30 000 ft overview of ancient solutions

  • Speaker: Mat Langford ( University of Newcastle, University of Tennessee Knoxville )
  • Date: 2 September 2020

Abstract

Introduction to ancient solutions of heat type equations and flows of hypersurfaces, such as the curve shortening flow and the mean curvature flow.

References

  1. Angenent, Sigurd; Daskalopoulos, Panagiota; Sesum, Natasa. Unique asymptotics of ancient convex mean curvature flow solutions. J. Differential Geom. 111 (2019), no. 3, 381--455. doi:10.4310/jdg/1552442605
  2. Sigurd Angenent. Formal asymptotic expansions for symmetric ancient ovals in mean curvature flow. Networks & Heterogeneous Media, 2013, 8 (1) : 1-8. doi:10.3934/nhm.2013.8.1
  3. Uniqueness of two-convex closed ancient solutions to the mean curvature flow Pages 353-436 from Volume 192 (2020), Issue 2 by Sigurd Angenent, Panagiota Daskalopoulos, Natasa Sesum doi:10.4007/annals.2020.192.2.2
  4. Uniqueness of convex ancient solutions to mean curvature flow in higher dimensions. Brendle, K. Choi arXiv:1804.00018
  5. Angenent, S., Brendle, S., Daskalopoulos, P. and Šešum, N. (2020), Unique Asymptotics of Compact Ancient Solutions to Three‐Dimensional Ricci Flow. Comm. Pure Appl. Math. doi:10.1002/cpa.21955
  6. Unique asymptotics of ancient compact non-collapsed solutions to the 3-dimensional Ricci flow, Sigurd Angenent, Panagiota Daskalopoulos, Natasa Sesum. arXiv:1906.11967
  7. Uniqueness of compact ancient solutions to three-dimensional Ricci flow, Simon Brendle, Panagiota Daskalopulos, Natasa Sesum. arXiv:2002.12240
  8. Brendle, S., Choi, K. Uniqueness of convex ancient solutions to mean curvature flow in \(\mathbb{R}^3\). Invent. math. 217, 35–76 (2019). doi:10.1007/s00222-019-00859-4

Classification of convex, ancient solutions to the Curve Shortening Flow

  • Speaker: Mat Langford ( University of Newcastle, University of Tennessee Knoxville )
  • Date: 9 September 2020

Abstract

Convex, ancient solutions to the Curve Shortening Flow are classified as circles, angenent ovals and grim reapers.

References

  1. Bourni, T., Langford, M. & Tinaglia, G. Convex ancient solutions to curve shortening flow. Calc. Var. 59, 133 (2020). 10.1007/s00526-020-01784-8
  2. Wang, Xu-Jia. Convex solutions to the mean curvature flow. Ann. of Math. (2) 173 (2011), no. 3, 1185--1239. 10.4007/annals.2011.173.3.1
  3. Daskalopoulos, Panagiota; Hamilton, Richard; Sesum, Natasa. Classification of compact ancient solutions to the curve shortening flow. J. Differential Geom. 84 (2010), no. 3, 455--464. MR2669361 10.4310/jdg/1279114297

An Introduction to ancient Ricci flows

Abstract

Aspects of Ricci solitons with a view to classifying 3d shrinkning, gradient Ricci solitons.

References

Classification of 3d Ricci Solitons

Abstract

3d Ricci Solitons are classified as the sphere, cylinder, Gaussian shrinker and various quotients thereof.

References

Harmonic functions with polynomial growth

  • Speaker: Ben Andrews ( Australian National University )
  • Date: 30 September 2020

Abstract

Using comparison geometry, the space of harmonic maps with polynomial growth of given degree are shown to be harmonic polynomials, hence finite dimensional.

References

  1. Colding, Tobias H.; Minicozzi, II, William P. Harmonic functions with polynomial growth. J. Differential Geom. 46 (1997), no. 1, 1--77. doi:10.4310/jdg/1214459897
  2. Colding, Tobias H.; Minicozzi, William P., II. Harmonic functions on manifolds. Ann. of Math. (2) 146 (1997), no. 3, 725--747. doi:10.2307/2952459
  3. Li, Peter. Harmonic sections of polynomial growth. Math. Res. Lett. 4 (1997), no. 1, 35--44. doi:10.4310/MRL.1997.v4.n1.a4

Harmonic functions with polynomial growth 2

  • Speaker: Ben Andrews ( Australian National University )
  • Date: 7 October 2020

Abstract

Ben Andrews continues to describe harmonic functions of polynomial growth on complete, non-compact manifolds of nonnegative Ricci curvature. He describes Yau's gradient estimate, the Harnack inequality and the mean value inequality and begins to apply these to the problem at hand.

References

  1. Colding, Tobias H.; Minicozzi, II, William P. Harmonic functions with polynomial growth. J. Differential Geom. 46 (1997), no. 1, 1--77. doi:10.4310/jdg/1214459897
  2. Colding, Tobias H.; Minicozzi, William P., II. Harmonic functions on manifolds. Ann. of Math. (2) 146 (1997), no. 3, 725--747. doi:10.2307/2952459
  3. Li, Peter. Harmonic sections of polynomial growth. Math. Res. Lett. 4 (1997), no. 1, 35--44. doi:10.4310/MRL.1997.v4.n1.a4

Transcript

What follows is a very rough transcript taken during the lecture. It has only been edited to ensure the TeX compiles. No effort has been expended as yet to ensure accuracy. There are surely many errors. Please send any comments, suggestions, typos etc. to Paul Bryan

Download transcript PDF

\((M^n, g)\) is a complete, non-compact Riemannian manifold with \(\Ric \geq 0\).

  1. Laplacian Comparison: \(\Delta \rho^2 \leq 2n\), \(\rho(x) = d(x, x_0)\)
  2. Volume comparision: \(\abs{B_R(x_0)} \leq \left(\frac{R}{r}\right)^n \abs{B_r(x_0)}\) for \(0 < r \leq R\)
  3. Poincaré inequality: \(\int_{B_r(x_0)} \abs{\nabla \varphi}^2 \geq \frac{C(n)}{r^2} \int_{B_r(x_0)} \varphi^2\) for \(\varphi|_{B_r(x_0)} = 0\).

If \(u > 0\), \(\Delta u = 0\) on \(B_r(x_0) \subset M\) then \[ (r^2 -\rho(x)^2) \abs{\nabla \ln u(x)} \leq c(n) r \] on \(B_r(x_0)\).

