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

Limits of Solutions to a Parabolic Monge-Ampère Equation

John Loftin* and Mao-Pei Tsui
(November 13, 2007)

*Department of Mathematics and Computer Science ,

 Rugters University, Newark, NJ 07102

email: loftin@rutgers.edu

**Department of Mathematics, University of Toledo, Toledo, OH 43606,USA

   Taida Institute for Mathematical Sciences, Taipei, Taiwan

email: Mao-Pei.Tsui@Utoledo.edu

1 Introduction

In [10], we study solutions to the affine normal flow for an initial hypersurface n+1\mathcal{L}\subset\mathbb{R}^{n+1} which is a convex, properly embedded, noncompact hypersurface. The method we used was to consider an exhausting sequence i\mathcal{L}_{i} of smooth, strictly convex, compact hypersurfaces so that each i\mathcal{L}_{i} is contained in the convex hull of i+1\mathcal{L}_{i+1} for each ii, and so that i\mathcal{L}_{i}\to\mathcal{L} locally uniformly. If the compact i\mathcal{L}_{i} is the initial hypersurface, the affine normal flow i(t)\mathcal{L}_{i}(t) is well-defined for all time tt from 0 to the extinction time TiT_{i} [7]. Then for all positive tt, we define the affine normal flow for initial hypersurface \mathcal{L} as a limit (t)=limii(t)\mathcal{L}(t)=\lim_{i\to\infty}\mathcal{L}_{i}(t). Ben Andrews extensively studies the affine normal flow for compact initial hypersurfaces [1, 2].

The method of proof in [10] is to consider the support functions si=sis_{\mathcal{L}_{i}}=s_{i} and to take the limit as ii\to\infty. For each Yn+1Y\in\mathbb{R}^{n+1}, the support function is defined by

s(Y)=s(Y)=supxx,Y,s(Y)=s_{\mathcal{L}}(Y)=\sup_{x\in\mathcal{L}}\langle x,Y\rangle,

for ,\langle\cdot,\cdot\rangle the Euclidean inner produce on n+1\mathbb{R}^{n+1}. It is immediate that ss is a convex function of homogeneity one on n+1\mathbb{R}^{n+1}. The homogeneity property means that it suffices to study the behavior of ss when restricted to the unit sphere 𝕊nn+1\mathbb{S}^{n}\subset\mathbb{R}^{n+1}. Also, ss restricted to an affine hyperplane not touching the origin in n+1\mathbb{R}^{n+1} determines ss on a half-space of n+1\mathbb{R}^{n+1}. We consider ss in this setting primarily: If Y=(y,1)Y=(y,-1) for yny\in\mathbb{R}^{n}, then ss evolves under the affine normal flow by

st=(det2syiyj)1n+2.\frac{\partial s}{\partial t}=-\left(\det\frac{\partial^{2}s}{\partial y^{i}\partial y^{j}}\right)^{-\frac{1}{n+2}}. (1.1)

Note that this setting of considering the restriction of ss to the hyperplane {yn+1=1}\{y^{n+1}=1\} has its roots in the Minkowski problem (see Cheng-Yau [4]).

In the present paper, we consider our previous result primarily the point of view of Equation (1.1)—in other words, from more of a classical PDE point of view as opposed to the largely tensorial point of view in [10]. Also, to the extent possible, we phrase the proofs in analytic terms, and try not to rely too much on the affine geometry. In particular, consider the support function sis_{i} of i\mathcal{L}_{i}. Then as ii\to\infty, si(Y)s_{i}(Y) increases to the limit s(Y)s(Y) for all Yn+1Y\in\mathbb{R}^{n+1} (this follows by the exhaustion property of i\mathcal{L}_{i}\to\mathcal{L}). The noncompactness of \mathcal{L} implies that ss is equal to ++\infty on at least a half-space of n+1\mathbb{R}^{n+1}. Let 𝒟(s)\mathcal{D}^{\circ}(s) be the largest open subset of n+1\mathbb{R}^{n+1} on which s<s<\infty. (𝒟(s)\mathcal{D}^{\circ}(s) is then the interior of the domain of ss, which is defined by 𝒟(s)={Y:s(Y)<+}\mathcal{D}(s)=\{Y:s(Y)<+\infty\}.) Since 𝒟(s)\mathcal{D}^{\circ}(s) is contained in an open half-space of n+1\mathbb{R}^{n+1}, we may (by choosing new coordinates if necessary) restrict to the affine hyperplane {Y=(y,1):yn}\{Y=(y,-1):y\in\mathbb{R}^{n}\} and consider the limit siss_{i}\nearrow s.

We make the following nondegeneracy assumptions about \mathcal{L} and thus ss. First, assume that \mathcal{L} does not contain any lines. This is equivalent to

𝒟(s)\mathcal{D}^{\circ}(s)\neq\emptyset (1.2)

(see e.g. Rockafellar [12]). Also assume that \mathcal{L} is a hypersurface, and not a lower-dimensional set. So, in particular, the convex hull ^\hat{\mathcal{L}} has nonempty interior, and thus contains a small ball Bϵ(P)B_{\epsilon}(P). Thus s=s=s^sBϵ(P)s=s_{\mathcal{L}}=s_{\hat{\mathcal{L}}}\geq s_{B_{\epsilon}(P)}, and there are Pn+1P\in\mathbb{R}^{n+1} and ϵ>0\epsilon>0 so that for all Yn+1Y\in\mathbb{R}^{n+1},

s(Y)ϵ|Y|+P,Ys(Y)\geq\epsilon|Y|+\langle P,Y\rangle (1.3)

For Y=(y,1)Y=(y,-1), this assumption becomes that there are ϵ>0\epsilon>0, pnp\in\mathbb{R}^{n} and cRc\in R so that for all yny\in\mathbb{R}^{n},

s(y)ϵ|y|2+1+p,ycs(y)\geq\epsilon\sqrt{|y|^{2}+1}+\langle p,y\rangle-c (1.4)

Also note that equation (1.3) may be computed using the following useful transformation law for the support function: If A𝐆𝐋(n+1,)A\in\mathbf{GL}(n+1,\mathbb{R}) and bn+1b\in\mathbb{R}^{n+1}, then

sA+b(Y)=s(AY)+b,Y.s_{A\mathcal{L}+b}(Y)=s_{\mathcal{L}}(A^{\top}Y)+\langle b,Y\rangle. (1.5)

This rule is particularly useful, since the affine normal flow is invariant under all affine volume-preserving maps of n+1\mathbb{R}^{n+1}. Note also that (1.5) is equivalent to a projective transformation of ss when restricted to {yn+1=1}\{y^{n+1}=-1\}.

In terms of the support function functions, we consider siss_{i}\nearrow s, where the si:n+1s_{i}\!:\mathbb{R}^{n+1}\to\mathbb{R} are all convex functions of homogeneity one on n+1{0}\mathbb{R}^{n+1}\setminus\{0\} which are smooth and strictly convex on each affine hyperplane in n+1\mathbb{R}^{n+1} which does not pass through the origin. Then the affine normal flow si(t)s_{i}(t) may be defined by solving (1.1) on affine coordinate hyperplanes {yi=±1}\{y^{i}=\pm 1\} and patching together the solutions. More simply, si|𝕊ns_{i}|_{\mathbb{S}^{n}} solves a parabolic equation, and thus we have existence and uniqueness for a short time (as noted by Chow [7] originally). Then we let siss_{i}\to s pointwise everywhere in n+1\mathbb{R}^{n+1}, given the nondegeneracy assumptions (1.2) and (1.4) and as well that the interior of the domain 𝒟(s)\mathcal{D}^{\circ}(s) is contained in the half-space {yn+1<0}\{y^{n+1}<0\}.

Now for the affine normal flow, si(t)s(t)s_{i}(t)\nearrow s(t) as ii\to\infty. On 𝒟(s)\mathcal{D}^{\circ}(s), this is an increasing limit of smooth strictly convex functions (and so s(t)s(t) is Lipschitz a priori). Our problem is then to examine which properties of the solutions si(t)s_{i}(t) to (1.1) survive in the limit si(t)s(t)s_{i}(t)\nearrow s(t) on 𝒟(s)\mathcal{D}^{\circ}(s). This will determine the regularity properties of s(t)s(t). In particular, there are locally uniform spacelike C0,1C^{0,1} estimates on sis_{i} on 𝒟(s)\mathcal{D}^{\circ}(s) just by convexity. Uniform spacelike C2C^{2} and ellipticity estimates follow by a global speed estimate of Andrews [2] which survives in the limit as siss_{i}\to s and a local Pogorelov-type estimate of Gutiérrez-Huang [8]. We also use a barrier due to Calabi [3] to ensure we can apply Gutiérrez-Huang’s estimate to get locally uniform spacelike C2C^{2} estimates on sis_{i} for all positive tt. Then Evans-Krylov theory applies to get locally uniform parabolic C2+α,1+α/2C^{2+\alpha,1+\alpha/2} estimates and standard bootstrapping implies local CC^{\infty} convergence of siss_{i}\to s for positive time tt.

There is also an important estimate of Ben Andrews [1] on |C|2|C|^{2} associated to sis_{i} the support function a compact, smooth, strictly convex hypersurfaces i\mathcal{L}_{i}, for a tensor CC called the cubic form. This estimate shows that for any ancient solution to the affine normal flow, |C|2=0|C|^{2}=0, which implies by a classical theorem of Berwald that \mathcal{L} is a quadratic hypersurface. In Section 7 below, we reproduce this classical theorem from the point of view of the support function ss.

The first author is grateful for the support of the NSF under Grant DMS0405873, and to the organizers of the 2007 Conference on Geometric Analysis in Taipei, for the opportunity to speak and for their kind hospitality during the conference. The second author wishes to express his gratitude to Taida Institute for Mathematical Sciences for providing an excellent research environment and the support of C.S. Lin and Y.I. Lee. We would both like to thank D.H. Phong for his support and for his suggestion to write up our results in this way.

2 Support Function

In this section, we compute some of the basic quantities of affine differential geometry in terms of the support function ss. In the end of this section, we show that (1.1) is equivalent to affine normal flow.

Let FF be a smooth embedding of a strictly convex hypersurface in terms of an extended Gauss map. This means F=F(Y)F=F(Y) for any vector YY equal to a negative multiple of the inward-pointing unit normal vector ν\nu to the image of FF. So FF is a function from an collection of open rays in n+1{0}\mathbb{R}^{n+1}\setminus\{0\} to n+1\mathbb{R}^{n+1} which is homogeneous of degree 0. In particular, we have

s(Y)=F,Y.s(Y)=\langle F,Y\rangle.

The affine normal ξ\xi is a transverse vector field to the image of FF which is invariant under the action of all volume-preserving affine maps in n+1\mathbb{R}^{n+1}. We recall the basic tensors and structure equations of affine differential geometry: For each yy in the domain of FF, consider the basis F1,,Fn,ξF_{1},\dots,F_{n},\xi of n+1\mathbb{R}^{n+1}, write the derivatives of these basis elements in terms of the same basis:

Fij\displaystyle F_{ij} =\displaystyle= (Γijk+Cijk)Fk+gijξ,\displaystyle(\Gamma_{ij}^{k}+C_{ij}^{k})F_{k}+g_{ij}\xi,
ξi\displaystyle\xi_{i} =\displaystyle= AijFj.\displaystyle-A_{i}^{j}F_{j}.

