Skip to main content
Log in

Ricci Solitons, Conical Singularities, and Nonuniqueness

  • Published:
Geometric and Functional Analysis Aims and scope Submit manuscript

Abstract

In dimension \(n=3\), there is a complete theory of weak solutions of Ricci flow—the singular Ricci flows introduced by Kleiner and Lott (Acta Math 219(1):65–134, 2017, in: Chen, Lu, Lu, Zhang (eds) Geometric analysis. Progress in mathematics, vol 333. Birkhäuser, Cham, 2018)—which Bamler and Kleiner (Uniqueness and stability of Ricci flow through singularities, arXiv:1709.04122v1, 2017) proved are unique across singularities. In this paper, we show that uniqueness should not be expected to hold for Ricci flow weak solutions in dimensions \(n\ge 5\). Specifically, for any integers \(p_1,p_2\ge 2\) with \(p_1+p_2\le 8\), and any \(K\in {\mathbb N}\), we construct a complete shrinking soliton metric \(g_K\) on \(\mathcal S^{p_1}\times {\mathbb R}^{p_2+1}\) whose forward evolution \(g_K(t)\) by Ricci flow starting at \(t=-1\) forms a singularity at time \(t=0\). As \(t\nearrow 0\), the metric \(g_K(t)\) converges to a conical metric on \(\mathcal S^{p_1}\times \mathcal S^{p_2}\times (0,\infty )\). Moreover there exist at least K distinct, non-isometric, forward continuations by Ricci flow expanding solitons on \(\mathcal S^{p_1}\times {\mathbb R}^{p_2+1}\), and also at least K non-isometric, forward continuations expanding solitons on \({\mathbb R}^{p_1+1}\times \mathcal S^{p_2}\). In short, there exist smooth complete initial metrics for Ricci flow whose forward evolutions after a first singularity forms are not unique, and whose topology may change at the singularity for some solutions but not for others.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. To obtain the full strength of their results, Kleiner and Lott impose three other conditions (Definition 1.6) on the singular Ricci flows they consider involving 3-manifolds and PIC 4-manifolds. We omit those conditions because they are not relevant to our work on higher-dimensional manifolds in this paper.

  2. Completeness here is understood with respect to Kleiner and Lott’s spacetime metric \(g_{\mathcal M} = {{\hat{g}}} +(\mathrm d\mathfrak t)^2\), where \({{\hat{g}}}\) is the extension of g to a quadratic form on \(T\mathcal M\) satisfying \(\partial _t\in \ker ({{\hat{g}}})\).

  3. This means there is a real-analytic immersion \(\Upsilon :\mathfrak D\rightarrow {W^{u}({{\mathfrak {g}} {\mathfrak {f}}})}\), where \(\mathfrak D\) is an open neighborhood of the origin in \({\mathbb R}^3\), with the property that \(\Upsilon (0)={{\mathfrak {g}} {\mathfrak {f}}}\) and \(g^t(\Upsilon (\varvec{x})) = \Upsilon (e^{2t}\varvec{x})\) for all sufficiently small \(\varvec{x}\in \mathfrak D\).

  4. Throughout this paper, we denote the result of a one-form \(\varrho :{\mathbb R}^6\rightarrow T^*{\mathbb R}^6\) acting on a vector \(V\in T_{{{\mathfrak {p}}}}{\mathbb R}^6\) by \(\varrho _{{{\mathfrak {p}}}}\cdot V\), or just \(\varrho \cdot V\), if the point \({{\mathfrak {p}}}\in {\mathbb R}^6\) of evaluation can be deduced from the context.

  5. If \(V_1\), \(V_2\pm iV_3\), and \(V_4\) are eigenvectors of \(dX_{{\mathfrak {r}}{\mathfrak {f}}{\mathfrak {c}}}\) corresponding to the eigenvalues \(-1\), \(-A\pm i\Omega \), and \(-n+1\), then define \(\langle \cdot ,\cdot \rangle _s\) by declaring \(\{V_1, V_2, V_3, V_4\}\) to be orthonormal.

  6. Palis originally published his \(\lambda \)-lemma in [Pal68]. The more modern, equivalent formulation we use here is Lemma 7.1 in [PdM82, p. 80]. Finally, [FH19, Proposition 6.1.7] proves the \(\lambda \)-lemma or “inclination lemma” for flows rather than maps, which is the closest to our use.