WLOG \(r = 1\). We use \(\Delta u = 0\), \(u > 0\) and let \(v = \ln u\).

\begin{equation} \label{eq:Deltav} \Delta v = \nabla_i \left(\frac{\nabla_i u}{u}\right) = \frac{\Delta u}{u} - \abs{\frac{\nabla u}{u}}^2 = - \abs{\nabla v}^2. \end{equation}

Let \(F = \abs{\nabla v}^2\). Then differentiating \eqref{eq:Deltav}, \[ 0 = \nabla_i (\Delta v + F) = \Delta \nabla_i v - R^p_i \nabla_p u + \nabla_i F \] Note at a maximum point \(\abs{\nabla F} = 2\abs{\tfrac{\nabla \varphi}{\varphi}} F\).

Therefore

\begin{align} \label{eq:DeltaF} 0 &= \Delta F - 2 \abs{\nabla^2 v}^2 - 2\operatorname{Ric}(\nabla v, \nabla v) + 2 \nabla v \cdot \nabla F \\ &\leq \Delta F - \frac{2}{n} (\Delta v)^2 + 2 \sqrt{F}\abs{\nabla F} \notag \\ &= \Delta F - \frac{2}{n} F^2 + 2\sqrt{F} \abs{\nabla F} \notag \end{align}

Write \(\varphi = 1 - \rho^2\) with \(\rho = d(\cdot, x_0)\). Let \(G = \varphi^2 F\). On \(\partial B_1(x_0)\), \(G = 0\).

At \(x\),

\begin{align*} 0 = \nabla G = \varphi^2 \nabla_i F = 2 \varphi \nabla_i \varphi F \\ = \varphi^2\left(\nabla_i F + \frac{\nabla \varphi}{\varphi^3} G\right) \end{align*}

At \(x\), using \eqref{eq:DeltaF}

\begin{align*} 0 &\geq \Delta G = \varphi^2(\Delta F + (2 \frac{\Delta \varphi}{\varphi^2} - 6 \frac{\abs{\nabla \varphi}^2}{\varphi^4})\varphi^2 F) \\ &\geq \varphi^2(\frac{2}{n} F^2 - 2\sqrt{F} 2 \frac{\abs{\nabla \varphi}}{\varphi} F + (2 \frac{\Delta \varphi}{\varphi} - 6 \frac{\abs{\nabla \varphi^2}}{\varphi^2} F)) \\ &= F(\frac{2}{n} G - 4 \sqrt{G} \abs{\varphi} + 2 \varphi\Delta\varphi - 6 \abs{\nabla \varphi^2}) \\ &\geq F(\frac{1}{n} G - C \abs{\nabla \varphi}^2 + 2\varphi\Delta\varphi) \end{align*}

using Cauchy-Schwarz in the last line.

From \(\varphi = 1 - \rho^2\), \(\nabla \varphi = -2\rho\abs{\nabla \rho} \Rightarrow \abs{\nabla \varphi} \leq 4\rho^2 \leq 4\) and \(\Delta \varphi = - \Delta \rho^2 \geq -2n\). Thus \[ G \leq C(n). \]

If \(u > 0\), \(\Delta u = 0\) on \(B_r(x_0) \subseteq M\) then \[ \sup_{B_{r_1}(x_0)} u \leq C(n) \inf B_{r_2}(x_0) u \]

WLOG \(r = 1\). By the gradient estimate, \(\abs{\nabla \ln u} \leq C(n)\), \[ \frac{u(y)}{u(x)} \leq e^{C(n)} d(x, y) \leq e^{C(n)}. \]

Assume \((M, g)\) is complete, non-compact with \(\operatorname{Ric} \geq 0\).

If \(v\) is a non-negative, subharmonic function \((\Delta v \geq 0)\) on \(B_r(x_0)\), then \[ v(x_0) \leq c(n) \frac{1}{\abs{B_r(x_0)}} \int_{B_r(x_0)} v. \]

[Proof in the case \(v = u^2\), \(\Delta u = 0\).] \[ \Delta v = 2 \abs{\nabla u}^2 \geq 0 \]

  • Case 1: If \(h\) is harmonic and \(h \geq 0\), the Harnack inequality gives
\begin{align*} h(x_0) &\leq \sup_{B_{r/2}(x_0)} h \leq C(n) \inf_{B_{r/2}} h(x_0) \\ &\leq C(n) \frac{1}{B_{r/2}(x_0)} \int_{B_{r/2}(x_0)} h \\ &\leq C(n) \frac{1}{B_{r}(x_0)} \int_{B_{r/2}(x_0)} h \frac{B_{r}(x_0)}{B_{r/2}(x_0)} \\ &\leq 2^n C(n) \frac{1}{B_{r(x_0)}} \int_{B_r(x_0)} h \end{align*}

using volume comparison.

  • Case 2: \(v = u^2\), \(\Delta u = 0\). Let \(h\) be the harmonic function on \(B_{r/2}\) with \(h|_{\partial B_r} = \abs{u}\). Note the Harnack gives \(h > 0\) on the interior.

Note \(\abs{u}\) is subharmonic hence \(\abs{u}(x_0) \leq h(x_0)\) so

\begin{align*} u^2(x_0) &\leq h^2(x_0) \leq C(n) \left(\frac{1}{B_{r/2}(x_0)} \int_{B_{r/2}(x_0)} h\right)^2 \\ &\leq C(n) \frac{1}{B_{r/2} (x_0)} \int_{B_{r/2} (x_0)} h^2 \\ \end{align*}

by Hölder's inequality.

Write \[ \int_{B_{r/2}} h^2 = \int_{B_r} (h - \abs{u} + \abs{u})^2 \leq 2 \int_{B_{r/2}} (h - \abs{u})^2 + 2 \int_{B_{r/2}} u^2 \] noting that \(h - \abs{u} = 0\) on the boundary. The using the Poincare inequality, \[ \int_{B_{r/2}} (h - \abs{u})^2 \leq C(n) \int_{B_{r/2}} \abs{\nabla h - \nabla \abs{u}}^2 \leq C \int_{B_{r/2}} \abs{\nabla h}^2 + \abs{\nabla u}^2 \] Since \(h\) is harmonic, it minimises the Dirichlet energy among maps with the same energy and hence \[ \int_{B_{r/2}} \abs{\nabla h}^2 \leq \int_{B_{r/2}} \abs{\nabla{\abs{u}}}^2 \leq \int_{B_{r/2}} \abs{\nabla u}^2 \] where we note \(u\) is smooth hence in \(W^{2,2}\).