Here gijg_{ij} the affine metric, or affine second fundamental form, is positive definite for strictly convex hypersurfaces; Γijk\Gamma_{ij}^{k} is its Levi-Civita connection; CijkC_{ij}^{k} is the cubic form; and AijA_{i}^{j} is the affine curvature, or affine shape operator.

Now we derive the formula for the cubic form CijkC^{k}_{ij} in terms of the support function ss:

Under the extended Gauss map, the inward-pointing Euclidean unit normal ν\nu satisfies

ν=Y|Y|=(y1,,yn,1)1+|y|2,\nu=-\frac{Y}{|Y|}=\frac{(-y^{1},\dots,-y^{n},1)}{\sqrt{1+|y|^{2}}}, (2.1)

and the (Euclidean) second fundamental form is given by

hij=sij1+|y|2,hij=1+|y|2sij.h_{ij}=\frac{s_{ij}}{\sqrt{1+|y|^{2}}},\qquad h^{ij}=\sqrt{1+|y|^{2}}s^{ij}. (2.2)

The scalar function ϕ\phi is defined to be (dethij)1n+2(detg¯ij)1n+2(\det h_{ij})^{\frac{1}{n+2}}(\det\bar{g}_{ij})^{-\frac{1}{n+2}} for g¯ij=Fi,Fj\bar{g}_{ij}=\langle F_{i},F_{j}\rangle the induced metric from Euclidean n+1\mathbb{R}^{n+1}. We compute (using the formula for g¯ij\bar{g}_{ij} below)

ϕ=(1+|y|2)12D1n+2,\phi=(1+|y|^{2})^{-\frac{1}{2}}D^{-\frac{1}{n+2}},

for D=detsijD=\det s_{ij}, and

ϕk=yk(1+|y|2)32D1n+21n+2(1+|y|2)12D1n+2(lnD)k,\phi_{k}=-y^{k}(1+|y|^{2})^{-\frac{3}{2}}D^{-\frac{1}{n+2}}-\frac{1}{n+2}(1+|y|^{2})^{-\frac{1}{2}}D^{-\frac{1}{n+2}}(\ln D)_{k},

where (lnD)k=spqspqk(\ln D)_{k}=s^{pq}s_{pqk}. We also define the vector field ZiZ^{i} by

Zi=hikϕk=1+|y|2sik[yk(1+|y|2)32D1n+21n+2(1+|y|2)12D1n+2(lnD)k]=D1n+2[(1+|y|2)1sikyk+1n+2sik(lnD)k]\begin{split}&Z^{i}\\ =&-h^{ik}\phi_{k}\\ =&-\sqrt{1+|y|^{2}}s^{ik}[-y^{k}(1+|y|^{2})^{-\frac{3}{2}}D^{-\frac{1}{n+2}}-\frac{1}{n+2}(1+|y|^{2})^{-\frac{1}{2}}D^{-\frac{1}{n+2}}(\ln D)_{k}]\\ =&D^{-\frac{1}{n+2}}[(1+|y|^{2})^{-1}s^{ik}y^{k}+\frac{1}{n+2}s^{ik}(\ln D)_{k}]\end{split} (2.3)
F=(s1,,sn,(slyl)s),F=(s_{1},\dots,s_{n},(s_{l}y^{l})-s),
Fi=(s1i,,sni,sliyl)F_{i}=(s_{1i},\dots,s_{ni},s_{li}y^{l}) (2.4)
Fij=(s1ij,,snij,slijyl+sij)F_{ij}=(s_{1ij},\dots,s_{nij},s_{lij}y^{l}+s_{ij}) (2.5)

In terms of the scalar function ϕ\phi and the vector field ZiZ^{i} defined above, we define the affine normal ξ\xi as

ξ=ϕν+ZiFi=(1+|y|2)12D1n+2(y1,,yn,1)1+|y|2+D1n+2[(1+|y|2)1sikyk+1n+2sik(lnD)k](s1i,,sni,sliyl)=D1n+2(1+|y|2)1(y1,,yn,1)+D1n+2(1+|y|2)1(y1,,yn,|y|2)+D1n+2n+2((lnD)1,,(lnD)n,(lnD)iyi)=1n+2D1n+2((lnD)1,,(lnD)n,(n+2)+(lnD)iyi).\begin{split}&\xi\\ =\ &\phi\nu+Z^{i}F_{i}\\ =\ &(1+|y|^{2})^{-\frac{1}{2}}D^{-\frac{1}{n+2}}\cdot\frac{(-y^{1},\dots,-y^{n},1)}{\sqrt{1+|y|^{2}}}\\ +\ &D^{-\frac{1}{n+2}}[(1+|y|^{2})^{-1}s^{ik}y^{k}+\frac{1}{n+2}s^{ik}(\ln D)_{k}](s_{1i},\dots,s_{ni},s_{li}y^{l})\\ =\ &D^{-\frac{1}{n+2}}(1+|y|^{2})^{-1}\cdot(-y^{1},\dots,-y^{n},1)\\ +\ &D^{-\frac{1}{n+2}}(1+|y|^{2})^{-1}(y^{1},\dots,y^{n},|y|^{2})\\ +\ &\frac{D^{-\frac{1}{n+2}}}{n+2}((\ln D)_{1},\dots,(\ln D)_{n},(\ln D)_{i}y^{i})\\ =\ &\frac{1}{n+2}D^{-\frac{1}{n+2}}((\ln D)_{1},\dots,(\ln D)_{n},(n+2)+(\ln D)_{i}y^{i}).\\ \end{split} (2.6)

The affine normal ξ\xi is invariant under volume-preserving affine actions on n+1\mathbb{R}^{n+1}. The affine metric (also called the affine second fundamental form) gijg_{ij} is invariant under the same group, and is given by gij=ϕ1hijg_{ij}=\phi^{-1}h_{ij}. So compute

gij=D1n+2sij.g_{ij}=D^{\frac{1}{n+2}}s_{ij}.

In terms of s=s(y,1)=s(Y)s=s(y,-1)=s(Y), the embedding FF is given by

F=(s1,,sn,(siyi)s),F=(s_{1},\dots,s_{n},(s_{i}y^{i})-s),
gij=(detsk)1n+2sijg_{ij}=(\det s_{k\ell})^{\frac{1}{n+2}}s_{ij}
mgij=(detsk)1n+2(1n+2spqspqmsij+sijm)\partial_{m}g_{ij}=(\det s_{k\ell})^{\frac{1}{n+2}}(\frac{1}{n+2}s^{pq}s_{pqm}s_{ij}+s_{ijm})
gij=(detsk)1n+2sijg^{ij}=(\det s_{k\ell})^{-\frac{1}{n+2}}s^{ij}
Γijk=12gkl(igj+jgigij)=12skl(1n+2smpsmpisjl+sjli+1n+2smpsmpjsil+silj1n+2smpsmplsijsijl)=12(1n+2smpsmpjδik+1n+2smpsmpiδjk+sksij1n+2sksmpsmpsij)=12(1n+2(lnD)jδik+1n+2(lnD)iδjk+sksij1n+2sk(lnD)sij),\begin{split}&\Gamma^{k}_{ij}\\ =\ &{\textstyle\frac{1}{2}}g^{kl}(\partial_{i}g_{j\ell}+\partial_{j}g_{i\ell}-\partial_{\ell}g_{ij})\\ =\ &{\textstyle\frac{1}{2}}s^{kl}(\frac{1}{n+2}s^{mp}s_{mpi}s_{jl}+s_{jli}+\frac{1}{n+2}s^{mp}s_{mpj}s_{il}+s_{ilj}-\frac{1}{n+2}s^{mp}s_{mpl}s_{ij}-s_{ijl})\\ =\ &\frac{1}{2}(\frac{1}{n+2}s^{mp}s_{mpj}\delta^{k}_{i}+\frac{1}{n+2}s^{mp}s_{mpi}\delta^{k}_{j}+s^{k\ell}s_{ij\ell}-\frac{1}{n+2}s^{k\ell}s^{mp}s_{mp\ell}s_{ij})\\ =\ &\frac{1}{2}(\frac{1}{n+2}(\ln D)_{j}\delta^{k}_{i}+\frac{1}{n+2}(\ln D)_{i}\delta^{k}_{j}+s^{k\ell}s_{ij\ell}-\frac{1}{n+2}s^{k\ell}(\ln D)_{\ell}s_{ij}),\end{split} (2.7)

where we define D=detsijD=\det s_{ij}. Now compute the metric induced from the Euclidean metric g¯ij\bar{g}_{ij}.

g¯ij\displaystyle\bar{g}_{ij} =\displaystyle= Fyi,Fyj\displaystyle\left\langle\frac{\partial F}{\partial y^{i}},\frac{\partial F}{\partial y^{j}}\right\rangle
=\displaystyle= k,l=1n2syiyk(ykyl+δkl)2syjyl,\displaystyle\sum_{k,l=1}^{n}\frac{\partial^{2}s}{\partial y^{i}\partial y^{k}}(y^{k}y^{l}+\delta^{kl})\frac{\partial^{2}s}{\partial y^{j}\partial y^{l}},
g¯ij\displaystyle\bar{g}^{ij} =\displaystyle= k,l=1nsik(ykyl1+|y|2+δkl)slj,where smn is the inverse of sij,\displaystyle\sum_{k,l=1}^{n}s^{ik}(-\frac{y^{k}y^{l}}{1+|y|^{2}}+\delta_{kl})s^{lj},\text{where }s^{mn}\text{ is the inverse of }s_{ij},
detg¯ij\displaystyle\det{\bar{g}_{ij}} =\displaystyle= det(2syiyk)det(ykyl+δkl)det(2syjyl)\displaystyle\det\left(\frac{\partial^{2}s}{\partial y^{i}\partial y^{k}}\right)\det(y^{k}y^{l}+\delta^{kl})\det\left(\frac{\partial^{2}s}{\partial y^{j}\partial y^{l}}\right)
=\displaystyle= (1+|y|2)det(2syiyj)2.\displaystyle(1+|y|^{2})\det\left(\frac{\partial^{2}s}{\partial y^{i}\partial y^{j}}\right)^{2}.

Recall that ξ=ϕν+ZkFk\xi=\phi\nu+Z^{k}F_{k}, hij=ϕgijh_{ij}=\phi g_{ij} and

Fij=gij(ϕν+ZkFk)+(Γijk+Cijk)Fk.F_{ij}=g_{ij}(\phi\nu+Z^{k}F_{k})+(\Gamma^{k}_{ij}+C^{k}_{ij})F_{k}.