  7. The difference system (30) for \((x_{12}, y_{12})\) implies that \(\zeta \) satisfies a differential equation of the form \(\zeta '=\mathcal L(\zeta )\zeta \), where \(\mathcal L(\zeta ):{\mathbb C}\rightarrow {\mathbb C}\) is a real linear transformation that depends smoothly on \(\zeta \). Every real linear transformation of \({\mathbb C}\) is of the form \(\zeta \mapsto M\zeta +N{{\bar{\zeta }}}\), for suitable \(M,N\in {\mathbb C}\). To verify the differential equation above, we must show that \(M(0)=-A+i\Omega \), and \(N(0)=0\). The coefficients M(0) and N(0) are determined by the linearization \(\zeta '=\mathcal L(0)\zeta \) of (30) at zero. Keeping in mind that \(A=(n-1)/2\) and \(A^2+\Omega ^2=2(n-1)\) (see (27), (28)), we compute this equation in terms of \(\zeta \):

    $$\begin{aligned} \zeta '&=(A+i\Omega )x_{12}'-(A^2+\Omega ^2)y_{12}'\\&=(A+i\Omega )\bigl (-(A^2+\Omega ^2)y_{12}\bigr ) -(A^2+\Omega ^2)(x_{12}-2Ay_{12})\\&=-(A^2+\Omega ^2)\bigl (x_{12}+(-A+i\Omega )y_{12}\bigr )\\&=(-A+i\Omega )\bigl ((A+i\Omega )x_{12}-(A^2+\Omega ^2)y_{12}\bigr )\\&=(-A+i\Omega )\zeta . \end{aligned}$$

    Thus \(\mathcal L(0)\) is complex linear, with \(M(0)=-A+i\Omega \), \(N(0)=0\).

  8. Our notation here is as follows: the first parameter denotes the spherical factor(s) involved, while the second indicates the highest-order derivative that appears.

  9. Compare to Equation (1.8) of [CCGGIIKLLN07], with \(\epsilon =2\lambda \).

  10. The subscript 4 in \(Z_4\) is intended to remind us of the power of 1/s that appears in the definition of \(Z_4\).

References

  1. S.B. Angenent, M.C. Caputo, and D. Knopf. Minimally invasive surgery for Ricci flow singularities. J. Reine Angew. Math. 672 (2012), 39–87

    MathSciNet  MATH  Google Scholar 

  2. S.B. Angenent, T. Ilmanen, and J.J.L. Velázquez. Fattening from smooth initial data in mean curvature flow. In preparation

  3. V.I. Arnol’d. Geometrical methods in the theory of ordinary differential equations. Second edition. Grundlehren der Mathematischen Wissenschaften, 250. Springer, New York (1988).

  4. R.H. Bamler and B. Kleiner. Uniqueness and stability of Ricci flow through singularities. arXiv:1709.04122v1.

  5. C. Böhm. Inhomogeneous Einstein metrics on low-dimensional spheres and other low-dimensional spaces. Invent. Math. (1)134 (1998), 145–176

    Article  MathSciNet  Google Scholar 

  6. C. Böhm. Non-compact cohomogeneity one Einstein manifolds. Bull. Soc. Math. France (1)127 (1999), 135–177

    Article  MathSciNet  Google Scholar 

  7. T. Carson. Ricci flow emerging from rotationally symmetric degenerate neckpinches. Int. Math. Res. Not. (IMRN) (12) (2016), 3678–3716

  8. T. Carson. Ricci flow from some spaces with asymptotically cylindrical singularities. arXiv:1805.09401.

  9. H.-D. Cao. Limits of solutions to the Kähler-Ricci flow. J. Differ. Geom. 45 (1997), 257–272

    Article  Google Scholar 

  10. X. Chen, G. Tian, and Z. Zhang. On the weak Kähler–Ricci flow. Trans. Am. Math. Soc. (6)363 (2011), 2849–2863

    Article  Google Scholar 

  11. B. Chow, S.-C. Chu, D. Glickenstein, C. Guenther, J. Isenberg, T. Ivey, D. Knopf, P. Lu, F. Luo, and L. Ni. The Ricci flow: techniques and applications. Part I, Mathematical Surveys and Monographs, 135. American Mathematical Society, Providence (2007).

  12. C.C. Conley and R.W. Easton. Isolated invariant sets and isolating blocks. Trans. Am. Math. Soc. (1)158 (1971), 35–61

    Article  MathSciNet  Google Scholar 

  13. C.C. Conley. Isolated Invariant Sets and the Morse Index. CBMS Regional Conference Series in Mathematics,38. American Mathematical Society, Providence (1978).

  14. A.S. Dancer, S.J. Hall, and M.Y. Wang. Cohomogeneity one shrinking Ricci solitons: An analytic and numerical study. Asian J. Math., 17 (2013), 33–61

    Article  MathSciNet  Google Scholar 

  15. A.S. Dancer and M.Y. Wang. Non-Kähler expanding Ricci solitons. Int. Math. Res. Not. 6 (2009), 1107–1133

    Article  Google Scholar 

  16. A.S. Dancer and M.Y. Wang. Some new examples of non-Kähler Ricci solitons. Math. Res. Lett. 16:2 (2009), 349–363

    Article  MathSciNet  Google Scholar 

  17. A. Deruelle. Asymptotic estimates and compactness of expanding Ricci solitons. Ann. Sc. Norm. Super. Pisa CL. Sci. (5)XVII (2017), 485–530

  18. P. Eyssidieux, V. Guedj, and A. Zeriahi. Weak solutions to degenerate complex Monge-Ampère flows II. Adv. Math. 293 (2016), 37–80

    Article  MathSciNet  Google Scholar 

  19. M. Feldman, T. Ilmanen, and D. Knopf. Rotationally symmetric shrinking and expanding gradient Kähler–Ricci solitons. J. Differ. Geom. (2)65 (2003), 169–209

    Article  Google Scholar 

  20. T. Fisher and B. Hasselblatt. Hyperbolic flows, Zurich Lectures in Advanced Mathematics (ZLAM). Europe Mathematical Society Publishing House, Zürich (2019).