Let \(\Phi \in C^{\infty}_c(B_1)\), \(\Phi \equiv 1\) on \(B_{r/2}\), \(\abs{\nabla \Phi} \leq C\).

\begin{align*} \int_{B_1} \Phi \abs{\nabla u}^2 &= - \int_{B_1} \Phi^2 u \Delta u - \int_{B_1} 2 \Phi u \nabla \Phi \cdot \nabla u \\ &= 2\left(\int_{B_1} \Phi^2 \abs{\nabla u}^2\right)^{1/2} \left(\int_{B_1} u^2 \abs{\nabla \Phi}^2\right)^{1/2} \end{align*}

Thus \[ \int_{B_{1/2}} \abs{\nabla u}^2 \leq \int_{B_1} \Phi^2 \abs{\nabla u}^2 \leq 4 \int_{B_1} u^2 \abs{\nabla \Phi}^2 \leq C \int_{B_1} u^2 \] giving \[ \int_{B_{r/2}} h^2 \leq C \int_{B_1} u^2 \] Using volume comparison, \[ u^2(x_0) \leq \frac{C}{\abs{B_{r/2} (x_0)}} \int_{B_1} u^2 \leq 2^n C \frac{1}{\abs{B_1}} \int_{B_1} u^2. \]

\[ \operatorname{dim} \mathcal{H}_p (M) \leq C(n) p^{n-1} \] where \[ \mathcal{H}_p(M) = \{u \in C^{\infty}(M) : \Delta u = 0, \abs{u(x)} \leq C(1 + d(x, x_0)^p\} \]

We aim to bound the dimension of any finite dimensional subspace, \(K\) of \(\mathcal{H}_p(M)\) which will give the result. First we have an estimate on how harmonic functions can be "packed" into a ball.

Let \(K\) be any finite dimensional subspace of \(\mathcal{H}_p(M)\) of \(\{u \in C^{\infty}(M) : \Delta u = 0\}\). Let \(\{u_i\}_{i=1}^k\) be any orthonormal basis of \(K\) with respect to \(L^2(B_r(x))\).

Then for any \(0 < \epsilon < 1/2\), \[ \int_{B_{(1-\epsilon)r}(x)} \sum_{i=1}^k u_i^2 \leq C(n) \epsilon^{-(n-1)}. \]

  1. The right hand side, \(C(n) \epsilon^{-(n-1)}\) is independent of \(K\). Thus while the space of all harmonic functions on a complete, non-compact manifold is infinite dimensional, any finite dimensional space must be concentrated towards the outer edge of the ball. We cannot fit too many harmonic functions into a small space.
  2. \(\int_{B_r(x)} u_i u_j = \delta_{ij} \Rightarrow \int_{B_r(x)} \sum_{i=1}^k u_i^2 = k\).

WLOG \(r = 1\).

If \(y \in B_{1-\epsilon}(x)\) by Gram-Schmidt, choose a rotation, \(\Theta\) of \(\mathbb{R}^k\) such that \[ \Theta \begin{pmatrix} u_1(y) \\ \vdots \\ u_k(y) \end{pmatrix} = \sqrt{\sum_{i=1}^k u_i^2(y)} \begin{pmatrix} 1 \\ 0 \\ \vdots \\ 0 \end{pmatrix} \]

Let \(u = \sum_i \Theta_i^i u_i\) (harmonic!). The MVI gives \[ u^2(y) \leq C \frac{1}{\abs{B_{1-\rho(y)}}} \int_{B_{1-\rho(y)}} u^2 \leq C \frac{1}{\abs{B_1(x)}} \int_{B_1(x)} u^2 \frac{\abs{B_1(x)}}{\abs{B_{1-\rho(y)}(y)}}. \] Volume comparison doesn't directly apply since we have different centres. But we get \[ u^2(y) \leq C \frac{\abs{B_{1+\rho}(y)}}{\abs{B_{1-\rho}(y)}} \frac{1}{\abs{B_1(x)}} \int_{B_1(x)} u^2 \leq C\left(\frac{1+\rho}{1-\rho}\right)^n \frac{1}{\abs{B_1(x)}} \int_{B_1(x)} u^2. \] Easy estimate: on \(B_{1-\epsilon}\), \(\tfrac{1+\rho}{1-\rho} \leq \tfrac{2}{\epsilon}\) giving \[ \sum_{i=1}^k u_i^2 (y) = u^2(y) \leq \frac{C \epsilon^{-n}}{\abs{B_1(x)}} \] and hence \[ \int_{B_{1-\epsilon}(x)} \sum_{i=1}^k u_i^2 \leq C \epsilon^{-n} \frac{\abs{B_{1-\epsilon}}}{\abs{B_1}}. \]

More work is required to bump up the estimate to \(\epsilon^{-(n-1)}\).

Harmonic functions with polynomial growth 3

  • Speaker: Ben Andrews ( Australian National University )
  • Date: 14 October 2020

Abstract

Ben Andrews continues to describe harmonic functions of polynomial growth on complete, non-compact manifolds of nonnegative Ricci curvature. He describes lower and upper bounds leading to the main theorem and briefly notes the situation for minimal submanifolds of Euclidean space.

References

  1. Colding, Tobias H.; Minicozzi, II, William P. Harmonic functions with polynomial growth. J. Differential Geom. 46 (1997), no. 1, 1--77. doi:10.4310/jdg/1214459897
  2. Colding, Tobias H.; Minicozzi, William P., II. Harmonic functions on manifolds. Ann. of Math. (2) 146 (1997), no. 3, 725--747. doi:10.2307/2952459
  3. Li, Peter. Harmonic sections of polynomial growth. Math. Res. Lett. 4 (1997), no. 1, 35--44. doi:10.4310/MRL.1997.v4.n1.a4

Transcript

What follows is a very rough transcript taken during the lecture. It has only been edited to ensure the TeX compiles. No effort has been expended as yet to ensure accuracy. There are surely many errors. Please send any comments, suggestions, typos etc. to Paul Bryan

Download transcript PDF

We continue proving the lemma from last week.

\label{upperbound} Let \(K\) be any finite dimensional subspace of \(\mathcal{H}(M)\) of \(\{u \in C^{\infty}(M) : \Delta u = 0\}\). Let \(\{u_i\}_{i=1}^k\) be any orthonormal basis of \(K\) with respect to \(L^2(B_r(x))\).