So

Fij,Fl=gijZkg¯kl+Γijkg¯kl+Cijkg¯kl,\langle F_{ij},F_{l}\rangle=g_{ij}Z^{k}\overline{g}_{kl}+\Gamma^{k}_{ij}\overline{g}_{kl}+C^{k}_{ij}\overline{g}_{kl},
Fij,Flg¯lm=gijZm+Γijm+Cijm,\langle F_{ij},F_{l}\rangle\overline{g}^{lm}=g_{ij}Z^{m}+\Gamma^{m}_{ij}+C^{m}_{ij},
Cijm=Fij,Flg¯lmgijZmΓijmC^{m}_{ij}=\langle F_{ij},F_{l}\rangle\overline{g}^{lm}-g_{ij}Z^{m}-\Gamma^{m}_{ij}

and, lowering the index by the affine metric Cijk=CijlglkC_{ijk}=C_{ij}^{l}g_{lk},

Cijk=Fij,Flg¯lmgmkgijZmgmkΓijmgmk.C_{ijk}=\langle F_{ij},F_{l}\rangle\overline{g}^{lm}g_{mk}-g_{ij}Z^{m}g_{mk}-\Gamma^{m}_{ij}g_{mk}. (2.8)

First, we compute

gijZk\displaystyle-g_{ij}Z^{k}
=\displaystyle=\ (detsk)1n+2sij(detsrs)1n+2[(1+|y|2)1sklyl+1n+2skl(p,qspqspql)]\displaystyle-(\det s_{k\ell})^{\frac{1}{n+2}}s_{ij}(\det s_{rs})^{-\frac{1}{n+2}}[(1+|y|^{2})^{-1}s^{kl}y^{l}+\frac{1}{n+2}s^{kl}(\sum_{p,q}s^{pq}s_{pql})]
=\displaystyle=\ sij[(1+|y|2)1sklyl+1n+2skl(lnD)l]\displaystyle-s_{ij}[(1+|y|^{2})^{-1}s^{kl}y^{l}+\frac{1}{n+2}s^{kl}(\ln D)_{l}]

So

gijZlglk=sij[(1+|y|2)1slmym+1n+2slm(p,qspqspqm)](detsrs)1n+2slk=sij(detsrs)1n+2[(1+|y|2)1yk+1n+2(p,qspqspqk)]=(detsrs)1n+2[sij(1+|y|2)1yk+sijn+2(lnD)k]\begin{split}&-g_{ij}Z^{l}g_{lk}\\ =\ &-s_{ij}[(1+|y|^{2})^{-1}s^{lm}y^{m}+\frac{1}{n+2}s^{lm}(\sum_{p,q}s^{pq}s_{pqm})](\det s_{rs})^{\frac{1}{n+2}}s_{lk}\\ =\ &-s_{ij}(\det s_{rs})^{\frac{1}{n+2}}[(1+|y|^{2})^{-1}y^{k}+\frac{1}{n+2}(\sum_{p,q}s^{pq}s_{pqk})]\\ =\ &-(\det s_{rs})^{\frac{1}{n+2}}[s_{ij}(1+|y|^{2})^{-1}y^{k}+\frac{s_{ij}}{n+2}(\ln D)_{k}]\\ \end{split} (2.9)

Now, we compute

g¯lmgmk\displaystyle\overline{g}^{lm}g_{mk}
=\displaystyle=\ slp(ypyq(1+y2)1+δpq)sqmD1n+2smk\displaystyle s^{lp}(-y^{p}y^{q}(1+y^{2})^{-1}+\delta_{pq})s^{qm}D^{\frac{1}{n+2}}s_{mk}
=\displaystyle=\ D1n+2(slpypyk(1+y2)1+slk)\displaystyle D^{\frac{1}{n+2}}(-s^{lp}y^{p}y^{k}(1+y^{2})^{-1}+s^{lk})

and

Fij,Fl\displaystyle\langle F_{ij},F_{l}\rangle
=\displaystyle=\ (s1ij,,snij,srijyr+sij),(s1k,,snk,smkym)\displaystyle\langle(s_{1ij},\dots,s_{nij},s_{rij}y^{r}+s_{ij}),(s_{1k},\dots,s_{nk},s_{mk}y^{m})\rangle
=\displaystyle=\ pspijspl+srijyrsmlym+sijsmlym.\displaystyle\sum_{p}s_{pij}s_{pl}+s_{rij}y^{r}s_{ml}y^{m}+s_{ij}s_{ml}y^{m}.

So

Fij,Flg¯lmgmk=D1n+2(pspijspl+srijyrsmlym+sijsmlym)(slqyqyk(1+y2)1+slk)=D1n+2(pspijypyk(1+y2)1srij|y|2(1+y2)1yryksij|y|2(1+y2)1yk+skij+srijyryk+sijyk))=D1n+2(pspijypyksij|y|2(1+y2)1yk+skij+srijyryk+sijyk))\begin{split}&\langle F_{ij},F_{l}\rangle\overline{g}^{lm}g_{mk}\\ =\ &D^{\frac{1}{n+2}}(\sum_{p}s_{pij}s_{pl}+s_{rij}y^{r}s_{ml}y^{m}+s_{ij}s_{ml}y^{m})(-s^{lq}y^{q}y^{k}(1+y^{2})^{-1}+s^{lk})\\ =\ &D^{\frac{1}{n+2}}(-\sum_{p}s_{pij}y^{p}y^{k}(1+y^{2})^{-1}-s_{rij}|y|^{2}(1+y^{2})^{-1}y^{r}y^{k}-s_{ij}|y|^{2}(1+y^{2})^{-1}y^{k}\\ +\ &s_{kij}+s_{rij}y^{r}y^{k}+s_{ij}y^{k}))\\ =\ &D^{\frac{1}{n+2}}(-\sum_{p}s_{pij}y^{p}y^{k}-s_{ij}|y|^{2}(1+y^{2})^{-1}y^{k}+s_{kij}+s_{rij}y^{r}y^{k}+s_{ij}y^{k}))\\ \end{split} (2.10)
Γijmgmk=12(1n+2(lnD)jδim+1n+2(lnD)iδjm+smsij1n+2sm(lnD)sij)D1n+2smk=D1n+22(1n+2(lnD)jsik+1n+2(lnD)isjk+sijk1n+2(lnD)ksij)\begin{split}&\Gamma^{m}_{ij}g_{mk}\\ =\ &\frac{1}{2}(\frac{1}{n+2}(\ln D)_{j}\delta^{m}_{i}+\frac{1}{n+2}(\ln D)_{i}\delta^{m}_{j}+s^{m\ell}s_{ij\ell}-\frac{1}{n+2}s^{m\ell}(\ln D)_{\ell}s_{ij})D^{\frac{1}{n+2}}s_{mk}\\ =\ &\frac{D^{\frac{1}{n+2}}}{2}(\frac{1}{n+2}(\ln D)_{j}s_{ik}+\frac{1}{n+2}(\ln D)_{i}s_{jk}+s_{ijk}-\frac{1}{n+2}(\ln D)_{k}s_{ij})\end{split} (2.11)

From (LABEL:C1), (LABEL:C2), (2.11) and (2.8), we have

Cijk=D1n+2(pspijypyksij|y|2(1+y2)1yk+skij+srijyryk+sijyk))(detsrs)1n+2[sij(1+|y|2)1yk+sijn+2(lnD)k]D1n+22(1n+2(lnD)jsik+1n+2(lnD)isjk+sijk1n+2(lnD)ksij)=D1n+2[12sijk12(n+2)ski(lnD)j12(n+2)skj(lnD)isij2(n+2)(lnD)k]\begin{split}&C_{ijk}\\ =\ &D^{\frac{1}{n+2}}\Big{(}-\sum_{p}s_{pij}y^{p}y^{k}-s_{ij}|y|^{2}(1+y^{2})^{-1}y^{k}\\ +\ &s_{kij}+s_{rij}y^{r}y^{k}+s_{ij}y^{k})\Big{)}-(\det s_{rs})^{\frac{1}{n+2}}[s_{ij}(1+|y|^{2})^{-1}y^{k}+\frac{s_{ij}}{n+2}(\ln D)_{k}]\\ -\ &\frac{D^{\frac{1}{n+2}}}{2}(\frac{1}{n+2}(\ln D)_{j}s_{ik}+\frac{1}{n+2}(\ln D)_{i}s_{jk}+s_{ijk}-\frac{1}{n+2}(\ln D)_{k}s_{ij})\\ =\ &D^{\frac{1}{n+2}}\Big{[}\frac{1}{2}s_{ijk}-\frac{1}{2(n+2)}s_{ki}(\ln D)_{j}-\frac{1}{2(n+2)}s_{kj}(\ln D)_{i}-\frac{s_{ij}}{2(n+2)}(\ln D)_{k}\Big{]}\end{split} (2.12)

Now we prove that (1.1) is equivalent to the affine normal flow.

Proposition 2.1

The affine normal flow

tF=ξ\frac{\partial}{\partial t}F=\xi

is equivalent to the evolution of the support function

st=(det2syiyj)1n+2.\frac{\partial s}{\partial t}=-\left(\det\frac{\partial^{2}s}{\partial y^{i}\partial y^{j}}\right)^{-\frac{1}{n+2}}.

Proof We first compute sts_{t} from Ft=ξF_{t}=\xi: Recall that ν=Y|Y|=(y1,,yn,1)1+|y|2\nu=-\frac{Y}{|Y|}=\frac{(-y^{1},\dots,-y^{n},1)}{\sqrt{1+|y|^{2}}} which is independent of time in our coordinate system, since tν=0{\partial_{t}}\nu=0 (see [10]). Using the definition s=F,Ys=\langle F,Y\rangle, ξ=(1+|y|2)12D1n+2ν+ZiFi\xi=(1+|y|^{2})^{-\frac{1}{2}}D^{-\frac{1}{n+2}}\nu+Z^{i}F_{i} and tY=0{\partial_{t}}Y=0, we have

ts\displaystyle{\partial_{t}}s =\displaystyle= tF,Y+F,tY\displaystyle\langle{\partial_{t}}F,Y\rangle+\langle F,{\partial_{t}}Y\rangle
=\displaystyle= det(1+|y|2)12D1n+2ν+ZiFi,(1+|y|2)12ν\displaystyle\left\langle\det(1+|y|^{2})^{-\frac{1}{2}}D^{-\frac{1}{n+2}}\nu+Z^{i}F_{i},-(1+|y|^{2})^{\frac{1}{2}}\nu\right\rangle
=\displaystyle= D1n+2.\displaystyle-D^{-\frac{1}{n+2}}.

For good measure, we also compute FtF_{t} from st=D1n+2s_{t}=-D^{-\frac{1}{n+2}}: Recall that the position function FF can be expressed by the support function

F=(s1,,sn,(slyl)s).F=(s_{1},\dots,s_{n},(s_{l}y^{l})-s).

Recall D=det(2syiyj)D=\det(\frac{\partial^{2}s}{\partial y^{i}\partial y^{j}}). Note that

st\displaystyle s_{t} =\displaystyle= D1n+2,\displaystyle-D^{-\frac{1}{n+2}},
sit\displaystyle s_{it} =\displaystyle= sti=(D1n+2)i\displaystyle s_{ti}=(-D^{-\frac{1}{n+2}})_{i}
=\displaystyle= [1n+2(lnD)iD1n+2]\displaystyle\left[\frac{1}{n+2}(\ln D)_{i}D^{-\frac{1}{n+2}}\right]

Compute

tF=(st1,,stn,stlylst)=1n+2D1n+2((lnD)1,,(lnD)n,(lnD)lyl+n+2).\begin{split}\frac{\partial}{\partial t}F&=(s_{t1},\dots,s_{tn},s_{tl}y^{l}-s_{t})\\ &=\frac{1}{n+2}D^{-\frac{1}{n+2}}((\ln D)_{1},\dots,(\ln D)_{n},(\ln D)_{l}y^{l}+n+2).\end{split} (2.13)

Recall that

ξ=1n+2D1n+2((lnD)1,,(lnD)n,(n+2)+(lnD)iyi)\xi=\frac{1}{n+2}D^{-\frac{1}{n+2}}((\ln D)_{1},\dots,(\ln D)_{n},(n+2)+(\ln D)_{i}y^{i})

from (2.6). Therefore tF=ξ\frac{\partial}{\partial t}F=\xi.

