Skip to main content

Advertisement

Log in

On the Morse–Bott property of analytic functions on Banach spaces with Łojasiewicz exponent one half

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

It is a consequence of the Morse–Bott Lemma (see Theorems 2.10 and 2.14) that a \(C^2\) Morse–Bott function on an open neighborhood of a critical point in a Banach space obeys a Łojasiewicz gradient inequality with the optimal exponent one half. In this article we prove converses (Theorems 1, 2, and Corollary 3) for analytic functions on Banach spaces: If the Łojasiewicz exponent of an analytic function is equal to one half at a critical point, then the function is Morse–Bott and thus its critical set nearby is an analytic submanifold. The main ingredients in our proofs are the Łojasiewicz gradient inequality for an analytic function on a finite-dimensional vector space (Łojasiewicz in Ensembles Semi-analytiques, Institut des Hautes Etudes Scientifiques, Bures-sur-Yvette, 1965) and the Morse Lemma (Theorems 4 and 5) for functions on Banach spaces with degenerate critical points that generalize previous versions in the literature, and which we also use to give streamlined proofs of the Łojasiewicz–Simon gradient inequalities for analytic functions on Banach spaces (Theorems 9 and 10).

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.

Similar content being viewed by others

Notes

  1. The first page number refers to the version of Łojasiewicz’s original manuscript mimeographed by IHES while the page number in parentheses refers to the cited LaTeX version of his manuscript prepared by M. Coste and available on the Internet.

  2. We omit the pair of possible signs, ±, when \(\mathbb {K}=\mathbb {C}\).

  3. I am indebted to Michael Greenblatt and András Némethi for pointing out to me that this should be a key analytical tool.

  4. For example, see the discussion prior to Theorem 2.10.

  5. As explained in Remark 1.6, this assumption is automatically obeyed for the same reasons as in Definition 1.5.

  6. The property that \({\text {Ker}}f''(x_0)\subset {\mathscr {X}}\) has a closed complement \({\mathscr {X}}_0 \subset {\mathscr {X}}\) is automatically obeyed for the reasons explained in Remark 1.6.

  7. I am grateful to Otis Chodosh for reminding me of this condition.

  8. I am grateful to Leon Simon for explaining the relevant key ideas from [4, 108, 109].

  9. By relaxing the hypotheses that the continuous embedding \({\mathscr {X}}\subset {\mathscr {X}}^*\) be definite, that one has a continuous embedding \({\mathscr {X}}\subset \tilde{{\mathscr {X}}}\), and that \(\mathbb {K}=\mathbb {R}\).

  10. I am grateful to Graeme Wilkin for drawing my attention to the results of Simpson and Goldman–Millson described here.

  11. To avoid notational clutter, we omit explicit notation, such as \(\iota :\tilde{{\mathscr {X}}} \subset {\mathscr {X}}^*\), for the continuous embedding.

  12. Because \(AA^{-1}={\mathrm {id}}_{\mathscr {X}}=A^{-1}A\) and by [103, Exercise 4.8], one has \((A^{-1})^*A^*={\mathrm {id}}_{\mathscr {X}}=A^*(A^{-1})^*\), so \((A^*)^{-1}=(A^{-1})^*\).

  13. Lang [73, Theorem 5.2] and Palais [95, p. 969] use a power series argument to define F rather than apply the Implicit Mapping Theorem for analytic maps.

  14. In other words, f is Morse at the point \(0 \in {\mathscr {X}}\).

  15. Note that \({\mathscr {K}}\) is finite-dimensional.

  16. While for clarity we have restricted our attention in this article to functions f which are \(C^{p+2}\) with \(p\ge 1\), the Morse–Bott Lemma holds for \(C^2\) functions on Euclidean space: see Banyaga and Hurtubise [13, Theorem 2].

  17. This appendix is a revised version of Feehan and Maridakis [50, Appendix A].

References

  1. Abraham, R.H.: Lectures of Smale on differential topology, unpublished manuscript, (1963)

  2. Abraham, R.H.: Bumpy metrics, Global analysis. In: Proceedings of Symposia in Pure Mathematics, vol. XIV, pp. 1–3, Berkeley, California, 1968. American Mathematical Society, Providence (1970)

  3. Abraham, R.H., Marsden, J.E., Ratiu, T.S.: Manifolds, Tensor Analysis, and Applications, 2nd edn. Springer, New York (1988)

    MATH  Google Scholar 

  4. Adams, D., Simon, L.: Rates of asymptotic convergence near isolated singularities of geometric extrema. Indiana Univ. Math. J. 37, 225–254 (1988)

    MathSciNet  MATH  Google Scholar 

  5. Adams, R.A., Fournier, J.J.F.: Sobolev Space, 2nd edn. Elsevier/Academic Press, Amsterdam (2003)

    MATH  Google Scholar 

  6. Allard, W.K., Almgren Jr., F.J.: On the radial behavior of minimal surfaces and the uniqueness of their tangent cones. Ann. Math. (2) 113(2), 215–265 (1981)

    MathSciNet  MATH  Google Scholar 

  7. Andrews, B., Baker, C.: Mean curvature flow of pinched submanifolds to spheres. J. Differ. Geom. 85(3), 357–395 (2010)

    MathSciNet  MATH  Google Scholar 

  8. Ang, D.D., Tuan, V.T.: An elementary proof of the Morse–Palais lemma for Banach spaces. Proc. Am. Math. Soc. 39, 642–644 (1973)

    MathSciNet  MATH  Google Scholar 

  9. Antoine, P.: Lemme de Morse et calcul des variations. Bull. Soc. Math. France Mém. (1979), no. 60, pp. 25–29, Analyse non convexe (Proceedings of the Pau Colloquium 1977)