Then for any \(0 < \epsilon < 1/2\), \[ \int_{B_{(1-\epsilon)r}(x)} \sum_{i=1}^k u_i^2 \leq C(n) \epsilon^{-(n-1)}. \]

Recall

WLOG, \(r = 1\). If \(y \in B_{(1-\epsilon)} (x)\), then \[ \sum_{i=1}^k u_i^2(y) \leq c(n) \left(\frac{1+\rho(y)}{1-\rho(y)}\right)^n \avgint_{B_1} u^2. \] Recall we rotated in \(k\) so \(u^2(y) = \sum_{i=1}^k u_i^2(y)\) and we had the normalisation \(\int_{B_1} u^2 = 1\). Thus \[ \int_{B_{1-\epsilon}} \sum_{i=1}^k u_i^2 \leq \frac{C(n)}{\abs{B_{1-\epsilon}(x)}} \int_{B_{(1-\epsilon)} (x)} \left(\frac{1+\rho(y)}{1-\rho(y)}\right)^n. \]

Now we use the Laplace comparison theorem to bump the estimate up to the desired power \(\epsilon^{-(1-n)}\). We integrate by parts on the annulus,

\begin{align*} \int_{B_{1-\epsilon} \backslash B_{1/2}} (1-\rho)^{-n} &= \int_{B_{1-\epsilon} \backslash B_{1/2}} (1-\rho)^{-n} \abs{\nabla \rho}^2 \\ &= \frac{1}{n-1} \int_{B_{1-\epsilon} \backslash B_{1/2}} \nabla\left((1-\rho)^{-(n-1)} - \epsilon^{-(1-n)}\right) \cdot \nabla \rho \\ &= -\frac{1}{n-1} \int_{\partial B_{1/2}} \left((1-\rho)^{-(n-1)} - \epsilon^{-(1-n)}\right) \\ &\quad - \frac{1}{n-1} \int_{B_{1-\epsilon} \backslash B_{1/2}} \left((1-\rho)^{-(n-1)} - \epsilon^{-(1-n)} \Delta \rho \right) \\ &\leq C\epsilon^{-(n-1)} \frac{n-1}{\rho} \leq C\epsilon^{-(n-1)} \frac{n-1}{2} \end{align*}

where we use the Bishop-Gromov again to control the size of the region of integration.

Note that lemma \ref{upperbound} did not use the polynomial growth hypothesis. Now we make use of the hypothesis for the next lemma. The assumption means that such harmonic functions cannot grow too much "all the time" - a sort of converse to the previous lemma.

\label{lowerbound} Let \(K\) be any finite dimensional subspace of \(\mathcal{H}_p(M)\) of \(\{u \in C^{\infty}(M) : \Delta u = 0, \abs{u(y)} \leq C(1 + d(y, x_0))^p\}\).

Then for any \(x \in M\), \(\epsilon \in (0, 1/2]\), \(r_0 > 0\), \(\delta > 0\), there exists \(r > r_0\) such that if \(\{u_i\}_{i=1}^k\) is an orthonormal basis of \(K\) with respect to \(L^2(B_r(x))\), we have \[ \sum_{i=1}^k \int_{B_{(1-\epsilon)r}(x)} \abs{u_i}^2 \geq k(1-\epsilon)^{2p + n + \delta} = (1-\epsilon)^{2p+n+\delta} \int_{B_r(x)} \sum_{i=1}^k u_i^2. \]

Fix \(x \in M\), \(\epsilon \in (0, 1/2]\), \(r_0 > 0\), \(\delta > 0\). The \(\delta\) is here just to give a little extra room to obtain a strict inequality.

Define \(r_{\alpha} = r_0(1-\epsilon)^{-\alpha}\) for \(\alpha \in \NN\).

To obtain a contradiction, suppose the claimed inequality does not hold for \(r = r_{\alpha}\). Let \(G_{\alpha}\) be the inner-product on \(K\) coming from \(L^2(B_{r_{\alpha}}(x)\). If \(\{u_i\}_{i=1}^k\) is an orthonormal basis for \(K\) with respect to \(L^2(B_{r_{\alpha}}(x)\) so \((G_{\alpha})_{ij} = \delta_{ij}\). Then \[ (G_{\alpha-1})_{ij} = \int_{B_{(1-\epsilon)r_{\alpha}(x)}} u_i u_j \] hence \[ \Tr (G_{\alpha}^{-1} \circ G_{\alpha-1}) = \int_{B_{(1-\epsilon)r_{\alpha}}} \sum_{i=1}^k \abs{u_i}^2 < (1-\epsilon)^{2p+n+\delta} \] by the previous lemma, \ref{upperbound}. By the arithmetic-geometric inequality, \[ \det (G_{\alpha}^{-1} \circ G_{\alpha-1}) \leq (\tfrac{1}{k} \Tr (G_{\alpha}^{-1} \circ G_{\alpha-1}))^k < (1-\epsilon)^{(2\rho+n+\delta)k}. \] Thus

\begin{equation} \label{det} \det (G_{\alpha}^{-1} \circ G_0) = \det (G_{\alpha}^{-1} \circ G_{\alpha-1}) \det (G_{\alpha-1}^{-1} \circ G_{\alpha-2}) \det (G_{1}^{-1} \circ G_{0}) < (1-\epsilon)^{(2\rho + n+\delta)k\alpha}. \end{equation}

Fix an o/n basis for \(G_0\), \(\{u_i\}_{i=1}^k\) so \(\int_{B_{r_0}(x)} u_i u_j = \delta_{ij}\). Since \(u_i \in K \subseteq \mathcal{H}_p\), \[ \abs{u_i(y)} \leq C (1 + d(x,y))^p. \]

Again using the arithmetic-geometric inequality and Bishop-Gromov,

\begin{align*} k \det (G_0^{-1} \circ G_{\alpha})^{1/k} &\leq \Tr (G_0^{-1} \circ G_{\alpha}) = \sum_{i=1}^k \int_{B_{r_0} (x)} \abs{u_i}^2 \\ &\leq C r_{\alpha}^{2p} k \abs{B_{r_{\alpha}}} \\ &\leq C k\abs{B_{r_0}} (1-\epsilon)^{-n\alpha} r_0^{2p}(1-\epsilon)^{1-2p} \end{align*}

But by \eqref{det} the left hand side satisfies \[ k \det (G_0^{-1} \circ G_{\alpha})^{1/k} > k(1-\epsilon)^{-(2p+n+\delta)\alpha} \] giving \[ k(1-\epsilon)^{-(2p+n+\delta)\alpha} < C k\abs{B_{r_0}} (1-\epsilon)^{-n\alpha} r_0^{2p}(1-\epsilon)^{1-2p}. \] Simplifying gives for any \(\alpha\), \[ (1-\epsilon)^{-\alpha\delta} \leq C \] with \(C\) independent of \(\alpha\). Sending \(\alpha \to \infty\) gives a contradiction.

