{\Lambda} contains at least 𝒞~\tilde{\mathcal{C}}, the shaded area.

The proof is similar to that of 3.1, with minor differences. Since the minimal height of unit normals is τ-\tau, in Fig. 3, a=(τ2,τ212,0,,0)a=(-\frac{\tau}{2},-\frac{\sqrt{\tau^{2}-1}}{2},0,\cdots,0). The vector ac\vec{ac} is parallel to ξ=(τ21,τ,0,,0)\xi=(\sqrt{\tau^{2}-1},\tau,0,\cdots,0). From this we derive b=(12τ,0,0)b=(-\frac{1}{2\tau},0\cdots,0) and c=(τ+τ212,ττ212,0,0)c=(\frac{-\tau+\sqrt{\tau^{2}-1}}{2},\frac{\tau-\sqrt{\tau^{2}-1}}{2},0\cdots,0). Thus 𝒞~\tilde{\mathcal{C}} has volume

|𝒞~|=ωn(ττ212)n12τ1n+12=2n(n+1)1ωnτ1(τ+τ21)n,|\tilde{\mathcal{C}}|=\omega_{n}\left(\frac{\tau-\sqrt{\tau^{2}-1}}{2}\right)^{n}\cdot\frac{1}{2\tau}\cdot\frac{1}{n+1}\cdot 2=2^{-n}(n+1)^{-1}\omega_{n}\tau^{-1}(\tau+\sqrt{\tau^{2}-1})^{-n},

from which it follows that

lim infrrn1|𝒜r||𝒞~|C~n,τ,\liminf_{r\to\infty}r^{-n-1}{|\mathcal{A}_{r}|}\geq|\tilde{\mathcal{C}}|\geq\tilde{C}_{n,\tau},

completing the proof. ∎

Lemma 4.2.

We have for any (x,y)r(x,y)\in\mathcal{B}_{r}

det(iju(x)+y,hij+δij/r)(f1n1+1/r)n.\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right)\leq\left(f^{\frac{1}{n-1}}+1/r\right)^{n}.
Proof.

By definition, 2u+y,h+g/r0\nabla^{2}u+\langle y,h\rangle+g/r\geq 0 for any (x,y)r(x,y)\in\mathcal{B}_{r}. Therefore

det(iju+y,hij+δij/r)(Δu+y,Hn+1r)n.\det\left(\nabla_{i}\nabla^{j}u+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right)\leq\left(\frac{\Delta u+\langle y,H\rangle}{n}+\frac{1}{r}\right)^{n}.

Moreover, we have

y,fHf,u\displaystyle\langle y,fH\rangle-\langle\nabla f,\nabla u\rangle 12fH+|f||u||f|2+f2H2|u|2+1/4\displaystyle\leq\frac{1}{2}f\|H\|+|\nabla f||\nabla u|\leq\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}}\sqrt{|\nabla u|^{2}+1/4}
|f|2+f2H21/4+1/4|f|2+f2H2.\displaystyle\leq\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}}\sqrt{1/4+1/4}\leq\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}}.

Combined with Eq. 4.1, it follows that

Δu+y,H=nf1n1f1(|f|2+f2H2+f,uy,fH)nf1n1,\Delta u+\langle y,H\rangle=nf^{\frac{1}{n-1}}-f^{-1}\left(\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}}+\langle\nabla f,\nabla u\rangle-\langle y,fH\rangle\right)\leq nf^{\frac{1}{n-1}},

giving the assertion. ∎

Proof of 1.2.

By area/coarea formula, we have

|𝒜r|rn+1rJ(Φr)rn+1𝑑y𝑑μgΣ(f1n1+1/r)n𝑑μg.\frac{|\mathcal{A}_{r}|}{r^{n+1}}\leq\int_{\mathcal{B}_{r}}\frac{J(\Phi_{r})}{r^{n+1}}dyd\mu_{g}\leq\int_{\Sigma}\left(f^{\frac{1}{n-1}}+1/r\right)^{n}d\mu_{g}.

Sending rr\to\infty, we derive Σfnn1C~n,τ\int_{\Sigma}f^{\frac{n}{n-1}}\geq\tilde{C}_{n,\tau}. In the view of Eq. 4.2, we compute

Σf+Σ|f|2+f2H2=nΣfnn1\displaystyle\int_{\partial\Sigma}f+\int_{\Sigma}\sqrt{|\nabla f|^{2}+f^{2}\|H\|^{2}}=n\int_{\Sigma}f^{\frac{n}{n-1}}
=n(Σfnn1)n1n(Σfnn1)1nnC~n,τ1n(Σfnn1)n1n.\displaystyle=n\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{1}{n}}\geq n\tilde{C}_{n,\tau}^{\frac{1}{n}}\left(\int_{\Sigma}f^{\frac{n}{n-1}}\right)^{\frac{n-1}{n}}.