\Box

3 Andrews’s Speed Estimate

In this section, we repeat, for the reader’s convenience, our version of a speed estimate of Andrews [2].

Proposition 3.1

Let ss be the support function of a smooth strictly convex compact hypersurface evolving under affine normal flow. If s(Y,t)r>0s(Y,t)\geq r>0 for all Y𝕊nY\in\mathbb{S}^{n} and t[0,T]t\in[0,T], then

|ts|(C+Ctn2n+2)s|\partial_{t}s|\leq\left(C+C^{\prime}t^{-\frac{n}{2n+2}}\right)s

on 𝕊n×[0,T]\mathbb{S}^{n}\times[0,T], where CC and CC^{\prime} are constants only depending on rr and nn.

Proof Consider the function

q=tssr/2.q=\frac{-\partial_{t}s}{s-r/2}.

We apply the maximum principle to logq=log|ts|log(sr/2)\log q=\log|\partial_{t}s|-\log(s-r/2). In particular, at a fixed time t[0,T]t\in[0,T], consider a point Y𝕊nY\in\mathbb{S}^{n} at which qq attains its maximum. By changing coordinates, we may assume that this point Y=(0,,0,1)Y=(0,\dots,0,-1) is the south pole. Then, as in Tso [13], consider the coordinates y=(y1,,yn)y=(y^{1},\dots,y^{n}) for ss restricted to the hyperplane {(y1,,yn,1)}\{(y^{1},\dots,y^{n},-1)\}. At y=0y=0, we have for i=1,,ni=1,\dots,n

(logq)i=0stist=sisr/2(\log q)_{i}=0\quad\Longleftrightarrow\quad\frac{s_{ti}}{s_{t}}=\frac{s_{i}}{s-r/2} (3.1)

The condition for (logq)|𝕊n(\log q)|_{\mathbb{S}^{n}} to have a maximum at the south pole is

(logq)ij+(logq)n+1δij0(\log q)_{ij}+(\log q)_{n+1}\delta_{ij}\leq 0 (3.2)

as a symmetric matrix. Here we use subscripts to denote ordinary differentiation fi=yiff_{i}=\partial_{y^{i}}f and ft=tff_{t}=\partial_{t}f.

To compute the second term in (3.2), use Euler’s identities for a function of homogeneity one

i=1n+1yisti=st,i=1n+1yisi=s\sum_{i=1}^{n+1}y^{i}s_{ti}=s_{t},\qquad\sum_{i=1}^{n+1}y^{i}s_{i}=s

at the point Y=(0,,0,1)Y=(0,\dots,0,-1) to conclude stn+1=sts_{tn+1}=-s_{t}, sn+1=ss_{n+1}=-s, and

(logq)n+1=r/2sr/2.(\log q)_{n+1}=\frac{r/2}{s-r/2}.

For the first term in (3.2), compute

(logq)ij\displaystyle(\log q)_{ij} =\displaystyle= stijststistjst2sijsr/2+sisj(sr/2)2\displaystyle\frac{s_{tij}}{s_{t}}-\frac{s_{ti}s_{tj}}{s_{t}^{2}}-\frac{s_{ij}}{s-r/2}+\frac{s_{i}s_{j}}{(s-r/2)^{2}}
=\displaystyle= stijstsijsr/2\displaystyle\frac{s_{tij}}{s_{t}}-\frac{s_{ij}}{s-r/2}

at y=0y=0 by (3.1). Thus (3.2) becomes at y=0y=0

r/2sr/2δij+stijstsijsr/20.\frac{r/2}{s-r/2}\delta_{ij}+\frac{s_{tij}}{s_{t}}-\frac{s_{ij}}{s-r/2}\leq 0. (3.3)

Now, we compute using the flow equation (1.1)

(logq)t\displaystyle(\log q)_{t} =\displaystyle= tlog|ts|tlog(sr/2)\displaystyle\partial_{t}\log|\partial_{t}s|-\partial_{t}\log(s-r/2)
=\displaystyle= 1n+2tlogdet(sij)stsr/2\displaystyle-\frac{1}{n+2}\,\partial_{t}\log\det(s_{ij})-\frac{s_{t}}{s-r/2}
=\displaystyle= 1n+2sijstijstsr/2\displaystyle-\frac{1}{n+2}\,s^{ij}s_{tij}-\frac{s_{t}}{s-r/2}

for sijs^{ij} the inverse matrix of sijs_{ij}. Then (3.3) implies that

(logq)t\displaystyle(\log q)_{t} \displaystyle\leq r/2n+2stsr/2δijsij2nn+2stsr/2\displaystyle\frac{r/2}{n+2}\cdot\frac{s_{t}}{s-r/2}\delta_{ij}s^{ij}-\frac{2n}{n+2}\cdot\frac{s_{t}}{s-r/2}
=\displaystyle= r/2n+2qδijsij+2nn+2q,\displaystyle-\frac{r/2}{n+2}\,q\,\delta_{ij}s^{ij}+\frac{2n}{n+2}\,q,
qt\displaystyle q_{t} \displaystyle\leq r/2n+2q2δijsij+2nn+2q2.\displaystyle-\frac{r/2}{n+2}\,q^{2}\,\delta_{ij}s^{ij}+\frac{2n}{n+2}\,q^{2}.

Now if we let μi\mu_{i} be the eigenvalues of sijs^{ij}, or equivalently the reciprocals of the eigenvalues of sijs_{ij}, then we see

|st|=(detsij)1n+2=(i=1nμi)1n+2(1ni=1nμi)nn+2=(1nδijsij)nn+2|s_{t}|=(\det s_{ij})^{-\frac{1}{n+2}}=\left(\prod_{i=1}^{n}\mu_{i}\right)^{\frac{1}{n+2}}\leq\left(\frac{1}{n}\sum_{i=1}^{n}\mu_{i}\right)^{\frac{n}{n+2}}=\left(\frac{1}{n}\,\delta_{ij}s^{ij}\right)^{\frac{n}{n+2}}

by the arithmetic-geometric mean inequality. Therefore,

δijsijn|st|n+2n=nqn+2n(sr/2)n+2nnqn+2n(r/2)n+2n\delta_{ij}s^{ij}\geq n|s_{t}|^{\frac{n+2}{n}}=nq^{\frac{n+2}{n}}(s-r/2)^{\frac{n+2}{n}}\geq nq^{\frac{n+2}{n}}(r/2)^{\frac{n+2}{n}}

since srs\geq r. And so finally, at y=0y=0, and thus at any maximum point of q|𝕊nq|_{\mathbb{S}^{n}},

qtn(r/2)2n+2nn+2q3n+2n+2nn+2q2.q_{t}\leq-\frac{n(r/2)^{\frac{2n+2}{n}}}{n+2}\,q^{\frac{3n+2}{n}}+\frac{2n}{n+2}\,q^{2}. (3.4)

Now define Q(t)=maxY𝕊nq(Y,t)Q(t)=\max_{Y\in\mathbb{S}^{n}}q(Y,t). Then (3.4) implies that

QtQ2(cnr2n+2nQn+2ncn)Q_{t}\leq-Q^{2}\left(c_{n}r^{\frac{2n+2}{n}}Q^{\frac{n+2}{n}}-c_{n}^{\prime}\right)

for constants cn,cnc_{n},c_{n}^{\prime} depending only on nn. Therefore,

Qmax{cnr2n+2n+2,cnr1tn2n+2}Q\leq\max\left\{c_{n}r^{-\frac{2n+2}{n+2}},c_{n}^{\prime}r^{-1}t^{-\frac{n}{2n+2}}\right\} (3.5)

for cn,cnc_{n},c_{n}^{\prime} new constants depending only on nn. The result easily follows.

Remark 1

QQ may not be differentiable as a function of tt, but the above estimate (3.5) still holds—see e.g. Hamilton [9, Section 3].

\Box

4 Gutiérrez-Huang’s Hessian Estimate

Again, for the convenience of the reader, we reproduce our version of Gutiérrez-Huang’s Pogorelov-type estimate [8] for solutions to the Monge-Ampère equation.

First we define a bowl-shaped domain in spacetime and its parabolic boundary. A set Ωn×\Omega\subset\mathbb{R}^{n}\times\mathbb{R} is bowl-shaped if there are constants t0<Tt_{0}<T so that

Ω=t0tTΩt×{t},\Omega=\bigcup_{t_{0}\leq t\leq T}\Omega_{t}\times\{t\},

where each Ωt\Omega_{t} is convex and Ωt1Ωt2\Omega_{t_{1}}\subset\Omega_{t_{2}} whenever t1<t2t_{1}<t_{2}. The parabolic boundary of Ω\Omega is then Ω(ΩT×{T}).\partial\Omega\setminus(\Omega_{T}\times\{T\}).

Proposition 4.1

Let ss be a smooth solution to (1.1) which is convex in yy, and let Ω\Omega be a bowl-shaped domain in space-time n×\mathbb{R}^{n}\times\mathbb{R} so that s=0s=0 on the parabolic boundary of Ω\Omega. Let β\beta be any unit direction in space.

Then at the maximum point PP of the function

w=|s|ββ2se12(βs)2,w=|s|\,\,\partial^{2}_{\beta\beta}s\,\,e^{\frac{1}{2}(\partial_{\beta}s)^{2}},

ww is bounded by a constant depending on only s(P)s(P), s(P)\nabla s(P) and nn.

Proof Choose coordinates so that β=(1,0,,0)\beta=(1,0,\dots,0) and so that at a maximum point PP of ww, sijs_{ij} is diagonal (in order to bound all second derivatives sββs_{\beta\beta}, it suffices to focus only on the eigendirections of the Hessian of ss).

Since ww is positive in Ω\Omega and 0 on the parabolic boundary, there is a point PP outside the parabolic boundary of Ω\Omega at which ww assumes its maximum value. We work with logw\log w instead of ww. Then at PP,

(logw)i=0,(logw)t0,(logw)ij0.(\log w)_{i}=0,\qquad(\log w)_{t}\geq 0,\qquad(\log w)_{ij}\leq 0.