We see in the last inequality where the \(\delta > 0\) is required and that the proof does not work with \(\delta = 0\).

Now we may prove the main theorem by playing the two lemmas off against each other.

\[ \operatorname{dim} \mathcal{H}_p (M) \leq C(n) p^{n-1} \] where \[ \mathcal{H}_p(M) = \{u \in C^{\infty}(M) : \Delta u = 0, \abs{u(x)} \leq C(1 + d(x, x_0)^p\} \]

Let \(K \subseteq \mathcal{H}_p\) be any finite dimensional subspace, \(\epsilon \in (0, 1/2]\), \(r_0 = 1\) and any \(\delta > 0\) as in Lemma \ref{lowerbound} such that \[ \int_{B_{r(1-\epsilon)}} \sum_{i=1}^k \abs{u_i}^2 \geq k(1-\epsilon)^{2p + n + \delta}. \] On the other hand by Lemma \ref{upperbound} \[ \int_{B_{r(1-\epsilon)}} \sum_{i=1}^k \abs{u_i}^2 \leq C(n) \epsilon^{-(n-1)}. \] Combining these gives, \[ k \leq C(n) \epsilon^{-(n-1)}(1-\epsilon)^{-(2p+n+\delta)}. \] Now we choose \(\epsilon = \tfrac{1}{2p}\) giving, \[ k \leq C(n) 2^{n-1}p^{n-1}(1-\tfrac{1}{2p})^{-2p} s^{n\epsilon\delta} \leq \tilde{C}(n) p^{n-1} \] where we bounded \((1-\tfrac{1}{2p})^{-2p} \leq e\).

Colding-Minicozzi also consider the situation \(M^n \subset \RR^N\) a minimal submanifold with Euclidean volume growth: \(V(B_r\cap M) \leq C_0 r^n\). Then the dimensions of the space of harmonic functions of polynomial growth of order \(p\) is bounded by \(C(n, C_0)p^{n-1}\).

The idea is to use the Michael-Simon inequality of harmonic functions for minimal submanifolds to show that if \(\Delta u = 0\), we have the mean value inequality \[ u(x) \leq \frac{C}{r^n} \int_{B_r \cap M} u^2 \]

By the monotonicity formula, \[ \frac{d}{dr} \frac{\abs{B_r \cap M}}{r^n} \geq 0. \] Combined with Euclidean volume growth, we get \[ \abs{B_r \cap M} \sim r^n. \]

Then the mean value inequality is now a real mean value inequality and \[ \abs{\frac{B_R \cap M}{B_r \cap M}} \leq C\left(\frac{R}{r}\right)^n \] and the proof of the main theorem goes through as before.

In the situation of the remark, \(M\) is contained in a subspace of \(\RR^N\) of dimension depending only on \(n, C_0\).

Since \(M\) is minimal, the coordinate functions are harmonic and of linear growth rate. Thus each \(x^i \in \mathcal{H}_1(M)\). Then apply dimension bound.

Ancient solutions of the heat equation of polynomial growth

  • Speaker: Glen Wheeler ( University of Wollongong )
  • Date: 21 October 2020

Abstract

Glen Wheeler discusses non-negative ancient solutions to the heat equation on Euclidean space. He works through the representation theorem of Lin and Zhang for such solutions, which expresses the solution pointwise as integrals of Borel measures in a specific form. This is used later in part 2 to bound the dimension of the space of ancient solutions with a specific (polynomial) growth rate and obtain a beautiful decomposition into a polynomial in time with coefficients given by functions polyharmonic in space.

References

  1. Lin, F. and Zhang, Q.S. (2019), On Ancient Solutions of the Heat Equation. Comm. Pure Appl. Math., 72: 2006-2028. doi:10.1002/cpa.21820

Ancient solutions of the heat equation of polynomial growth 02

  • Speaker: Glen Wheeler ( University of Wollongong )
  • Date: 28 October 2020

Abstract

Glen Wheeler continues the ten-step approach to non-negative ancient solutions to the heat equation on Euclidean space. He works through the representation theorem of Lin and Zhang for such solutions, which expresses the solution pointwise as integrals of Borel measures in a specific form. This is used later in part 2 to bound the dimension of the space of ancient solutions with a specific (polynomial) growth rate and obtain a beautiful decomposition into a polynomial in time with coefficients given by functions polyharmonic in space.

References

  1. Lin, F. and Zhang, Q.S. (2019), On Ancient Solutions of the Heat Equation. Comm. Pure Appl. Math., 72: 2006-2028. doi:10.1002/cpa.21820

Transcript

What follows is a very rough transcript taken during the lecture. It has only been edited to ensure the TeX compiles. No effort has been expended as yet to ensure accuracy. There are surely many errors. Please send any comments, suggestions, typos etc. to Paul Bryan

Download transcript PDF
  1. Monotonicity (via Li-Yau differential Harnack) (proof sketch)
  2. Bernstein theorem for completely monotone functions (won't prove)
  3. Apply Fubini
  4. \textbf{Apply inverse Laplace transform}
  5. Integral estimate
  6. PDE for \(h(\cdot, t)\)
  7. Harnack inequality for measures
  8. Radon-Nikodym
  9. PDE for \(\tfrac{d\nu_x}{d\nu_0}\)
  10. Apply representation formula (Caffarelli-Littman) (won't prove)

So \[ \int_0^{\infty} e^{-ts} h(x, s) ds = \frac{1}{t} u(x, -t). \] Bernstein gives extension of \(t\) into right half complex plane and heat equation estimates give Laplace transform is invertible, \[ h(x, t) = \frac{1}{2\pi i} \int_{-1-i\infty}^{-1+it} e^{st} \frac{u(x, -s)}{s} ds. \]