Finally we write Cn,τ=nC~n,τ1n=n21(n+1)1nωn1nτ1n(τ+τ21)1.C_{n,\tau}=n\tilde{C}_{n,\tau}^{\frac{1}{n}}=n2^{-1}(n+1)^{-\frac{1}{n}}\omega_{n}^{\frac{1}{n}}\tau^{-\frac{1}{n}}(\tau+\sqrt{\tau^{2}-1})^{-1}.

5. Proof of 1.3

We fix such a section νΓ(TΣ)\nu\in\Gamma(T^{\bot}\Sigma), with |ν|2=1|\nu|^{2}=-1; that is, ν\nu is a unit timelike normal vector field along Σ\Sigma. At each point xΣx\in\Sigma, we have decompostion

TxΣ=Tx,1ΣTx,2Σ,T_{x}^{\bot}\Sigma=T_{x}^{\bot,1}\Sigma\oplus T_{x}^{\bot,2}\Sigma,

where Tx,2Σ=Span{ν(x)}T_{x}^{\bot,2}\Sigma=\operatorname{Span}\{\nu(x)\}. Note that Tx,1ΣT_{x}^{\bot,1}\Sigma is a spacelike subspace. Accordingly, we write a normal vector as y=y,1+y,2y=y^{\bot,1}+y^{\bot,2}. The equation we consider now is

(5.1) {div(fu)=mfmm1|f|2+f2(|H,1|2+H,22),ΣΣ,u,η=1,Σ.\begin{cases}\operatorname{div}(f\nabla u)=mf^{\frac{m}{m-1}}-\sqrt{|\nabla f|^{2}+f^{2}(|H^{\bot,1}|^{2}+\|H^{\bot,2}\|^{2})},&\Sigma\setminus\partial\Sigma,\\ \langle\nabla u,\eta\rangle=1,&\partial\Sigma.\end{cases}

We normalize by

(5.2) Σf=Σmfmm1Σ|f|2+f2(|H,1|2+H,22)\int_{\partial\Sigma}f=\int_{\Sigma}mf^{\frac{m}{m-1}}-\int_{\Sigma}\sqrt{|\nabla f|^{2}+f^{2}(|H^{\bot,1}|^{2}+\|H^{\bot,2}\|^{2})}

so that the PDE has a solution. The domains and codomains are

Λr(x)={pn,1:|(pX(x))|2+|(pX(x)),1|2<r24,r2pX(x),ν(x)r2},\displaystyle\Lambda_{r}(x)=\left\{p\in\mathbb{R}^{n,1}:|(p-X(x))^{\top}|^{2}+|(p-X(x))^{\bot,1}|^{2}<\frac{r^{2}}{4},-\frac{r}{2}\leq\langle p-X(x),\nu(x)\rangle\leq\frac{r}{2}\right\},
𝒜r=xΣΛr(x),\displaystyle\mathcal{A}_{r}=\cap_{x\in\Sigma}\Lambda_{r}(x),
𝒰={(x,y):xΣΣ,yTxΣ,|u(x)|2+|y,1|2<14,12y,2,ν12},\displaystyle\mathcal{U}=\left\{(x,y):x\in\Sigma\setminus\partial\Sigma,y\in T_{x}^{\bot}\Sigma,|\nabla u(x)|^{2}+|y^{\bot,1}|^{2}<\frac{1}{4},-\frac{1}{2}\leq\langle y^{\bot,2},\nu\rangle\leq\frac{1}{2}\right\},
r={(x,y)𝒰:r2u(x)+ry,h(x)+g(x)0}.\displaystyle\mathcal{B}_{r}=\{(x,y)\in\mathcal{U}:r\nabla^{2}u(x)+r\langle y,h(x)+g(x)\geq 0\rangle\}.

Finally we take Φr:TΣn,1:(x,y)X(x)+r(u(x)+y)\Phi_{r}:T^{\bot}\Sigma\to\mathbb{R}^{n,1}:(x,y)\mapsto X(x)+r(\nabla u(x)+y). As in Section 3,

  • The inclusion Φr(r)𝒜r\Phi_{r}(\mathcal{B}_{r})\supseteq\mathcal{A}_{r}; and

  • The Jacobian J(Φr)=rn+1det(iju(x)+y,hij+δij/r)J(\Phi_{r})=r^{n+1}\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right).

The proofs are almost identical.

Lemma 5.1.

We have asymptotic behavior

(5.3) lim infrrn1|𝒜r|C~n,τ,\liminf_{r\to\infty}r^{-n-1}{|\mathcal{A}_{r}|}\geq\tilde{C}_{n,\tau},

where C~n,τ=2n(n+1)1ωnτ1(τ+τ21)n\tilde{C}_{n,\tau}=2^{-n}(n+1)^{-1}\omega_{n}\tau^{-1}(\tau+\sqrt{\tau^{2}-1})^{-n}.