  21. A. Gastel and M. Kronz. A family of expanding Ricci solitons. Variational problems in Riemannian geometry, 81–93, Progr. Nonlinear Differential Equations Appl., 59. Birkhäuser, Basel (2004).

  22. G. Giesen and P.M. Topping. Existence of Ricci flows of incomplete surfaces. Commun. Partial Differ. Equ. 3610 (2011), 1860–1880

    Article  MathSciNet  Google Scholar 

  23. G. Giesen and P.M. Topping. Ricci flows with unbounded curvature. Math. Z. (1–2)273 (2013), 449–460

    Article  MathSciNet  Google Scholar 

  24. D.M. Grobman. Topological classification of neighborhoods of a singularity in \(n\)-space. (Russian) Mat. Sb. (N.S.) 56(98) (1962), 77–94

  25. R.S. Hamilton. Four-manifolds with positive isotropic curvature. Commun. Anal. Geom. 5(1) (1997), 1–92

    Article  MathSciNet  Google Scholar 

  26. P. Hartman. A lemma in the theory of structural stability of differential equations. Proc. Am. Math. Soc. 11 (1960), 610–620

    Article  MathSciNet  Google Scholar 

  27. R. Haslhofer and A. Naber. Characterizations of the Ricci flow. J. Eur. Math. Soc. (JEMS)(5)20 (2018), 1269–1302

    Article  MathSciNet  Google Scholar 

  28. R. Haslhofer. Uniqueness and stability of singular Ricci flows in higher dimensions. arXiv:2110.03412v2.

  29. T. Ilmanen. Lectures on Mean Curvature Flow and Related Equations. (Trieste notes, originally entitled “Some Techniques in Geometric Heat Flow”). https://people.math.ethz.ch/~ilmanen/papers/notes.pdf.

  30. Y. Ilyashenko and S. Yakovenko. Lectures on Analytic Differential Equations, Graduate Studies in Mathematics86. American Mathematical Society (2008).

  31. T. Ivey. New examples of complete Ricci solitons. Proc. Am. Math. Soc. (1)122 (1994), 241–245

    Article  MathSciNet  Google Scholar 

  32. H. Koch and T. Lamm. Geometric flows with rough initial data. Asian J. Math. (2)16 (2012), 209–235

    Article  MathSciNet  Google Scholar 

  33. B. Kleiner and J. Lott. Singular Ricci flows I. Acta Math. (1)219 (2017), 65–134

    Article  MathSciNet  Google Scholar 

  34. B. Kleiner and J. Lott. Singular Ricci flows II. Chapter in: Chen J., Lu P., Lu Z., Zhang Z. (eds.) Geometric Analysis. Progress in Mathematics, vol. 333. Birkhäuser, Cham (2018).

  35. B. Kotschwar and L. Wang. Rigidity of asymptotically conical shrinking gradient Ricci solitons. J. Differ. Geom. (1)100 (2015), 55–108

    Article  MathSciNet  Google Scholar 

  36. R.J. McCann and P.M. Topping. Ricci flow, entropy and optimal transportation. Am. J. Math. (3)132 (2010), 711–730

    Article  MathSciNet  Google Scholar 

  37. F.W.J. Olver. Uniform, exponentially improved, asymptotic expansions for the confluent hypergeometric function and other integral transforms. SIAM J. Math. Anal. (5)22 (1991), 1475–1489

    Article  MathSciNet  Google Scholar 

  38. J. Palis. On Morse–Smale dynamical systems. Topology 8 (1968), 385–404

    Article  MathSciNet  Google Scholar 

  39. J. Palis and W. de Melo.Geometric theory of dynamical systems. An introduction. Translated from the Portuguese by A. K. Manning. Springer-Verlag, New York-Berlin, 1982. xii+198 pp. ISBN: 0-387-90668-1.