Here we use i,j,ti,j,t subscripts for partial derivatives in yiy^{i}, yjy^{j} and tt, and the last inequality is as a symmetric matrix. These equations become, at PP,

sis+s11is11+s1s1i=0,\displaystyle\displaystyle\frac{s_{i}}{s}+\frac{s_{11i}}{s_{11}}+s_{1}s_{1i}=0, (4.1)
sts+s11ts11+s1s1t0,\displaystyle\displaystyle\frac{s_{t}}{s}+\frac{s_{11t}}{s_{11}}+s_{1}s_{1t}\geq 0, (4.2)
sijssisjs2+s11ijs11s11is11js112+s1js1i+s1s1ij0.\displaystyle\displaystyle\frac{s_{ij}}{s}-\frac{s_{i}s_{j}}{s^{2}}+\frac{s_{11ij}}{s_{11}}-\frac{s_{11i}s_{11j}}{s_{11}^{2}}+s_{1j}s_{1i}+s_{1}s_{1ij}\leq 0. (4.3)

To use (4.2), we compute, for D=detsijD=\det s_{ij},

s1t\displaystyle s_{1t} =\displaystyle= (D1n+2)1=1n+2D1n+2sijsij1,\displaystyle\left(D^{-\frac{1}{n+2}}\right)_{1}=\frac{1}{n+2}\,D^{-\frac{1}{n+2}}s^{ij}s_{ij1},
s11t\displaystyle s_{11t} =\displaystyle= D1n+2[1(n+2)2(sijsij1)21n+2siksjlskl1sij1+1n+2sijsij11].\displaystyle D^{-\frac{1}{n+2}}\left[-\frac{1}{(n+2)^{2}}(s^{ij}s_{ij1})^{2}-\frac{1}{n+2}\,s^{ik}s^{jl}s_{kl1}s_{ij1}+\frac{1}{n+2}\,s^{ij}s_{ij11}\right].

Now plug into (4.2) and divide out by D1n+2D^{-\frac{1}{n+2}} to find

1s11[1(n+2)2(sijsij1)21n+2siksjlskl1sij1+1n+2sijsij11]\displaystyle\displaystyle\frac{1}{s_{11}}\left[-\frac{1}{(n+2)^{2}}(s^{ij}s_{ij1})^{2}-\frac{1}{n+2}\,s^{ik}s^{jl}s_{kl1}s_{ij1}+\frac{1}{n+2}\,s^{ij}s_{ij11}\right]
1s+s1(1n+2sijsij1)0\displaystyle\displaystyle{}-\frac{1}{s}+s_{1}(\frac{1}{n+2}\,s^{ij}s_{ij1})\geq 0 (4.4)

The last term of the first line of (4.4) leads us to contract (4.3) with the positive-definite matrix sijs^{ij} so that at PP:

0\displaystyle 0 \displaystyle\geq sij(sijssisjs2+s11ijs11s11is11js112+s1js1i+s1s1ij)\displaystyle s^{ij}\left(\frac{s_{ij}}{s}-\frac{s_{i}s_{j}}{s^{2}}+\frac{s_{11ij}}{s_{11}}-\frac{s_{11i}s_{11j}}{s_{11}^{2}}+s_{1j}s_{1i}+s_{1}s_{1ij}\right)
=\displaystyle= ns2sijsisjs2+sijs11ijs11sijsis1s1jssijsjs1s1is\displaystyle\frac{n}{s}-\frac{2s^{ij}s_{i}s_{j}}{s^{2}}+\frac{s^{ij}s_{11ij}}{s_{11}}-\frac{s^{ij}s_{i}s_{1}s_{1j}}{s}-\frac{s^{ij}s_{j}s_{1}s_{1i}}{s}
sijs12s1is1j+sijs1js1i+sijs1s1ij(by (4.1))\displaystyle{}-s^{ij}s_{1}^{2}s_{1i}s_{1j}+s^{ij}s_{1j}s_{1i}+s^{ij}s_{1}s_{1ij}\qquad\qquad\mbox{(by (\ref{grad-w-zero}))}
=\displaystyle= ns2sijsisjs2+sijs11ijs112s12ss12s11+s11+sijs1s1ij\displaystyle\frac{n}{s}-\frac{2s^{ij}s_{i}s_{j}}{s^{2}}+\frac{s^{ij}s_{11ij}}{s_{11}}-\frac{2s_{1}^{2}}{s}-s_{1}^{2}s_{11}+s_{11}+s^{ij}s_{1}s_{1ij}
(since sijs_{ij} is diagonal at PP)
\displaystyle\geq ns2sijsisjs22s12ss12s11+s11+sijs1s1ij+n+2s\displaystyle\frac{n}{s}-\frac{2s^{ij}s_{i}s_{j}}{s^{2}}-\frac{2s_{1}^{2}}{s}-s_{1}^{2}s_{11}+s_{11}+s^{ij}s_{1}s_{1ij}+\frac{n+2}{s}
s1sijsij1+(sijsij1)2(n+2)s11+siksjlskl1sij1s11\displaystyle{}-s_{1}s^{ij}s_{ij1}+\frac{(s^{ij}s_{ij1})^{2}}{(n+2)s_{11}}+\frac{s^{ik}s^{jl}s_{kl1}s_{ij1}}{s_{11}}
(by (4.4))
\displaystyle\geq 2n+2s2i=1nsi2s2sii2s12ss12s11+s11+i,j=1nsij12s11siisjj\displaystyle\frac{2n+2}{s}-2\sum_{i=1}^{n}\frac{s_{i}^{2}}{s^{2}s_{ii}}-\frac{2s_{1}^{2}}{s}-s_{1}^{2}s_{11}+s_{11}+\sum_{i,j=1}^{n}\frac{s_{ij1}^{2}}{s_{11}s_{ii}s_{jj}}

by collecting terms, completing the square, and since sijs_{ij} is diagonal at PP. Continue computing

0\displaystyle 0 \displaystyle\geq 2n+2s2i=1nsi2s2sii2s12ss12s11+s11+s1112s113+2i=2ns11i2s112sii\displaystyle\frac{2n+2}{s}-2\sum_{i=1}^{n}\frac{s_{i}^{2}}{s^{2}s_{ii}}-\frac{2s_{1}^{2}}{s}-s_{1}^{2}s_{11}+s_{11}+\frac{s_{111}^{2}}{s_{11}^{3}}+2\sum_{i=2}^{n}\frac{s_{11i}^{2}}{s_{11}^{2}s_{ii}}
=\displaystyle= 2n+2s2s12s2s112s12ss12s11+s11+s12s11s2+2s12s+s12s11\displaystyle\frac{2n+2}{s}-\frac{2s_{1}^{2}}{s^{2}s_{11}}-\frac{2s_{1}^{2}}{s}-s_{1}^{2}s_{11}+s_{11}+\frac{s_{1}^{2}}{s_{11}s^{2}}+\frac{2s_{1}^{2}}{s}+s_{1}^{2}s_{11}

by (4.1) and since sijs_{ij} is diagonal at PP. Finally, collect terms so that

0s11+2n+2s+1s11(s12s2)0\geq s_{11}+\frac{2n+2}{s}+\frac{1}{s_{11}}\left(-\frac{s_{1}^{2}}{s^{2}}\right)

and multiply each side of the inequality by s2s11es12s^{2}s_{11}e^{s_{1}^{2}} to find a quadratic inequality

w2+aw+b0w^{2}+aw+b\leq 0

for w=|s|s11e12s12w=|s|s_{11}e^{\frac{1}{2}s_{1}^{2}} at PP the point in Ω\Omega at which the maximum of ww is achieved. The coefficients aa and bb involve only nn, s(P)s(P) and s1(P)s_{1}(P), and so there is an upper bound of ww on Ω\Omega depending on only these quantities. \Box

This bounds sijs_{ij} away from infinity, which, together with Andrews’s speed estimate, shows that the ellipticity is locally uniformly controlled in the interior of appropriate bowl-shaped domains. In the next section, we use barriers essentially due to Calabi [3] to ensure that appropriate bowl-shaped domains exist, and so Gutiérrez-Huang’s estimate applies.

5 Barriers

We will use two soliton solutions to the affine normal flow as inner and outer barriers. First of all, the unit sphere is a shrinking soliton, and we use its affine images, ellipsoids, as inner barriers. Since the ellipsoids are compact, their support functions are finite and smooth on all n+1\mathbb{R}^{n+1}, and the usual maximum principle applies: If for an ellipsoid EE, sEsis_{E}\leq s_{i} on all n+1\mathbb{R}^{n+1} (which is equivalent to the inclusion of convex hulls E^i^\hat{E}\subset\widehat{\mathcal{L}_{i}} for i\mathcal{L}_{i} the hypersurface whose support function is sis_{i}), then the maximum principle for parabolic equations on 𝕊n\mathbb{S}^{n} shows that sE(t)si(t)s_{E}(t)\leq s_{i}(t) for all positive tt before the extinction time of sE(t)s_{E}(t).

The outer barrier we use is an expanding soliton due to Calabi [3]. Upon taking an affine transformation, its support function s𝒞s_{\mathcal{C}} has 𝒟(s𝒞)\mathcal{D}^{\circ}(s_{\mathcal{C}}) an open cone over a simplex, and has the value of a linear function there. (Outside its domain, recall the support function is ++\infty.) Moreover, under the affine normal flow, s𝒞(t)s_{\mathcal{C}}(t) satisfies Dirichlet conditions on the boundary, and is continuous and finite on the closure of its domain. These properties make Calabi’s example very useful as an outer barrier (as exploited by Cheng-Yau [5, 6] for the elliptic real Monge-Ampère equation).

Recall that siss_{i}\nearrow s, where sis_{i} are the support functions of strictly convex smooth compact hypersurfaces i\mathcal{L}_{i} which approach \mathcal{L}. On 𝒟(s)\mathcal{D}^{\circ}(s), as siss_{i}\nearrow s uniformly on compact subsets, and since the sis_{i} are convex, we automatically have uniform C0C^{0} and C1C^{1} estimates on compact subsets of 𝒟(s)\mathcal{D}^{\circ}(s). We define s(t)=limisi(t)s(t)=\lim_{i\to\infty}s_{i}(t) for positive tt also, and so we have locally uniform C0C^{0} and C1C^{1} estimates for positive tt as well.

To get similar uniform local ellipticity bounds for small positive tt, we need to check the hypotheses of Propositions 3.1 and 4.1 as well. For Proposition 3.1, we must ensure that si(Y)rs_{i}(Y)\geq r for all large ii, t[0,T]t\in[0,T], and Y𝕊nY\in\mathbb{S}^{n}. The affine normal flow of a sphere provides a lower barrier to show this. In particular, we have the solution corresponding to the affine normal flow of a sphere centered at the origin. For any r0>0r_{0}>0, let

u(Y,t)=r(t)|Y|,r(t)=(r02n+2n+22n+2n+2t)n+22n+2.u(Y,t)=r(t)|Y|,\qquad r(t)=\left(r_{0}^{\frac{2n+2}{n+2}}-\frac{2n+2}{n+2}\,t\right)^{\frac{n+2}{2n+2}}. (5.1)

Then uu satisfies the affine normal flow equation for a support function. Now the nondegeneracy assumption (1.3) shows that we can use the transformation law (1.5) with AA the identity matrix and b=Pb=-P to show s(Y)ϵs(Y)\geq\epsilon for all Y𝕊nY\in\mathbb{S}^{n}. Thus (5.1) and the maximum principle show that for r=ϵ/2r=\epsilon/2 there is a T>0T>0 so that for all t[0,T]t\in[0,T] and Y𝕊nY\in\mathbb{S}^{n}, and large ii, we have si(Y,t)rs_{i}(Y,t)\geq r. Thus we can apply Andrews’s estimate for all time in t[0,T]t\in[0,T].