Proof.

We write n,1=TxΣTxΣ=(TxΣTx,1Σ)Tx,2Σ\mathbb{R}^{n,1}=T_{x}\Sigma\oplus T_{x}^{\bot}\Sigma=(T_{x}\Sigma\oplus T_{x}^{\bot,1}\Sigma)\oplus T_{x}^{\bot,2}\Sigma. If pΛr(x)p\in\Lambda_{r}(x), then

|(pX(x))|2+|(pX(x)),1|2<r24 and |(pX(x)),2|2r24.|(p-X(x))^{\top}|^{2}+|(p-X(x))^{\bot,1}|^{2}<\frac{r^{2}}{4}\quad\text{ and }\quad|(p-X(x))^{\bot,2}|^{2}\leq\frac{r^{2}}{4}.

If ξTxΣTx,1Σ\xi\in T_{x}\Sigma\oplus T_{x}^{\bot,1}\Sigma is a unit vector and pX=sξ+tνp-X=s\xi+t\nu, then |s|,|t|r2|s|,|t|\leq\frac{r}{2}, the shape of Λr(x)\Lambda_{r}(x) and 𝒜r\mathcal{A}_{r} would be exactly like those in Fig. 3. Thus the same conclusion holds as in 4.1. ∎

Lemma 5.2.

We have for any (x,y)r(x,y)\in\mathcal{B}_{r}

det(iju(x)+y,hij+δij/r)(f1m1+1/r)m.\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right)\leq\left(f^{\frac{1}{m-1}}+1/r\right)^{m}.
Proof.

By geometric-arithmetic inequality,

det(iju(x)+y,hij+δij/r)(Δu+y,Hm+1r)m.\det\left(\nabla_{i}\nabla^{j}u(x)+\langle y,h_{i}^{j}\rangle+\delta_{i}^{j}/r\right)\leq\left(\frac{\Delta u+\langle y,H\rangle}{m}+\frac{1}{r}\right)^{m}.

Now by Eq. 5.1, we have

Δu+y,H\displaystyle\Delta u+\langle y,H\rangle =mf1m1f1(|f|2+f2(|H,1|2+H,22)+f,ufH,y).\displaystyle=mf^{\frac{1}{m-1}}-f^{-1}(\sqrt{|\nabla f|^{2}+f^{2}(|H^{\bot,1}|^{2}+\|H^{\bot,2}\|^{2})}+\langle\nabla f,\nabla u\rangle-\langle fH,y\rangle).

On the other hand, by the definition of r\mathcal{B}_{r},

f,u+fH,y\displaystyle-\langle\nabla f,\nabla u\rangle+\langle fH,y\rangle =f,u+fH,1,y,1+fH,2,y,2\displaystyle=\langle\nabla f,\nabla u\rangle+\langle fH^{\bot,1},y^{\bot,1}\rangle+\langle fH^{\bot,2},y^{\bot,2}\rangle
|f||u|+|fH,1||y,1|+12fH,2\displaystyle\leq|\nabla f||\nabla u|+|fH^{\bot,1}||y^{\bot,1}|+\frac{1}{2}f\|H^{\bot,2}\|
|f|2+f2|H,1|2+f2H,22|u|2+|y,1|2+14\displaystyle\leq\sqrt{|\nabla f|^{2}+f^{2}|H^{\bot,1}|^{2}+f^{2}\|H^{\bot,2}\|^{2}}\sqrt{|\nabla u|^{2}+|y^{\bot,1}|^{2}+\frac{1}{4}}
|f|2+f2(|H,1|2+H,22).\displaystyle\leq\sqrt{|\nabla f|^{2}+f^{2}(|H^{\bot,1}|^{2}+\|H^{\bot,2}\|^{2})}.

This implies that Δu+y,Hmf1m1\Delta u+\langle y,H\rangle\leq mf^{\frac{1}{m-1}}, completing the proof. ∎

Proof of 1.3.

Using area/coarea formula, we have

|𝒜r|rn+1Σ(f1m1+1/r)m.\frac{|\mathcal{A}_{r}|}{r^{n+1}}\leq\int_{\Sigma}\left(f^{\frac{1}{m-1}}+1/r\right)^{m}.

Taking rr\to\infty, we obtain

C~n,τΣfmm1.\tilde{C}_{n,\tau}\leq\int_{\Sigma}f^{\frac{m}{m-1}}.