  40. G. Perelman. The entropy formula for the Ricci flow and its geometric applications. arXiv:math/0211159.

  41. G. Perelman. Ricci flow with surgery on three- manifolds. arXiv:math/0303109.

  42. P. Petersen. Riemannian geometry. Third edition. Graduate Texts in Mathematics, 171. Springer, Cham (2016).

  43. P. Petersen and W. Wylie. Rigidity of gradient Ricci solitons. Pac. J. Math. (2)241 (2009), 329–345

    Article  MathSciNet  Google Scholar 

  44. M. Simon. Deformation of \(C^0\) Riemannian metrics in the direction of their Ricci curvature. Commun. Anal. Geom. 10(5) (2002), 1033–1047

    Article  MathSciNet  Google Scholar 

  45. M. Simon. Ricci flow of almost non-negatively curved three-manifolds. J. Reine Angew. Math. 630 (2009), 177–217

    MathSciNet  MATH  Google Scholar 

  46. M. Simon. Ricci flow of non-collapsed three-manifolds whose Ricci curvature is bounded from below. J. Reine Angew. Math. 662 (2012), 59–94

    MathSciNet  MATH  Google Scholar 

  47. F. Schulze and M. Simon. Expanding solitons with non-negative curvature operator coming out of cones. Math. Z. 275 (2013), 625–639

    Article  MathSciNet  Google Scholar 

  48. J. Song and G. Tian. The Kähler–Ricci flow through singularities. Invent. Math. (2)207 (2017), 519–595

    Article  MathSciNet  Google Scholar 

  49. K.-T. Sturm. Super-Ricci Flows for Metric Measure Spaces. arXiv:1603.02193

  50. P.M. Topping. Ricci flow compactness via pseudolocality, and flows with incomplete initial metrics. J. Eur. Math. Soc. (JEMS) (6)12 (2010), 1429–1451

    Article  MathSciNet  Google Scholar 

  51. P.M. Topping. Uniqueness and nonuniqueness for Ricci flow on surfaces: Reverse cusp singularities. Int. Math. Res. Not. (IMRN) 10 (2012), 2356–2376

    MathSciNet  MATH  Google Scholar 

  52. P.M. Topping. Uniqueness of instantaneously complete Ricci flows. Geom. Topol. (3)19 (2015), 1477–1492

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

DK thanks the NSF for support (DMS-1205270) during early work on this project. Both authors thank the Mathematisches Forschungsinstitut Oberwolfach for its hospitality in the 2016 Geometrie workshop, during which they made further progress on the project. Finally, the authors thank the anonymous referees for their careful reading of the manuscript and several detailed suggestions to improve the exposition.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sigurd B. Angenent.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A. Doubly-Warped Product Geometries

1.1 The general case.

It is well known (see, e.g., [Pet16]) that all curvatures of a doubly-warped product metric

$$\begin{aligned} g = (\mathrm ds)^2 + \varphi _1^2\,g_{\mathcal S^{p_1}} + \varphi _2^2\,g_{\mathcal S^{p_2}} \end{aligned}$$

on \(\mathbb R_+\times \mathcal S^{p_1}\times \mathcal S^{p_2}\) are convex linear combinations of the five sectional curvaturesFootnote 8

$$\begin{aligned} \kappa _{\alpha ,1}=\frac{1-\varphi _{\alpha ,s}^2}{\varphi _\alpha ^2},\quad \kappa _{\alpha ,2}=-\frac{\varphi _{\alpha ,ss}}{\varphi _\alpha },\quad \kappa _{12,1}=-\frac{\varphi _{1,s} \varphi _{2,s}}{\varphi _1\varphi _2}. \end{aligned}$$

We note that the functions \(\kappa _{\alpha ,1}\) (\(\alpha \in \{1,2\}\)) are the sectional curvatures of orthonormal planes tangent to \(\mathcal S^{p_\alpha }\). The functions \(\kappa _{\alpha ,2}\) are the sectional curvatures of orthonormal planes spanned by \(\frac{\partial }{\partial s}\) and vectors tangent to \(\mathcal S^{p_\alpha }\). And \(\kappa _{12,1}\) is the sectional curvature of a plane spanned by one vector tangent to \(\mathcal S^{p_1}\) and one tangent to \(\mathcal S^{p_2}\).

It follows easily that the Ricci tensor of such a metric is given by

$$\begin{aligned} {{\,\mathrm{Rc}\,}}&=\big \{p_1\kappa _{1,2} +p_2\kappa _{2,2}\big \}(\mathrm ds)^2\\&\quad +\big \{\kappa _{1,2}+ (p_1-1)\kappa _{1,1}+p_2\kappa _{12,1}\big \}\varphi _1^2g_{\mathcal S^{p_1}}\\&\quad +\big \{\kappa _{2,2}+(p_2-1)\kappa _{2,1}+p_1\kappa _{12,1}\big \}\varphi _2^2g_{\mathcal S^{p_2}}. \end{aligned}$$