Recall that \(\Delta f^x(t) + \partial_t f^x(t) = 0\) in \(\RR^n \times [0, \infty)\). Testing against a smooth cut-off function \(\phi \in C^{\infty}_c(\RR^n \to \RR)\),

\begin{align*} \partial_t \int_{\RR^n} f^x(t) \phi(x) dx &= - \int_{\RR^n} \phi(x) \Delta f^x dx \\ &= -\int_{\RR^n} \Delta \phi (x) f^x (t) dx \\ &= -\int_{\RR^n} \int_0^{\infty} s e^{-ts} \phi(x) d\nu(x, s). \end{align*}

Thus \[ \int_{\RR^n} \int_0^{\infty} se^{-ts} \phi(x) d\nu(x, s) = \int_{\RR^n}\int_0^{\infty} e^{-ts} \Delta \phi(x) d\nu(x, s) \] which is a Laplace transform again: \[ \int_0^{\infty} e^{-ts} \int_{\RR^n} (\Delta\phi(x) - s\phi(x)) dx d\nu(x, s) = 0 \] for all \(t > 0\) and \(\phi \in C^{\infty}_c(\RR^n \to \RR)\).

Define the signed measure \(\eta_{\phi}\) on \([0, \infty)\) by \[ d\eta_{\phi} = \int_{\RR^n} (\Delta \phi(x) - s \phi(x)) d\nu(x, s). \] We've shown, \[ \operatorname{LT}(\eta_{\phi}) = 0 \Rightarrow \eta_{\phi} = 0. \] That is, \[ 0 = \eta_{\phi}([0, t]) = \int_{\RR^n} \left[\Delta \phi(x) \underbrace{\int_0^t d\nu(x, s)}_{h(x, t)} - \phi(x) \int_0^t sd\nu(x, s) \right]dx. \]

Side note: \[ h(x, u) = \int_0^u d\nu(x, s) \] and so \(h = dx/d\nu(x, s)\) and we can integrate \[ \int_0^t s d\nu(x, s) \] by parts. Thus

\begin{align*} \int_{\RR^n} \Delta \phi(x) \int_0^t d\nu(x, s) &= \int_{\RR^n}\phi(x) \int_0^t s d\nu(x, s) \\ &= \int_{\RR^n} \phi(x) \left(th(x, t) - \int_0^t h(x, s)ds\right)dx \end{align*}

Because this is true for all \(\phi \in C^{\infty}_c(\RR^n \to \RR)\), \(h\) satisfies the integro-differential equation, \[ \Delta h(x, t) = t h(x, t) - \int_0^t h(x, s) ds \] hence \(h\) is smooth as the solution of \[ \left(\Delta - t\right) h(x, t) = - \int_0^t h(x, s) ds. \]

Fix \(x\) and consider the measure \(\nu_x\): \[ \nu_x([a, b]) = h(x, b) - h(x, a). \] Aim is to turn the equation for \(h\) into estimates for \(\Delta \nu\).

We have (writing \(I = [a, b]\)),

\begin{align*} \Delta \nu_x (I) &= \Delta h(x, b) - \Delta h(x, a) \\ &= b h(x, b) - \int_0^b h(x, s) ds \\ &\quad - ah(x, a) + \int_0^a h(x, s) ds \\ &= bh(x, b) - ah(x, a) - \int_a^b h(x, s) ds \\ &= b(\underbrace{h(x, b) - h(x, a)}_{\nu_x(I)}) - \int_a^b \underbrace{h(x, s) - h(x, a)}_{\geq 0} ds \end{align*}

using side calculation: \((b-a) h(x, a) = \int_a^b h(x, a) ds\) and recalling \(h\) is non-decreasing.

Thus, \[ \Delta \nu_x (I) \leq b \nu_x (I) \] or equivalently, \[ \frac{\Delta \nu_x(I)}{\nu_x(I)} \leq b. \]

Similarly, from below,

\begin{align*} \Delta \nu_x (I) &= b h(x, b) - ah(x, a) - \int_a^b h(x, s) ds \\ &= a(\underbrace{h(x, b) - h(x, a))}_{\nu_x(I)} - \int_a^b \underbrace{h(x, s) - h(x, b)}_{\leq 0} ds \\ &\geq a \nu_x(I). \end{align*}

Thus \[ \frac{\Delta \nu_x(I)}{\nu_x(I)} \geq a. \]

Let \[ c(x) = \frac{\Delta \nu_x (I)}{\nu_x (I)} \] satisfies, \[ \abs{c(x)} \leq b \]

The Harnack inequality for \(L = \Delta - C\) with \(I = [0, b]\) gives, \[ \nu_x(I) \leq c(\abs{x}, b) \nu_0(I). \] That is \(\nu_x\) is absolutely continuous with respect to \(\nu_0\).

Radon-Nikodym then gives, \[ \frac{d\nu_x}{d\nu_0}(s) \leq c(\abs{x}, b), \quad s \leq b. \]

From the integral estimate we obtained, \[ \int_0^{\infty} e^{-ts} \int_{\RR^n} (\Delta \phi(x) - s\phi(x)) d\nu_x (s) = 0. \] Now we are able to make a change of variables: \[ 0 = \int_0^{\infty} e^{-ts} \left(\int_{\RR^n} (\Delta \phi(x) - s\phi(x)) \frac{d\nu_x}{d\nu_0} (s) dx\right) d\nu_0(s) \] This is the Laplace transform of the signed measure \(\int_{\RR^n} (\Delta \phi(x) - s\phi(x)) \frac{d\nu_x}{d\nu_0} (s) dx\) so for \(\nu_0\)-a.e. \(s\) we have, \[ \int_{\RR^n} (\Delta \phi(x) - s\phi(x)) \frac{d\nu_x}{d\nu_0} (s) dx = 0 \] for every \(\phi \in C^{\infty}_c(\RR^n \to \RR)\). So, for \(\nu_0\)-a.e. \(s\), \[ (\Delta - s) \frac{d\nu_x}{d\nu_0} = 0 \] and \(\frac{d\nu_x}{d\nu_0}\) is smooth.

See also Kaprelovic `67 and Koranyi `79.

From \[ (\Delta - s) \frac{d\nu_x}{d\nu_0} = 0 \] we have the representation formula, \[ \frac{d\nu_x}{d\nu_0} (s) = \int_{\mathbb{S}^{n-1}} e^{x\cdot \xi\sqrt{s}} d\mu_s(\xi) \] for \(\mu_s\) a Borel measure on \(\mathbb{S}^{n-1}\).