Proposition 5.1

Let \mathcal{L} be a noncompact convex properly embedded hypersurface in n+1\mathbb{R}^{n+1} which contains no lines. Then the affine normal flow (t)\mathcal{L}(t) exists for all positive time t>0t>0.

Proof We will phrase this in terms of the support function. Since \mathcal{L} is noncompact, there is a ray R={v+tw:t0}R=\{v+tw:t\geq 0\} contained in the convex hull ^\hat{\mathcal{L}}. We may choose coordinates so that w=(0,1)n×w=(0,1)\in\mathbb{R}^{n}\times\mathbb{R}. Therefore, the support function

s(Y)=s(Y)sR(Y)={+foryn+1>0v,Yforyn+10s(Y)=s_{\mathcal{L}}(Y)\geq s_{R}(Y)=\left\{\begin{array}[]{c@{\quad\mbox{for}\quad}c}+\infty&y^{n+1}>0\\ \langle v,Y\rangle&y^{n+1}\leq 0\end{array}\right.

We will use this estimate, together with the nondegeneracy assumption (1.3) to provide a lower barrier. In particular, there is an ϵ>0\epsilon>0 so that s(Y)=+s(Y)=+\infty for yn+1>0y^{n+1}>0 and

s(Y)ϵ|Y|+v,Yforyn+10.s(Y)\geq\epsilon|Y|+\langle v,Y\rangle\quad\mbox{for}\quad y^{n+1}\leq 0.

The barrier we will use is, for y=(y1,,yn)y=(y^{1},\dots,y^{n}) and Y=(y,yn+1)Y=(y,y^{n+1}),

sEj(Y)=ϵ|y|2+(jyn+1)2+v,Y+jyn+1.s_{E_{j}}(Y)=\epsilon\sqrt{|y|^{2}+(jy^{n+1})^{2}}+\langle v,Y\rangle+jy^{n+1}.

This is the support function of an ellipsoid centered at P+(0,j)P+(0,j) with nn minor axes of length ϵ\epsilon and one major axis of length ϵj\epsilon j. Clearly for all j>1j>1, sEj(Y)s(Y)s_{E_{j}}(Y)\leq s(Y). As jj\to\infty, the ellipsoid is equivalent, under a volume-preserving affine map, to a sphere of radius ϵj1n+1\epsilon j^{\frac{1}{n+1}}, which also goes to infinity. Now (5.1) shows that the extinction time of the ellipsoid under the affine normal flow goes to infinity as jj\to\infty. Since the sEjs_{E_{j}} are all lower barriers to ss (which is equivalent to the ellipsoids EjE_{j} being inside the convex hull ^\hat{\mathcal{L}}), we have that the affine normal flow applied to ss must exist for all time. \Box

Now to find appropriate bowl-shaped domains to apply Proposition 4.1, we use an upper barrier due to Calabi. This barrier is first used in the real elliptic Monge-Ampère equation by Cheng-Yau [5, 6]. Calabi’s example is based on the fact that the hypersurface

𝒞(t)={(x1,,xn+1)n+1:xi>0,j=1n+1xj=k>0}\mathcal{C}(t)=\left\{(x^{1},\dots,x^{n+1})\in\mathbb{R}^{n+1}:x_{i}>0,\,\prod_{j=1}^{n+1}x^{j}=k>0\right\}

is an expanding soliton for the affine normal flow (which evolves by setting the parameter k=k(t)k=k(t) for an appropriate function). At time t=0t=0, we set the hypersurface

𝒞(0)={(x1,,xn+1)n+1:xi0,j=1n+1xj=0}\mathcal{C}(0)=\left\{(x^{1},\dots,x^{n+1})\in\mathbb{R}^{n+1}:x_{i}\geq 0,\,\prod_{j=1}^{n+1}x^{j}=0\right\}

the boundary of the first orthant in n+1\mathbb{R}^{n+1}. The support function of this example is given for cn=(n+1)12(2n+2)n+22c_{n}=(n+1)^{\frac{1}{2}}(\frac{2}{n+2})^{\frac{n+2}{2}} :

s𝒞(Y,t)={+ if any yi>0(n+1)(cntn+2ni=1n+1|yi|)1n+1 if all yi0s_{\mathcal{C}}(Y,t)=\left\{\begin{array}[]{cl}+\infty&\mbox{ if any }y^{i}>0\\ -(n+1)\left(c_{n}t^{\frac{n+2}{n}}\prod_{i=1}^{n+1}|y^{i}|\right)^{\frac{1}{n+1}}&\mbox{ if all }y^{i}\leq 0\end{array}\right. (5.2)

Note in particular that for time t=0t=0, s𝒞(Y,0)s_{\mathcal{C}}(Y,0) is 0 on the closed orthant on which all the yi0y^{i}\leq 0 and is ++\infty elsewhere. In order to find a more flexible class of barriers, we can apply (1.5) to transform s𝒞s_{\mathcal{C}} by a volume-preserving affine map Φ:xAx+b\Phi\!:x\mapsto Ax+b to be any linear function b,Y\langle b,Y\rangle on any linear image (A)1𝒞(A^{\top})^{-1}\mathcal{C}, and ++\infty elsewhere. In our standard affine coordinates Y=(y,1)Y=(y,-1), we find that the support function of 𝒞(0)\mathcal{C}(0) can be transformed to have its domain be a simplex (this is a projective image of the first orthant in n\mathbb{R}^{n}), and the value of sΦ𝒞(0)s_{\Phi\mathcal{C}}(0) is any affine function of yy on this domain. The graphs of these functions will give us the flexibility to create upper barriers for the support function which ensure that the function ss does move by a certain amount under the affine normal flow. This in turn gives a bowl-shaped domain in which to apply Gutiérrez-Huang’s interior estimates for the Hessian of ss.

Assume that the domain 𝒟(s)\mathcal{D}^{\circ}(s) is contained in the lower half-space of n+1\mathbb{R}^{n+1}. So since ss has homogeneity one, ss can be described by its behavior on the affine hyperplane ={(y,1):yn}\mathcal{H}=\{(y,-1):y\in\mathbb{R}^{n}\}. For the remainder of this section, we consider the domain 𝒟(s)\mathcal{D}^{\circ}(s) to be a subset of n\mathbb{R}^{n}, as identified with the affine plane \mathcal{H}.

Each x𝒟(s)x\in\mathcal{D}^{\circ}(s) has a convex neighborhood 𝒩\mathcal{N} on which siss_{i}\to s uniformly as an increasing sequence of convex functions, and so that the Lipschitz norms siC0,1(𝒩)\|s_{i}\|_{C^{0,1}}(\mathcal{N}) are bounded by a constant CC independent of ii. By adding linear functions (constant in tt) to the sis_{i}, we may assume si(x)=0s_{i}(x)=0 and si(x)=0\nabla s_{i}(x)=0. This normalization does not affect the Monge-Ampère equation (1.1) or the Hessian of sis_{i} (and so the C2C^{2} estimates we derive apply to the original sis_{i} as well). We can choose points p1,pn+1p_{1},\dots p_{n+1} so that 0si(y)C0\leq s_{i}(y)\leq C^{\prime} for CC^{\prime} a constant independent of ii and yy in the convex hull 𝒬\mathcal{Q} of the pjp_{j}. We may also assume that xx is in the interior of 𝒬\mathcal{Q}. Now consider the simplices 𝒮j\mathcal{S}_{j} to be the convex hull in n\mathbb{R}^{n} of the points

x,p1,,pj1,pj^,pj+1,,pn+1,x,p_{1},\dots,p_{j-1},\widehat{p_{j}},p_{j+1},\dots,p_{n+1},

where pjp_{j} is omitted from the list. Define PjP_{j} to be an affine function on each 𝒮j\mathcal{S}_{j} which is equal to CC^{\prime} on each of the pk𝒮jp_{k}\in\mathcal{S}_{j} and is equal to 0 at xx, and define PjP_{j} to be ++\infty outside 𝒮j\mathcal{S}_{j}. Then define P(y)=minjPj(y)P(y)=\min_{j}P_{j}(y). Then it is clear that PP is satisfies Pj(y)P(y)si(y)P_{j}(y)\geq P(y)\geq s_{i}(y) for all ii and for all yny\in\mathbb{R}^{n}.

We do not know the explicit solution to the Monge-Ampère equation (1.1) with initial value PP, but all we need to show to produce uniformly large bowl-shaped domains centered at xx for each of the sis_{i} is that P(x,t)<0P(x,t)<0 for positive tt. This can be verified as follows: By the discussion above, PjP_{j} is the image of Calabi’s example 𝒞(0)\mathcal{C}(0) under an affine transformation zAz+bz\mapsto Az+b of n+1\mathbb{R}^{n+1}. By the explicit solution (5.2) and the transformation law (1.5), we see that P(t,y)Pj(t,y)<0P(t,y)\leq P_{j}(t,y)<0 for small t>0t>0 and all yy near xx on the ray from xx to the barycenter of 𝒮j\mathcal{S}_{j}. Therefore, since P(t,y)P(t,y) is convex in yy and xx is in the convex hull of the barycenters of the 𝒮j\mathcal{S}_{j}, we have shown that P(t,x)<0P(t,x)<0 for all small positive tt.

By the maximum principle, each sub-level set of each sis_{i} contains a sub-level set of PP, which shows that x𝒟(s)x\in\mathcal{D}^{\circ}(s) has a uniformly large bowl-shaped domain around it for each sis_{i} independently of ii. So Gutiérrez-Huang’s Hessian estimates are uniform in every compact subset of 𝒟(s)×(0,T]\mathcal{D}^{\circ}(s)\times(0,T] for small TT.

By standard techniques, both Gutiérrez-Huang’s and Andrews’s estimates can be extended in time to be uniform in compact subsets of 𝒟(s)×(0,)\mathcal{D}^{\circ}(s)\times(0,\infty). These estimates uniformly control the spacelike C2C^{2} norm and the ellipticity of sis_{i}. Then the Monge-Ampère equation allows us to apply Krylov’s regularity theory to get local uniform C2+α,1+α/2C^{2+\alpha,1+\alpha/2} estimates, which can then be bootstrapped to show

Theorem 5.1

On 𝒟(s)×(0,)\mathcal{D}^{\circ}(s)\times(0,\infty), siss_{i}\to s in the ClocC^{\infty}_{\rm loc} topology.

Also note that in [10] we use the same inner and outer barriers to show

Proposition 5.2

Under the affine normal flow, ss satisfies a Dirichlet boundary condition on 𝒟(s)\partial\mathcal{D}(s).

This proposition holds regardless of the boundary regularity—ss can be infinite or finite and discontinuous on the boundary 𝒟(s)\partial\mathcal{D}(s) [12]. We also use the barriers to show

Proposition 5.3

For every t>0t>0, F=F(y,t)F=F(y,t) is properly embedded as a function of yy for (y,1)𝒟(s)(y,-1)\in\mathcal{D}^{\circ}(s). In other words, as (y,1)𝒟(s)(y,-1)\to\partial\mathcal{D}^{\circ}(s), at least one coordinate of

F(y)=(s1(y),,sn(y),sk(y)yks(y))F(y)=(s_{1}(y),\dots,s_{n}(y),s_{k}(y)y^{k}-s(y))

goes to ±\pm\infty.