Now let \(\mathfrak X = f(s)\,\frac{\partial }{\partial s}\) denote the gradient of a potential function F(s). Applying the general formula

$$\begin{aligned} (\mathcal L_{\mathfrak X}g)(V_1,V_2) = \mathfrak X\big \{g(V_1,V_2)\big \}+g(\nabla _{V_1}\mathfrak X,V_2)+g(V_1,\nabla _{V_2}\mathfrak X) \end{aligned}$$

for the Lie derivative of a covariant 2-tensor to this special case, one sees that

$$\begin{aligned} \mathcal L_{\mathfrak X} g = 2f_s(\mathrm d s)^2+2f\varphi _1\varphi _{1,s}\,g_{\mathcal S^{p_1}}+2f\varphi _2\varphi _{2,s}\,g_{\mathcal S^{p_2}}. \end{aligned}$$

In the main body of this work, we apply the equations above to the soliton condition

$$\begin{aligned} -2{{\,\mathrm{Rc}\,}}[g] = 2\lambda g + \mathcal L_{\mathfrak X}g, \end{aligned}$$

were \(\lambda \in \{-1,0,+1\}\) controls the rescaling of the soliton.Footnote 9

1.2 Cones.

In the case of a cone metric,

$$\begin{aligned} \varphi _1= s\,\sqrt{\frac{p_1-1}{{{\bar{x}}}_1}} \qquad \text{ and }\qquad \varphi _2= s\, \sqrt{\frac{p_2-1}{{{\bar{x}}}_2}} \end{aligned}$$

with asymptotic apertures \({{\bar{x}}}_1,\,{{\bar{x}}}_2\), the sectional curvature \(\kappa _{12,1}\) is given by

$$\begin{aligned} \kappa _{12,1} = -\frac{1}{s^2}. \end{aligned}$$

In particular, the norm of the curvature tensor of the cone becomes unbounded as \(s\searrow 0\).

Appendix B. A Representation as a Mechanical System on \({\mathbb R}^3\)

As a curiosity, we observe that our assumption that \(p_\alpha \ge 2\) allows us to define

$$\begin{aligned} u_\alpha = \log \varphi _\alpha -\tfrac{1}{2} \ln (p_\alpha -1),\quad \alpha \in \{1,2\}, \qquad \text { and }\qquad v = f - p_1\frac{\mathrm du_1}{\mathrm ds} - p_2\frac{\mathrm du_2}{\mathrm ds}. \end{aligned}$$

In these variables, the differential equations (8) become

$$\begin{aligned} {\ddot{u}}_\alpha&= v \dot{u}_\alpha + e^{-2u_\alpha }+\lambda ,\qquad \alpha \in \{1,2\}, \end{aligned}$$
(102a)
$$\begin{aligned} \dot{v}&= p_1\dot{u}_1^2 + p_2\dot{u}_2^2 - \lambda . \end{aligned}$$
(102b)

These equations can be interpreted as a mechanical system in which s is “time” and where (in this section only) we write s-derivatives as fluxions. In this interpretation, \(u_1\) and \(u_2\) are the coordinates of two unit-mass particles on the real line that are each subject to a force field given by \(F(u) = e^{-2u}+\lambda \), and whose motion is subject to friction with friction coefficient v. The only unusual aspect of this system from the point of view of mechanics is that the friction coefficient v can be either positive or negative, and that it is itself a function of time that satisfies an ode.

The derivation of the equations for \(u_1, u_2, v\) from (8) is a simple calculus exercise. Even though it appears simpler than the original equations (8), we will not use the mechanical system (102) in this paper. It does however make several cameo appearances. For example, the Ivey invariant for the stationary soliton flow can be interpreted as the energy dissipation in the mechanical system. Indeed, if \(\lambda =0\), then any solution of (102) satisfies

$$\begin{aligned} \frac{\mathrm d}{\mathrm ds}\left\{ \frac{p_1}{2} \bigl (\dot{u}_1^2 + e^{-2u_1} \bigr ) + \frac{p_2}{2} \bigl (\dot{u}_2^2 + e^{-2u_1} \bigr ) \right\} = v\dot{v} = \frac{\mathrm d\frac{1}{2} v^2}{\mathrm ds}, \end{aligned}$$

which implies that the quantity \(I = p_1(\dot{u}_1^2+e^{-2u_1}) + p_2(\dot{u}_2^2+e^{-2u_2}) - v^2\) is preserved along solutions of (102). Similarly, the non-obvious Lyapunov function W in Gastel and Kronz’ construction of the Böhm soliton (see Section 7.1) can be interpreted as Kinetic\(+\)Potential Energy for a renormalized version of the mechanical system (102).