This completes the proof because, \[ u(x, -t) = \int_0^{\infty} e^{-st} d\nu_x(s) \] implies \[ u(x, t) = \int_0^{\infty} \int_{\mathbb{S}^{n-1}} e^{st + x \cdot \xi \sqrt{s}} d\mu_s(\xi) d\nu_0(s). \] \qed

Replace \(\RR^n\) with a Riemannian manifold \((M^n, g)\).

For \((M, g)\) complete with \(\Ric \geq 0\) Li-Yau still applies to give \[ u(x, t) = \int_0^{\infty} e^{ts} h(s, x) dx \] where \(h\) solves \[ (\Delta - s) h(s, x) = 0. \]

Classifying solutions of this equation is the hard part.

Key concept: Martin boundary. (Martin '41, Murata ('86, '93, '02, '07)

The general theory gives, \[ h(x) = \int_{\Sigma_0} P_s(x, w) d\mu(w) \] for \(\mu\) a unique, non-negative Borel measure on \(\Sigma\) with \(\mu(\Sigma - \Sigma_0) = 0\).

Here, \[ \Sigma_0 = \{[w] : P(\cdot, w) \text{ is minimal}\} \] where the term minimal means that if \(0 \leq v(x) \leq P(x, w)\) then \(v(x) = P(x, w)\). Also,

\[ P_s(x, y) = \begin{cases} \frac{\Gamma_s(x, y)}{\Gamma_s(x_0, y)}, & y \ne x_0 \\ 0, & y = x_0, x\ne y \\ 1, & x = y = x_0. \end{cases} \]

\(\Gamma_s\) is the minimal Green's function for \(\Delta - s\).

[Lin-Zhang]

Let \((M, g)\) be a complete Riemannian manifold with \(\Ric \geq 0\). Suppose \(u\) is a non-negative, ancient solution of \(\partial_t u = \Delta_g u\). Then

  • \(u(x, -t)\) is completely monotone in \(t\),
  • \(\exists\) family of non-negative Borel measures \(\mu = \mu(\cdot, s)\) on the Martin boundary, \(\Sigma_s\) for \(\Delta - s\), and a Borel measure \(\rho\) on \([0, \infty)\) such that

\[ u(x, t) = \int_0^{\infty} \int_{\Sigma_s} e^{ts} P_s(x, w) d\mu(w, s) d\rho(s). \]

The condition \(\Ric\geq 0\) is not strictly necessary. It may be replaced by something like volume doubling and Poincare inequality.

  • What are conditions on \((M, g)\) such that \[ \Sigma_{s_1} \simeq \Sigma_{s_2} \] in some sense such as under the motion of a Killing or conformal Killing field.
  • Are there conditions determining the structure of \(\Sigma_s\) (compare Schoen for bounded sectional curvature).
  • What is \(P_s\)?

Ancient solutions of the heat equation of polynomial growth 03

  • Speaker: Glen Wheeler ( University of Wollongong )
  • Date: 4 November 2020

Abstract

Glen Wheeler provides a surprise proof that the space of ancient solutions to the heat equation with polynomial growth are finite dimensional. The proof was devised by Ben Andrews and arises by apply Li-Yau estimates to iterated time derivatives of the solution.

References

  1. Lin, F. and Zhang, Q.S. (2019), On Ancient Solutions of the Heat Equation. Comm. Pure Appl. Math., 72: 2006-2028. doi:10.1002/cpa.21820

Ancient solutions of the heat equation of semi-linear equations

  • Speaker: Kyeongsu Choi ( Korea Institute for Advanced Study )
  • Date: 11 November 2020

Abstract

Kyeongsu Choi gives an overview of classifying ancient solutions of semi-linear equations with some applications to geometric flows.

References

A family of 3d steady gradient solitons that are flying wings

  • Speaker: Yi Yai ( UC Berkeley )
  • Date: 25 November 2020

Abstract

Li Yai describes the construction of flying wing, steady gradient 3d Ricci solitions confirming a conjecture of Richard Hamilton and providing new examples of solitions not covered in classification results.

References

Transcript

What follows is a very rough transcript taken during the lecture. It has only been edited to ensure the TeX compiles. No effort has been expended as yet to ensure accuracy. There are surely many errors. Please send any comments, suggestions, typos etc. to Paul Bryan

Download transcript PDF

A steady gradient soliton is a metric \(g\) such that \[ \Ric = \frac{1}{2}\Lie_V g = \nabla^2 f. \] where \(V = \grad f\).

Given such a metric, let \[ g(t) = \varphi_t^{\ast} g \] where \(\varphi_t\) is the flow of \(-\grad f\) and \(t \in (-\infty, \infty)\). Then \[ \partial_t g = \varphi_t^{\ast} \Lie_{-\grad f} g = -2\varphi_t^{\ast} \Ric(g) = -2 \Ric(g(t)) \] so \(g\) is an eternal RF (Ricci Flow).

Hamilton's Cigar Solition \(\Sigma^2\): \[ g = dr^2 + \varphi^2(r) d\theta^2 \] with \(\theta \in [0, 2\pi)\) and where \(\lim_{r\to\infty}\varphi(r) < \infty\).

\[ (\Sigma, p_i) \to \RR \times \mathbb{S}^1 \] \(\Sigma\) collapsed.

Bryan solition: \(n \geq 3\). \[ g = dr^2 + \varphi^2(r) g_{\mathbb{S}^{n-1}} \] with \[ \text{warping asymptotics: } \varphi(r) \simeq r^{1/2} \] \[ \text{Scalar curvature asymptotics: } R \simeq r^{-1} \]

\[ (\Sigma, R(p_i) g, p_i) \to \RR \times \mathbb{S}^{n-1} \] non-collapsed.

Known 3d SGS (non-flat): \(\RR \times \operatorname{Cigar}\) (gradients) and Bryan solition \(n=3\).

  1. \(\Rm \not > 0\): By the maximum principle, \(\RR \times \operatorname{Cigar}\).
  2. (Brendle) non-collapsed implies Bryant soliton
  3. (Deng-Zhu) \(\frac{C^{-1}}{r} \leq R \leq \frac{C}{r}\) implies Bryant soliton.
  4. (Catino-Monticelli-Mastrolia) If \(\lim_{r\to\infty} \int_{B(p, r)} \frac{1}{r} R = 0\) implies quotients of \(\RR \times \operatorname{Cigar}\).

Hamilton's conjecture: there exists a 3d SGS that is a flying wing.