6 The evolution of |C|2|C|^{2}

Here we recall an estimate of Andrews [1] on the evolution of |C|2=gilgjmgkpCijkClmp|C|^{2}=g^{il}g^{jm}g^{kp}C_{ijk}C_{lmp}. For a compact strictly convex initial hypersurface evolving under the affine normal flow,

(t1n+2Δ)|C|22n(n+2)|C|4.\left(\partial_{t}-\frac{1}{n+2}\Delta\right)|C|^{2}\leq-\frac{2}{n(n+2)}|C|^{4}.

Then the maximum principle shows that for all t(0,T)t\in(0,T) for TT the extinction time,

|C|2n(n+2)2t|C|^{2}\leq\frac{n(n+2)}{2t} (6.1)

independently of initial conditions.

Since Theorem 5.1 above shows that siss_{i}\to s in ClocC^{\infty}_{\rm loc} on 𝒟(s)×(0,)\mathcal{D}^{\circ}(s)\times(0,\infty), the pointwise bound (6.1) survives in the limit for any solution to the affine normal flow beginning at time t=0t=0. If the flow begins at time τ\tau instead, then of course we have

|C|2n(n+2)2(tτ),|C|^{2}\leq\frac{n(n+2)}{2(t-\tau)},

and for an ancient solution (τ\tau\to-\infty), we must have |C|2=0|C|^{2}=0. In the following section, we give a proof of the classical theorem of Berwald that says that Cijk=0C_{ijk}=0 implies the hypersurface is quadric. Thus any ancient solution to the affine normal flow must be a quadric hypersurface. Since a hyperboloid cannot form part of an ancient solution, we have

Theorem 6.1

Any ancient solution to the affine normal flow is a paraboloid or an ellipsoid.

7 Quadric Hypersurfaces

Now we prove a classical theorem of Berwald, that the cubic form Cijk=0C_{ijk}=0 implies that the hypersurface is a quadric. The first step is to show that the hypersurface is an affine sphere (i.e., that ξ=aF+V\xi=aF+V for a constant scalar aa and a constant vector VV).

Compute for Cijk=0C_{ijk}=0

sijk=1n+2(sij(lnD)k+sjk(lnD)i+ski(lnD)j)s_{ijk}=\frac{1}{n+2}\Big{(}s_{ij}(\ln D)_{k}+s_{jk}(\ln D)_{i}+s_{ki}(\ln D)_{j}\Big{)} (7.1)

and differentiate to find

(n+2)sijkl=(sijl(lnD)k+sij(lnD)kl+sjkl(lnD)i+sjk(lnD)il+skil(lnD)j+ski(lnD)jl)=1n+2(sij(lnD)l(lnD)k+sjl(lnD)i(lnD)k+sli(lnD)j(lnD)k)+sij(lnD)kl+1n+2(sjk(lnD)l(lnD)i+skl(lnD)j(lnD)i+slj(lnD)k(lnD)i)+sjk(lnD)il+1n+2(ski(lnD)l(lnD)j+sil(lnD)k(lnD)j+slk(lnD)i(lnD)j)+ski(lnD)jl\begin{split}&(n+2)s_{ijkl}\\ =\ &\Big{(}s_{ijl}(\ln D)_{k}+s_{ij}(\ln D)_{kl}+s_{jkl}(\ln D)_{i}+s_{jk}(\ln D)_{il}+s_{kil}(\ln D)_{j}+s_{ki}(\ln D)_{jl}\Big{)}\\ =\ &\frac{1}{n+2}(s_{ij}(\ln D)_{l}(\ln D)_{k}+s_{jl}(\ln D)_{i}(\ln D)_{k}+s_{li}(\ln D)_{j}(\ln D)_{k})+s_{ij}(\ln D)_{kl}\\ +\ &\frac{1}{n+2}(s_{jk}(\ln D)_{l}(\ln D)_{i}+s_{kl}(\ln D)_{j}(\ln D)_{i}+s_{lj}(\ln D)_{k}(\ln D)_{i})+s_{jk}(\ln D)_{il}\\ +\ &\frac{1}{n+2}(s_{ki}(\ln D)_{l}(\ln D)_{j}+s_{il}(\ln D)_{k}(\ln D)_{j}+s_{lk}(\ln D)_{i}(\ln D)_{j})+s_{ki}(\ln D)_{jl}\end{split} (7.2)
(n+2)silkj=1n+2(sil(lnD)j(lnD)k+slj(lnD)i(lnD)k+sji(lnD)l(lnD)k)+sil(lnD)kj+1n+2(slk(lnD)j(lnD)i+skj(lnD)l(lnD)i+sjl(lnD)k(lnD)i)+slk(lnD)ij+1n+2(ski(lnD)j(lnD)l+sij(lnD)k(lnD)l+sjk(lnD)i(lnD)l)+ski(lnD)lj\begin{split}&(n+2)s_{ilkj}\\ =\ &\frac{1}{n+2}(s_{il}(\ln D)_{j}(\ln D)_{k}+s_{lj}(\ln D)_{i}(\ln D)_{k}+s_{ji}(\ln D)_{l}(\ln D)_{k})+s_{il}(\ln D)_{kj}\\ +\ &\frac{1}{n+2}(s_{lk}(\ln D)_{j}(\ln D)_{i}+s_{kj}(\ln D)_{l}(\ln D)_{i}+s_{jl}(\ln D)_{k}(\ln D)_{i})+s_{lk}(\ln D)_{ij}\\ +\ &\frac{1}{n+2}(s_{ki}(\ln D)_{j}(\ln D)_{l}+s_{ij}(\ln D)_{k}(\ln D)_{l}+s_{jk}(\ln D)_{i}(\ln D)_{l})+s_{ki}(\ln D)_{lj}\end{split} (7.3)

Using sijkl=silkjs_{ijkl}=s_{ilkj}, we have

sij((lnD)kl1n+2(lnD)k(lnD)l)+sjk((lnD)li1n+2(lnD)l(lnD)i)=sil((lnD)kj1n+2(lnD)k(lnD)j)+slk((lnD)ij1n+2(lnD)i(lnD)l)\begin{split}&s_{ij}((\ln D)_{kl}-\frac{1}{n+2}(\ln D)_{k}(\ln D)_{l})+s_{jk}((\ln D)_{li}-\frac{1}{n+2}(\ln D)_{l}(\ln D)_{i})\\ =\ &s_{il}((\ln D)_{kj}-\frac{1}{n+2}(\ln D)_{k}(\ln D)_{j})+s_{lk}((\ln D)_{ij}-\frac{1}{n+2}(\ln D)_{i}(\ln D)_{l})\end{split} (7.4)

Multiplying sijs^{ij} to previous equation, we get

n((lnD)kl1n+2(lnD)k(lnD)l)+((lnD)lk1n+2(lnD)l(lnD)k)=((lnD)kl1n+2(lnD)k(lnD)l)+slksij((lnD)ij1n+2(lnD)i(lnD)l)\begin{split}&n((\ln D)_{kl}-\frac{1}{n+2}(\ln D)_{k}(\ln D)_{l})+((\ln D)_{lk}-\frac{1}{n+2}(\ln D)_{l}(\ln D)_{k})\\ &=((\ln D)_{kl}-\frac{1}{n+2}(\ln D)_{k}(\ln D)_{l})+s_{lk}s^{ij}((\ln D)_{ij}-\frac{1}{n+2}(\ln D)_{i}(\ln D)_{l})\end{split} (7.5)

So

n((lnD)kl1n+2(lnD)k(lnD)l)=slksij((lnD)ij1n+2(lnD)i(lnD)l)n((\ln D)_{kl}-\frac{1}{n+2}(\ln D)_{k}(\ln D)_{l})=s_{lk}s^{ij}((\ln D)_{ij}-\frac{1}{n+2}(\ln D)_{i}(\ln D)_{l})

Let SS be the matrix (sij)(s_{ij}) and TT be the matrix with Tij=(lnD)ij(lnD)i(lnD)jn+2T_{ij}=(\ln D)_{ij}-\frac{(\ln D)_{i}(\ln D)_{j}}{n+2}. So we have T=gijTijnST=\frac{g^{ij}T_{ij}}{n}S. Denote trT=gijTij\mbox{tr}\,T=g^{ij}T_{ij}.

From (n+2)ξ=D1n+2((lnD)1,,(lnD)n,(n+2)+(lnD)iyi)(n+2)\xi=D^{-\frac{1}{n+2}}((\ln D)_{1},\dots,(\ln D)_{n},(n+2)+(\ln D)_{i}y^{i}). So for ξi\xi^{i} the ithi^{\rm th} component of ξ\xi,

(n+2)j(ξi)=j(D1n+2(lnD)i)=1n+2D1n+2(lnD)j(lnD)i+D1n+2(lnD)ij=D1n+2Tij=D1n+2trTnsij\begin{split}&(n+2)\partial_{j}(\xi^{i})\\ =\ &\partial_{j}(D^{-\frac{1}{n+2}}(\ln D)_{i})\\ =\ &-\frac{1}{n+2}D^{-\frac{1}{n+2}}(\ln D)_{j}(\ln D)_{i}+D^{-\frac{1}{n+2}}(\ln D)_{ij}\\ =\ &D^{-\frac{1}{n+2}}T_{ij}=\frac{D^{-\frac{1}{n+2}}\,\mbox{tr}\,T}{n}s_{ij}\end{split} (7.6)

for 1in1\leq i\leq n. Similarly,

(n+2)j(ξn+1)=D1n+2((lnD)ij1n+2(lnD)i(lnD)j)yi=D1n+2Tijyi(n+2)\partial_{j}(\xi^{n+1})=D^{-\frac{1}{n+2}}((\ln D)_{ij}-\frac{1}{n+2}(\ln D)_{i}(\ln D)_{j})y^{i}=D^{-\frac{1}{n+2}}T_{ij}y^{i}
=D1n+2trTnsijyi=\frac{D^{-\frac{1}{n+2}}\,\mbox{tr}\,T}{n}s_{ij}y^{i}

Therefore ξ,i=D1n+2trTnFi\xi_{,i}=\frac{D^{-\frac{1}{n+2}}\,\mbox{tr}\,T}{n}F_{i}

Recall that Fi=(s1i,,sni,sliyl)F_{i}=(s_{1i},\dots,s_{ni},s_{li}y^{l}). We have ξ,i=D1n+2trTnFi\xi_{,i}=\frac{D^{-\frac{1}{n+2}}\,\mbox{tr}\,T}{n}F_{i}. Affine curvature is defined by ξ,i=AikF,k\xi_{,i}=-A^{k}_{i}F_{,k}. So Aik=D1n+2trTnδik=aδik-A^{k}_{i}=\frac{D^{-\frac{1}{n+2}}\,\mbox{tr}\,T}{n}\delta^{k}_{i}=a\delta^{k}_{i} where a=D1n+2trTna=\frac{D^{-\frac{1}{n+2}}\,\mbox{tr}\,T}{n}

Now the affine structure equations, applied to the second ordinary derivative ξij\xi_{ij}, shows

ξij\displaystyle\xi_{ij} =\displaystyle= (aFi)j\displaystyle(aF_{i})_{j}
=\displaystyle= ajFi+aFij\displaystyle a_{j}F_{i}+aF_{ij}
=\displaystyle= ajFi+a(gijξ+(Γijk+Cijk)Fk\displaystyle a_{j}F_{i}+a(g_{ij}\xi+(\Gamma_{ij}^{k}+C_{ij}^{k})F_{k}
=\displaystyle= (ajδik+aΓijk)Fk+agijξ.\displaystyle(a_{j}\delta^{k}_{i}+a\Gamma_{ij}^{k})F_{k}+ag_{ij}\xi.