Appendix C. An Estimate for Orbits Near a Hyperbolic Fixed Point

1.1 A model nonlinear system.

Consider a system

$$\begin{aligned} x_-' = -A_- x_- + B_-(x)x, \qquad x_+' = A_+ x_+ + B_+(x)x, \end{aligned}$$
(103)

where \(x = (x_-, x_+) \in {\mathbb R}^{k_-}\times {\mathbb R}^{k_+}\), where \(A_- : {\mathbb R}^{k_-} \rightarrow {\mathbb R}^{k_-} \), \(A_+ : {\mathbb R}^{k_+}\rightarrow {\mathbb R}^{k_+} \) are constant linear maps, and where \(B_\pm \) are smooth functions on some neighborhood of the origin in \({\mathbb R}^{k_-+k_+} \) such that \(B_-(x) \) is a linear map from \({\mathbb R}^{k_-+k_+} \) to \({\mathbb R}^{k_-}\) and \(B_+(x) \) is a linear map from \({\mathbb R}^{k_-+k_+} \) to \({\mathbb R}^{k_+}\). We assume furthermore that \(B_\pm (0) = 0\).

The origin (0, 0) is a fixed point for our system (103). The linearization of this system at the origin has the matrix

$$\begin{aligned} \begin{bmatrix} -A_- &{} 0 \\ 0 &{} A_+ \end{bmatrix}. \end{aligned}$$

We make one more assumption, namely that the eigenvalues of both \(A_\pm \) all have strictly positive real parts.

1.2 An analysis lemma.

There is a constant \(C\in {\mathbb R}\) that only depends on the matrices \(A_\pm \) and the nonlinear functions \(B_\pm \), such that for all \(T>0\) and for any solution \(x:[0, T]\rightarrow {\mathbb R}^{k_-+k_+}\) of (103) such that \(\sup _{0\le t\le T} \Vert x(t)\Vert \) is sufficiently small, one has

$$\begin{aligned} \Vert x(t)\Vert \le C\bigl ( e^{-\epsilon t} + e^{-\epsilon (T-t)} \bigr )\sup _{0\le t\le T} \Vert x(t)\Vert \end{aligned}$$

and

$$\begin{aligned} \int _0^T \Vert x(t)\Vert \, {\mathrm d}t \le C \sup _{0\le t\le T} \Vert x(t)\Vert . \end{aligned}$$

Proof

Briefly, we use a Gronwall-type argument to establish an exponential upper bound for \(\Vert x(t)\Vert \) in the interval [0, T], and then integrate this upper bound to get the claimed estimate.

There is a \(\delta > 0 \) such that all eigenvalues \( \mu \) of \(A_+\) and \(A_-\) satisfy \({\mathrm {Re}}\mu \ge \delta \). Furthermore, there is a constant \(C_A>0\) such that

$$\begin{aligned} \Vert e^{-tA_\pm }\Vert \le C_A e^{-\delta t} \end{aligned}$$

holds for all \(t\ge 0\). Applying the variation of constants formula to the system (103), we find that on the interval [0, T], both \(x_+\) and \(x_-\) are given by

$$\begin{aligned} x_-(t)&= e^{-t A_-}x_-(0) + \int _0^t e^{-(t-s)A_-} B_-(x(s))x(s) \,{\mathrm d}s, \\ x_+(t)&= e^{-(T-t)A_+} x_+(T) - \int _t^T e^{-(s-t)A_+} B_+(x(s))x(s)\, {\mathrm d}s. \end{aligned}$$

Since \(B_\pm (0) = 0\), there is a constant \(C_B>0\) such that \(\Vert B_\pm (x)\Vert \le C_B\Vert x\Vert \) holds for all sufficiently small x. Thus we get

$$\begin{aligned} \Vert x_-(t)\Vert&\le C_A e^{-\delta t}\Vert x_-(0)\Vert + C_AC_B \int _0^t e^{-\delta (t-s)} \Vert x(s)\Vert ^2 \,{\mathrm d}s, \\ \Vert x_+(t)\Vert&\le C_A e^{-\delta (T-t)} \Vert x_+(T) \Vert + C_AC_B \int _t^T e^{-\delta (s-t)} \Vert x(s)\Vert ^2 \,{\mathrm d}s. \end{aligned}$$

For K to be fixed below, we may assume that \(\Vert x(s)\Vert \le K\) for all \(s\in [0,T]\), whence we get

$$\begin{aligned} \Vert x_-(t)\Vert&\le C_AKe^{-\delta t} + C_AC_BK \int _0^t e^{-\delta (t-s)} \Vert x(s)\Vert \,{\mathrm d}s, \\ \Vert x_+(t)\Vert&\le C_AKe^{-\delta (T-t)} + C_AC_BK \int _t^T e^{-\delta (s-t)} \Vert x(s)\Vert \,{\mathrm d}s. \end{aligned}$$

After adding these two inequalities, we find there exists \(C_0=C_0(C_A,C_B)\) such that that for all \(t\in [0, T]\), one has

$$\begin{aligned} \Vert x(t)\Vert&\le \Vert x_-(t)\Vert + \Vert x_+(t)\Vert \nonumber \\&\le C_0K \bigl (e^{-\delta t} + e^{-\delta (T-t)}\bigr ) + C_0K \int _0^T e^{-\delta |s-t|} \Vert x(s)\Vert \,{\mathrm d}s. \end{aligned}$$
(104)