A flying wing is a SGS \((M, g)\) with \(\Rm \geq 0\) such that the blow-down is \(C_{\infty} (g) = \operatorname{Cone}([0, \theta])\), \(\theta \in (0, \pi)\).

The previous examples are not flying wings: \(\RR \times \operatorname{Cigar}\) has \(\theta = \pi\) while the Bryant solition has \(\theta = 0\). Flying wings also do not fit into the classification results above.

  • Constructed by X.J Wang in \(\RR^n\), \(n\geq 3\)
  • Complete existence results by Hoffman-Ilmanen-Martin-White and Bourni-Langford-Tingalia
  • Uniquness by HIMW in \(n=3\) and by BLT in \(n \geq 3\).
                MCF                                       RF                                        
 \(n=2\)        Grim Reaper                               Cigar Soliton                             
 \(n=3\)      
              
              
   Collapsed: Flying Wings               
       New examples (BLT).               
    Noncollapsed:   Bowl soliton         
 Collapsed: Flying wing                    
    (Thm 1, \(n=3\) and Thm 2)             
      Noncollaped: Bryant soliton          
 \(n \geq 4\) 
              
 Collapsed: Flying wings. non-collapsed: 
                                         
 Collapsed: Flying wings?                  
 Non colapsed Thm 1 \(n \geq 4\)           

[Thm 1] \(\forall \alpha \in (0, 1)\) \(\exists\) a \(\ZZ_2 \times O(n-1)\)-symmetric SGS \((M, g, f, p)\) with \(\Rm > 0\) such that \[ \lambda_1 = \alpha\lambda_2 = \cdots = \alpha \lambda_n \] where \(\lambda_i\) are eigenvalues of \(\Ric\) at \(p\).

\(n=3\)

Let \(\{X_{iu}, u \in [0, 1]\}_{i=1}^{\infty}\) be a sequence of smooth families of metrics on \(\mathbb{S}^2\) such that

  1. \(\ZZ_2 \times O(2)\)-symmetric
  2. \(K(X_{iu}) > 1\)
  3. \(X_{i0} = c_i g_{\mathbb{S}^2}\), \(c_i > 0\),
  4. \(\operatorname{diam}(X_{i1} \to \pi\) as \(i \to\infty\),
  5. \(\sup_{u\in[0, 1]} \operatorname{vol} (X_{iu}) \to 0\) as \(i \to \infty\).

Recall Dernelle's result:

There exists a unique EGS (Expanding Gradient Soliton) \((M_{in}, g_{in}, p_{in})\) with \(\Rm > 0\) such that \(C_{\infty}(g_{in}) = C(X_{in})\) and

  • \(R(p_{in}) = 1\)
  • \(X_{in} \to X_{in} \overset{\text{Dernelle}}{\to} (M_{in}, g_{in}, p_{in})\)

A sequence of EGS with \(\ZZ_2 \times O(n-1)\)-symmetry and \(\Rm > 0\) with \(R(p_i) = 1\) and \(AVR(g_i) \to 0\) as \(i \to \infty\) then \[ (M_i, g_i, p_i) \overset{Cheeger-Gromov}{\to} (M, g, p) \quad \text{ SGS \(R(p) = 1\)}. \]

Using Dernelle's result and the lemma \[ (M_{i0}, g_{i0}, p_{i0}) \to \text{Bryant soliton (by rotational symmetry)}, \quad \frac{\lambda_1}{\lambda_3} = 1. \] and \[ (M_{i1}, g_{i1}, p_{i1}) \to \RR \times \operatorname{Cigar}, \quad \frac{\lambda_1}{\lambda_3} = 0. \]

For any \(\alpha_0 \in (0, 1)\) there exists \(u_i \in (0, 1)\) such that \(\tfrac{\lambda_1}{\lambda_3} (g_{iu_i}) = 0\). Applying the lemma \[ (M_{in_i}, g_{in_i}, p_{in_i}) \to (M, g, p) \text{ a SGS} \quad \frac{\lambda_1}{\lambda_3} (g) = \alpha_0 \text{ at } p \]

The next theorem excludes the blow down is a ray so the construction of Thm 1 gives flying wings.

[Thm 2] Let \((M, g, p)\) be a \(\ZZ_2 \times O(2)\)-symmetric 3d SGS, and \(C_{\infty}(g) = \text{a ray} = \operatorname{Cone} (\{0\})\). Then it is a Bryant soliton.

In order to obtain a contradiction, suppose it is not a Bryant soliton.

The profile curve \(\Gamma\) is fixed by the \(O(2)\) action and the section \(\Sigma\) is fixed by the \(\ZZ_2\) action. Take a geodesic in \(\Sigma\) with \(\gamma(0) = p\). Away from the edges \[ g = g_0 + \varphi^2 d\theta, \theta \in [0, \pi) \] with \(g_0\) totally geodesic.

Define

  • \(h_1(s) = d(\gamma(s), \Gamma)\)
  • \(h_2(s) = \varphi(\gamma(s))\)

Computing gives \[ \frac{h_2'(2)}{h_2(s)} \leq C R(\gamma(s)) \]

\[ R(\gamma(s)) \leq C h_1^{-2} (s) \]

\[ \frac{h_2(s)}{h_1(s)} \to 0, s \to \infty \]

\[ h_1(s) h_2(s) \geq C s, \quad s \to \infty \]

Using the lemmas, \[ h_2(s) \leq C s^{\epsilon} \forall \epsilon < 1/2 \] and \[ h_2(s) \leq C \to \lim_{s\to\infty} R(\Gamma(s)) > 0 \]

Let \((M, g)\) be a \(\ZZ_2 \times O(2)\)-symmetric SGS. If \[ C_{\infty}(g) = \operatorname{Cone}([0, \alpha]), \alpha \in [0,\pi] \] Then \[ \lim_{s\to\infty} R(\Gamma(s)) = R(p) \sin \tfrac{\alpha}{2} \]

We have

\begin{align*} 0 \simeq \ip{\grad f}{\sigma'(\gamma)}|_{-D}^D &= \int_{-D}^D \partial_r \ip{\grad f}{\sigma'(\gamma)} dr \\ &= \int_{-D}^D \Ric(\sigma'(\gamma), \sigma'(\gamma)) dr \\ &= 2R^{1/2}(\Gamma(s)) \end{align*}

Thus \[ \lim_{s\to\infty} R(\Gamma(s)) = 0 \] contradicting \(\Rm > 0\) for the case \(\alpha = 0\).