Taking Eq. 5.2 into consideration, we derive

mC~n,τ1m(Σfmm1)m1m\displaystyle m\tilde{C}_{n,\tau}^{\frac{1}{m}}\left(\int_{\Sigma}f^{\frac{m}{m-1}}\right)^{\frac{m-1}{m}} m(Σfmm1)1m(Σfmm1)m1m=mΣfmm1\displaystyle\leq m\left(\int_{\Sigma}f^{\frac{m}{m-1}}\right)^{\frac{1}{m}}\left(\int_{\Sigma}f^{\frac{m}{m-1}}\right)^{\frac{m-1}{m}}=m\int_{\Sigma}f^{\frac{m}{m-1}}
=Σf+Σ|f|2+f2(|H,1|2+H,22).\displaystyle=\int_{\partial\Sigma}f+\int_{\Sigma}\sqrt{|\nabla f|^{2}+f^{2}(|H^{\bot,1}|^{2}+\|H^{\bot,2}\|^{2})}.

Finally, we compute Cm,n,τ=mC~n,τ1m=m2nm(n+1)1mωn1mτ1m(τ+τ21)nmC_{m,n,\tau}=m\tilde{C}_{n,\tau}^{\frac{1}{m}}=m2^{-\frac{n}{m}}(n+1)^{-\frac{1}{m}}\omega_{n}^{\frac{1}{m}}\tau^{-\frac{1}{m}}(\tau+\sqrt{\tau^{2}-1})^{-\frac{n}{m}}. ∎

References

  • [Aub76] Thierry Aubin “Problèmes Isopérimétriques et Espaces de Sobolev” In Journal of Differential Geometry 11.4, 1976, pp. 573–598 DOI: 10.4310/jdg/1214433725
  • [BR33] E. F. Beckenbach and T. Rado “Subharmonic Functions and Surfaces of Negative Curvature” In Transactions of the American Mathematical Society 35.3, 1933, pp. 662 DOI: 10.2307/1989854
  • [Bre20] Simon Brendle “Sobolev Inequalities in Manifolds with Nonnegative Curvature”, 2020, pp. 1–25 arXiv: http://arxiv.org/abs/2009.13717
  • [Bre21] Simon Brendle “The Isoperimetric Inequality for a Minimal Submanifold in Euclidean Space” In Journal of the American Mathematical Society 34.2, 2021, pp. 595–603
  • [CG92] Jaigyoung Choe and Robert Gulliver “Isoperimetric Inequalities on Minimal Submanifolds of Space Forms” In Manuscripta Mathematica 77, 1992, pp. 169–189
  • [CG92a] Jaigyoung Choe and Robert Gulliver “The Sharp Isoperimetric Inequality for Minimal Surfaces with Radially Connected Boundary in Hyperbolic Space” In Inventiones Mathematicae 109.1, 1992, pp. 495–503 DOI: 10.1007/BF01232035
  • [Cab08] Xavier Cabré “Elliptic PDE’s in Probability and Geometry: Symmetry and Regularity of Solutions” In Discrete and Continuous Dynamical Systems 20.3, 2008, pp. 425–457 DOI: 10.3934/dcds.2008.20.425
  • [Cho05] Jaigyoung Choe “Isoperimetric Inequalities of Minimal Submanifolds” In Global theory of minimal surfaces, 2005, pp. 325–370
  • [Cro84] Christopher B. Croke “A Sharp Four Dimensional Isoperimetric Inequality” In Commentarii Mathematici Helvetici 59.1, 1984, pp. 187–192 DOI: 10.1007/BF02566344
  • [GS21] Mohammad Ghomi and Joel Spruck “Total Curvature and the Isoperimetric Inequality in Cartan-Hadamard Manifolds” arXiv, 2021 arXiv: http://arxiv.org/abs/1908.09814
  • [Gua] Pengfei Guan “Curvature Measures, Isoperimetric Type Inequalities and Fully Nonlinear Pdes”
  • [Kle92] Bruce Kleiner “An Isoperimetric Comparison Theorem” In Inventiones Mathematicae 108.1, 1992, pp. 37–47 DOI: 10.1007/BF02100598
  • [MS14] Sung Hong Min and Keomkyo Seo “Optimal Isoperimetric Inequalities for Complete Proper Minimal Submanifolds in Hyperbolic Space” In Journal fur die Reine und Angewandte Mathematik, 2014, pp. 203–214 DOI: 10.1515/crelle-2012-0119
  • [TW22] Chung-Jun Tsai and Kai-Hsiang Wang “An Isoperimetric-Type Inequality for Spacelike Submanifold in the Minkowski Space” In International Mathematics Research Notices 2022.1, 2022, pp. 128–139 DOI: 10.1093/imrn/rnaa084
  • [Wei26] André Weil “Sur Les Surfaces a Courbure Negative” In CR Acad. Sci. Paris 182.2, 1926, pp. 1069–1071
  • [Yau75] Shing-Tung Yau “Isoperimetric Constants and the First Eigenvalue of a Compact Riemannian Manifold” In Annales scientifiques de l’École normale supérieure 8.4, 1975, pp. 487–507 DOI: 10.24033/asens.1299