If we define

$$\begin{aligned} \rho (t) = \frac{1}{2\delta } \int _0^T e^{-\delta |s-t|} \Vert x(s)\Vert \,{\mathrm d}s, \end{aligned}$$

then since \(G(t) = (2\delta )^{-1}e^{-\delta |t|}\) is the Green’s function for \(-\frac{{\mathrm d}^2}{{\mathrm d}t^2} + \delta ^2\), the quantity \(\rho \) satisfies

$$\begin{aligned} -\rho ''(t) + \delta ^2 \rho (t) = \Vert x(t)\Vert \end{aligned}$$

for all \(t\in (0, T)\). We can therefore rewrite the integral inequality (104) as

$$\begin{aligned} -\rho ''(t) + \bigl (\delta ^2 - 2\delta C_0K \bigr )\rho (t) \le C_0 \bigl (e^{-\delta t} + e^{-\delta (T-t)}\bigr ), \qquad (0<t<T). \end{aligned}$$
(105)

Moreover, we have the boundary conditions

$$\begin{aligned} \rho (0) = \frac{1}{2\delta }\int _0^Te^{-\delta t}\Vert x(t)\Vert \,{\mathrm d}t \le \frac{K}{2\delta ^2}(1-e^{-\delta T})\quad \text{ and }\quad \rho (T) \le \frac{K}{2\delta ^2}(1-e^{-\delta T}). \nonumber \\ \end{aligned}$$
(106)

For any constant M, the function \({{\bar{\rho }}}(t) = M\bigl (e^{-\epsilon t}+e^{-\epsilon (T-t)}\bigr )\) satisfies \({{\bar{\rho }}}\,'' = \epsilon ^2{{\bar{\rho }}} \), so that

$$\begin{aligned} -{{\bar{\rho }}}\,''(t) +\bigl (\delta ^2 - 2\delta C_0K\bigr ){{\bar{\rho }}} = \bigl (\delta ^2 - 2\delta C_0K - \epsilon ^2\bigr ) {{\bar{\rho }}}. \end{aligned}$$

If we now choose K small enough that \(2C_0K < \frac{1}{2} \delta \) and then choose \(\epsilon = \frac{1}{2} \delta \), we have

$$\begin{aligned} \delta ^2 - 2\delta CK - \epsilon ^2 > \tfrac{1}{2} \delta ^2 - \epsilon ^2 = \tfrac{1}{4} \delta ^2. \end{aligned}$$

So the function \({{\bar{\rho }}}\) becomes a supersolution of the boundary-value problem (105, 106) provided that \(M=C_1K\), where \(C_1\) is a constant that depends on \(\delta ,\epsilon \), and \(C_0\). By the maximum principle, we conclude that \(\rho \le {{\bar{\rho }}}\), and thus that

$$\begin{aligned} \rho (t) \le C_1 K \bigl ( e^{-\epsilon t} + e^{-\epsilon (T-t)} \bigr ). \end{aligned}$$

Applying this to (104) and using the fact that \(\epsilon < \delta \), we find

$$\begin{aligned} e^{-\delta t} + e^{-\delta (T-t)} \le e^{-\epsilon t} + e^{-\epsilon (T-t)}. \end{aligned}$$

Since we may assume that \(K\le 1\), it follows that

$$\begin{aligned} \Vert x(t)\Vert \le (C_2K + C_2K^2) \bigl ( e^{-\epsilon t} + e^{-\epsilon (T-t)} \bigr ) \le C_3K \bigl ( e^{-\epsilon t} + e^{-\epsilon (T-t)} \bigr ). \end{aligned}$$

We complete the proof by integrating over [0, T], obtaining

$$\begin{aligned} \int _0^T \Vert x(t)\Vert {\mathrm d}t \le \frac{2C_3}{\epsilon } K = C_4 K, \end{aligned}$$

where the constant C only depends on \(\delta \), \(C_A\), and \(C_B\), but not on T. \(\square \)

Appendix D. Asymptotics of \(\chi (s)\) as \(s\rightarrow \infty \)

Asymptotic expansions for the solutions \(\chi \) of (39) are well documented and can be derived in a number of ways. Here we indicate one possible real-variable approach.

If \(\chi :{\mathbb R}\rightarrow {\mathbb R}\) is a solution of (39), namely

$$\begin{aligned} \chi _{ss} + \left( \frac{n}{s}+\lambda s\right) \chi _s + \frac{2(n-1)}{s^2} \chi = 0, \end{aligned}$$

then the function

$$\begin{aligned} Z(s) = \frac{\chi _s(s)}{s\chi (s)} \end{aligned}$$

satisfies

$$\begin{aligned} \mathcal G(s,Z) {\mathop {=}\limits ^{{\mathrm{def}}}} \frac{1}{s}\frac{{\mathrm d}Z}{{\mathrm d}s} = -\frac{2(n-1)}{s^4} - \frac{n+1}{s^2}Z - \lambda Z - Z^2. \end{aligned}$$
(107)

As \(s\rightarrow \infty \), this equation becomes

$$\begin{aligned} \frac{{\mathrm d}Z}{{\mathrm d}(s^2/2)} = - Z(Z+\lambda ) + {{\mathcal {O}}}(s^{-2}), \end{aligned}$$

which has two constant (approximate) solutions, \(Z=0\) and \(Z=-\lambda \).