So ajδika_{j}\delta^{k}_{i} must be symmetric in i,ji,j, and in particular, aiδkk=akδika_{i}\delta^{k}_{k}=a_{k}\delta^{k}_{i}. Since n2n\geq 2, we have ai=0a_{i}=0 for all ii. So aa is constant and ξk=aFk\xi_{k}=aF_{k} implies that ξ=aF+V\xi=aF+V, where VV is a constant vector.

So far, we have shown

Proposition 7.1

Let n2n\geq 2. If Cijk=0C_{ijk}=0 then ξ=aF+V\xi=aF+V for VV a constant vector and aa a constant scalar.

The rest of the proof of the following theorem follows Nomizu-Sasaki [11].

Theorem 7.1

Assume n2n\geq 2. If the cubic form Cijk=0C_{ijk}=0, then the hypersurface given by the image of FF is a quadric hypersurface. In other words, there is a second-degree polynomial map 𝒫:n+1\mathcal{P}\!:\mathbb{R}^{n+1}\to\mathbb{R} so that \mathcal{L} is an open subset of {𝒫=0}\{\mathcal{P}=0\}.

Proof Let \mathcal{L} denote our hypersurface with is (locally) the image of the embedding FF. For each x=F(y)x=F(y)\in\mathcal{L}, since {F1,,Fn,ξ}\{F_{1},\dots,F_{n},\xi\} is a basis of n+1\mathbb{R}^{n+1}, we can write each point PnP\in\mathbb{R}^{n} uniquely as

P=F(y)+UPi(y)Fi(y)+μP(y)ξ(y).P=F(y)+U^{i}_{P}(y)F_{i}(y)+\mu_{P}(y)\xi(y). (7.7)

Then the Lie quadric of \mathcal{L} at x=F(y)x=F(y) is defined as the locus

y={Pn+1:gijUiUjaμ22μ=0},\mathcal{F}_{y}=\{P\in\mathbb{R}^{n+1}:g_{ij}U^{i}U^{j}-a\mu^{2}-2\mu=0\},

where aa is the constant determined in Proposition 7.1 above and gij=gij(y)g_{ij}=g_{ij}(y). For each yy, y\mathcal{F}_{y} is clearly a quadric hypersurface in n+1\mathbb{R}^{n+1}.

Now we will show that for each xx\in\mathcal{L}, that x\mathcal{L}\subset\mathcal{F}_{x}. By dimension considerations, this show that \mathcal{L} is an open subset of the quadric x\mathcal{F}_{x}, and we are done. Now consider y0y_{0} for F(y0)=PF(y_{0})=P\in\mathcal{L}, and consider UiU^{i} and μ\mu defined in (7.7) above as functions of yy with y0y_{0} fixed. Now differentiate (7.7) to find for k=1,,nk=1,\dots,n and Uki=kUiU^{i}_{k}=\partial_{k}U^{i},

0=kP=UkiFi+UiFik+μkξ+μξk.0=\partial_{k}P=U^{i}_{k}F_{i}+U^{i}F_{ik}+\mu_{k}\xi+\mu\xi_{k}.

By Proposition 7.1, ξk=aFk\xi_{k}=aF_{k}, and also Fik=(Γikj+Cikj)Fj+gikξF_{ik}=(\Gamma^{j}_{ik}+C^{j}_{ik})F_{j}+g_{ik}\xi for CikjC^{j}_{ik} the cubic form and Γikj\Gamma^{j}_{ik} the Levi-Civita connection with respect to the affine metric gijg_{ij}. Since we assume the cubic form is zero, we have Fik=ΓikjFj+gikξF_{ik}=\Gamma^{j}_{ik}F_{j}+g_{ik}\xi. Thus

0=UkiFi+Ui(ΓikjFj+gikξ)+μkξ+μaFk,0=U^{i}_{k}F_{i}+U^{i}(\Gamma^{j}_{ik}F_{j}+g_{ik}\xi)+\mu_{k}\xi+\mu aF_{k},

and by splitting into the components on the basis {F1,,Fn,ξ}\{F_{1},\dots,F_{n},\xi\}, we find

Ukj\displaystyle U^{j}_{k} =\displaystyle= UiΓikj(1+aμ)δkjfor j,k=1,,n,\displaystyle-U^{i}\Gamma^{j}_{ik}-(1+a\mu)\delta^{j}_{k}\quad\mbox{for }j,k=1,\dots,n, (7.8)
μk\displaystyle\mu_{k} =\displaystyle= Uigikfor k=1,,n.\displaystyle-U^{i}g_{ik}\quad\mbox{for }k=1,\dots,n. (7.9)

Now define Φ:\Phi\!:\mathcal{L}\to\mathbb{R} by

Φ(y)=gijUiUjaμ22μ=D1n+2sijUiUjaμ22μ.\Phi(y)=g_{ij}U^{i}U^{j}-a\mu^{2}-2\mu=D^{\frac{1}{n+2}}s_{ij}U^{i}U^{j}-a\mu^{2}-2\mu.

Note Φ(y0)=0\Phi(y_{0})=0 since by definition Ui(y0)=μ(y0)=0U^{i}(y_{0})=\mu(y_{0})=0. So if we show Φk=0\Phi_{k}=0, then Φ(y)=0\Phi(y)=0 for all yy. By the definitions of Φ,Ui,μ\Phi,U^{i},\mu, then we will have shown y0yy_{0}\in\mathcal{F}_{y} and so y\mathcal{L}\subset\mathcal{F}_{y}.

So in order to complete the proof of the theorem, we must check Φk=0\Phi_{k}=0. So compute, using (7.8) and (7.9) above,

Φk\displaystyle\Phi_{k} =\displaystyle= 1n+2D1n+2(lnD)ksijUiUj+D1n+2sijkUiUj+2D1n+2sijUkiUj\displaystyle\frac{1}{n+2}\,D^{\frac{1}{n+2}}(\ln D)_{k}s_{ij}U^{i}U^{j}+D^{\frac{1}{n+2}}s_{ijk}U^{i}U^{j}+2D^{\frac{1}{n+2}}s_{ij}U^{i}_{k}U^{j}
2aμμk2μk\displaystyle{}-2a\mu\mu_{k}-2\mu_{k}
=\displaystyle= 1n+2D1n+2(lnD)ksijUiUj+D1n+2sijkUiUj+2(aμ+1)Uigik\displaystyle\frac{1}{n+2}\,D^{\frac{1}{n+2}}(\ln D)_{k}s_{ij}U^{i}U^{j}+D^{\frac{1}{n+2}}s_{ijk}U^{i}U^{j}+2(a\mu+1)U^{i}g_{ik}
+2D1n+2sijUj[UlΓlki(1+aμ)δki]\displaystyle{}+2D^{\frac{1}{n+2}}s_{ij}U^{j}[-U^{l}\Gamma^{i}_{lk}-(1+a\mu)\delta^{i}_{k}]
=\displaystyle= 1n+2D1n+2(lnD)ksijUiUj+D1n+2sijkUiUj2D1n+2sijUjUlΓlki\displaystyle\frac{1}{n+2}\,D^{\frac{1}{n+2}}(\ln D)_{k}s_{ij}U^{i}U^{j}+D^{\frac{1}{n+2}}s_{ijk}U^{i}U^{j}-2D^{\frac{1}{n+2}}s_{ij}U^{j}U^{l}\Gamma^{i}_{lk}
=\displaystyle= 1n+2D1n+2(lnD)ksijUiUj+D1n+2sijkUiUjD1n+2sij[1n+2(lnD)kδli\displaystyle\frac{1}{n+2}\,D^{\frac{1}{n+2}}(\ln D)_{k}s_{ij}U^{i}U^{j}+D^{\frac{1}{n+2}}s_{ijk}U^{i}U^{j}-D^{\frac{1}{n+2}}s_{ij}\left[\frac{1}{n+2}(\ln D)_{k}\delta^{i}_{l}\right.
+1n+2(lnD)lδki+simslkm1n+2sim(lnD)mslk]\displaystyle\left.{}+\frac{1}{n+2}(\ln D)_{l}\delta^{i}_{k}+s^{im}s_{lkm}-\frac{1}{n+2}s^{im}(\ln D)_{m}s_{lk}\right]
=\displaystyle= 0.\displaystyle 0.

This completes the proof of Theorem 7.1. \Box

References

  • [1] B. Andrews. Contraction of convex hypersurfaces by their affine normal. Journal of Differential Geometry, 43(2):207–230, 1996.
  • [2] B. Andrews. Motion of hypersurfaces by Gauss curvature. Pacific J. Math., 195(1):1–34, 2000.
  • [3] E. Calabi. Complete affine hyperspheres I. Instituto Nazionale di Alta Matematica Symposia Mathematica, 10:19–38, 1972.
  • [4] S. Y. Cheng and S. T. Yau. On the regularity of the solution of the nn-dimensional Minkowski problem. Comm. Pure Appl. Math., 29(5):495–516, 1976.
  • [5] S.-Y. Cheng and S.-T. Yau. On the regularity of the Monge-Ampère equation det((2u/xixj))=F(x,u)\det((\partial^{2}u/\partial x^{i}\partial x^{j}))={F}(x,u). Communications on Pure and Applied Mathematics, 30:41–68, 1977.
  • [6] S.-Y. Cheng and S.-T. Yau. On the real Monge-Ampère equation and affine flat structures. In Proceedings of the 1980 Beijing Symposium on Differential Geometry and Differential Equations, Vol. 1,2,3 (Beijing, 1980), pages 339–370. Science Press, 1982.
  • [7] B. Chow. Deforming convex hypersurfaces by the nnth root of the Gaussian curvature. J. Differential Geom., 22(1):117–138, 1985.
  • [8] C. E. Gutiérrez and Q. Huang. A generalization of a theorem by Calabi to the parabolic Monge-Ampère equation. Indiana Univ. Math. J., 47(4):1459–1480, 1998.
  • [9] R. S. Hamilton. Four-manifolds with positive curvature operator. J. Differential Geom., 24(2):153–179, 1986.
  • [10] J. Loftin and M.-P. Tsui. Ancient solutions of the affine normal flow. to appear, Journal of Differential Geometry, math.DG/0602484.
  • [11] K. Nomizu and T. Sasaki. Affine Differential Geometry: Geometry of Affine Immersions. Cambridge University Press, 1994.
  • [12] R. T. Rockafellar. Convex analysis. Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 1997. Reprint of the 1970 original, Princeton Paperbacks.
  • [13] K. Tso. Deforming a hypersurface by its Gauss-Kronecker curvature. Comm. Pure Appl. Math., 38(6):867–882, 1985.