Direct substitution reveals thatFootnote 10\(Z_4(s) {\mathop {=}\limits ^{{\mathrm{def}}}} 2(n-1)/s^4\) satisfies

$$\begin{aligned} \frac{1}{s}\frac{{\mathrm d}Z_4}{{\mathrm d}s} - \mathcal G(s, Z_4(s)) = {{\mathcal {O}}}(s^{-6}), \qquad (s\rightarrow \infty ), \end{aligned}$$

while

$$\begin{aligned} Z_4^{\pm }(s) {\mathop {=}\limits ^{{\mathrm{def}}}} \frac{2(n-1)\pm 1}{s^4} \end{aligned}$$

satisfies

$$\begin{aligned} \frac{1}{s}\frac{{\mathrm d}Z_4^\pm }{{\mathrm d}s} - \mathcal G(s, Z_4^\pm (s)) = \pm \frac{\lambda }{s^4} + {{\mathcal {O}}}(s^{-6}), \qquad (s\rightarrow \infty ). \end{aligned}$$

If \(\lambda >0\), this implies that \(Z_4^-(s) < Z_4^+(s)\) are lower and upper barriers for the ode (107), and therefore that there is a solution Z(s) with \(Z_4^-(s) \le Z(s) \le Z_4^+(s)\) for large s. In fact, if \(s_0\gg 1\), then any solution Z(s) of (107) that satisfies \(Z_4^-(s_0) \le Z(s_0) \le Z_4^+(s_0)\) will continue to satisfy \(Z_4^-(s) \le Z(s) \le Z_4^+(s)\) for all \(s\ge s_0\).

If \(\lambda <0\), then \(Z_4^-(s)\) is an upper barrier, and \(Z_4^+(s)\) is a lower barrier. Since \(Z_4^-(s) < Z_4^+(s)\) for all \(s\ge s_0\) if \(s_0\) is large enough, we can apply a Ważewski argument and conclude that there exists at least one \(Z^*\in \big (Z_4^-(s_0) , Z_4^+(s_0)\big )\) such that the solution of (107) with \(Z(s_0) = Z^*\) satisfies \(Z_4^-(s) \le Z(s) \le Z_4^+(s)\) for all \(s\ge s_0\).

In either case, the conclusion is that there exists a solution Z(s) of (107) with

$$\begin{aligned} Z(s) = \bigl (2(n-1)+{{\mathcal {O}}}(1)\bigr ) s^{-4}, \qquad (s\rightarrow \infty ). \end{aligned}$$

By repeating this argument, one finds that for any \(m\in {\mathbb N}\), there exists a solution that satisfies the expansion

$$\begin{aligned} Z(s) = \frac{A_4}{s^4} + \frac{A_6}{s^6} + \frac{A_8}{s^8} + \cdots + \frac{A_{2m}}{s^{2m}} + {{\mathcal {O}}}\bigl (s^{-2m-2}\bigr ), \qquad (s\rightarrow \infty ). \end{aligned}$$

The coefficients \(A_{2j}\) can be computed inductively by substituting the formal expansion; one finds for example that \(A_4 = 2\lambda (n-1)\).

Integration then shows that \(\chi \) satisfies

$$\begin{aligned} \chi (s) = e^ {\int Z(s) s{\mathrm d}s} = e^{C - \lambda \frac{n-1}{s^2} + {{\mathcal {O}}}(s^{-4})} = e^C \Bigl \{1 - \lambda \frac{n-1}{s^2} + {{\mathcal {O}}}(s^{-4})\Bigr \}, \qquad (s\rightarrow \infty ). \end{aligned}$$

As we noted above, there exists another solution \({{\tilde{Z}}}\) such that \({{\tilde{Z}}}(s)= -\lambda + o(1)\) for large s. Similar reasoning then leads to an expansion of the form

$$\begin{aligned} {{\tilde{Z}}}(s) = -\lambda - \frac{n+1}{s^2} + \frac{B_4}{s^4} + \cdots , \end{aligned}$$

which after integration leads to

$$\begin{aligned} \chi (s) = e^{-\lambda s^2/2} s^{-n-1} \bigl \{1 + {{\mathcal {O}}}(s^{-2})\bigr \}, \qquad (s\rightarrow \infty ). \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Angenent, S.B., Knopf, D. Ricci Solitons, Conical Singularities, and Nonuniqueness. Geom. Funct. Anal. 32, 411–489 (2022). https://doi.org/10.1007/s00039-022-00601-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00039-022-00601-y

Navigation