  10. Arnold, V.I., Gusein-Zade, S.M., Varchenko, A.N.: Singularities of differentiable maps. Volume 1, modern Birkhäuser Classics, Birkhäuser/Springer, New York (2012). Classification of critical points, caustics and wave fronts, Translated from the Russian by Ian Porteous based on a previous translation by Mark Reynolds, Reprint of the 1985 edition

  11. Austin, D.M., Braam, P.J.: Morse–Bott theory and equivariant cohomology, The Floer memorial volume, Progr. Math., vol. 133, pp. 123–183. Birkhäuser, Basel (1995)

  12. Baker, C.: The Mean Curvature Flow of Submanifolds of High Codimension. Australian National University, Canberra (2010). arXiv:1104.4409

    Google Scholar 

  13. Banyaga, A., Hurtubise, D.E.: A proof of the Morse–Bott lemma. Expo. Math. 22(4), 365–373 (2004)

    MathSciNet  MATH  Google Scholar 

  14. Bellettini, G.: Lecture Notes on Mean Curvature Flow, Barriers and Singular Perturbations, Appunti. Scuola Normale Superiore di Pisa (Nuova Serie) [Lecture Notes. Scuola Normale Superiore di Pisa (New Series)], vol. 12. Edizioni della Normale, Pisa (2013)

    Google Scholar 

  15. Berger, M.: Nonlinearity and Functional Analysis. Academic Press, New York (1977)

    MATH  Google Scholar 

  16. Bierstone, E., Milman, P.D.: Semianalytic and Subanalytic Sets, vol. 67, pp. 5–42. Institut des Hautes Hautes Études Scientifiques, Paris (1988)

    MATH  Google Scholar 

  17. Bierstone, E., Milman, P.D.: Canonical desingularization in characteristic zero by blowing up the maximum strata of a local invariant. Invent. Math. 128(2), 207–302 (1997)

    MathSciNet  MATH  Google Scholar 

  18. Bolton, J., Fernández, L.: On the Regularity of the Space of Harmonic 2-Spheres in the 4-Sphere, Harmonic Maps and Differential Geometry, Contemporary Mathematics, vol. 542, pp. 187–194. American Mathematical Society, Providence (2011)

    MATH  Google Scholar 

  19. Bott, R.H.: Nondegenerate critical manifolds. Ann. Math. (2) 60, 248–261 (1954)

    MathSciNet  MATH  Google Scholar 

  20. Bott, R.H.: The stable homotopy of the classical groups. Ann. Math. (2) 70, 313–337 (1959)

    MathSciNet  MATH  Google Scholar 

  21. Brézis, H.: Functional Analysis, Sobolev Spaces and Partial Differential Equations, Universitext. Springer, New York (2011)

    MATH  Google Scholar 

  22. Bröcker, T.: Differentiable Germs and Catastrophes, Cambridge University Press, Cambridge-New York-Melbourne: Translated from the German, last chapter and bibliography by L, p. 17. London Mathematical Society Lecture Note Series, No, Lander (1975)

  23. Bruveris, M.: Notes on Riemannian geometry of manifolds of maps, unpublished manuscript (2016)

  24. Calegari, D.: Topics in differential geometry minimal submanifolds, unpublished lecture notes

  25. Cambini, A.: Sul lemma di M. Morse. Boll. Un. Mat. Ital. (4) 7, 87–93 (1973)

    MathSciNet  MATH  Google Scholar 

  26. Caps, O.: Evolution Equations in Scales of Banach Spaces, Teubner-Texte zur Mathematik [Teubner Texts in Mathematics], vol. 140. B. G. Teubner, Stuttgart (2002)

    MATH  Google Scholar 

  27. Carlotto, A., Chodosh, O., Rubinstein, Y.A.: Slowly converging Yamabe flows. Geom. Topol. 19(3), 1523–1568 (2015). arXiv:1401.3738

    MathSciNet  MATH  Google Scholar 

  28. Chern, S.S., Goldberg, S.I.: On the volume decreasing property of a class of real harmonic mappings. Am. J. Math. 97, 133–147 (1975)

    MathSciNet  MATH  Google Scholar 

  29. Chill, R.: On the Łojasiewicz–Simon gradient inequality. J. Funct. Anal. 201, 572–601 (2003)

    MathSciNet  MATH  Google Scholar 

  30. Colding, T.H., Minicozzi II, W.P.: Minimal submanifolds. Bull. Lond. Math. Soc. 38(3), 353–395 (2006)

    MathSciNet  MATH  Google Scholar 

  31. Colding, T.H., Minicozzi II, W.P.: Łojasiewicz inequalities and applications. Surv. Differ. Geom. XIX, 63–82 (2014)

    MATH  Google Scholar 

  32. Colding, T.H., Minicozzi II, W.P.: Uniqueness of blowups and Łojasiewicz inequalities. Ann. Math. (2) 182(1), 221–285 (2015)

    MathSciNet  MATH  Google Scholar 

  33. Colding, T.H., Minicozzi II, W.P., Pedersen, E.K.: Mean curvature flow. Bull. Am. Math. Soc. (N.S.) 52(2), 297–333 (2015)

    MathSciNet  MATH  Google Scholar 

  34. Crawford, T.A.: The space of harmonic maps from the \(2\)-sphere to the complex projective plane. Can. Math. Bull. 40(3), 285–295 (1997)

    MathSciNet  MATH  Google Scholar 

  35. Dajczer, M., Tojeiro, R.: Submanifold Theory. Springer, New York (2019). (beyond an introduction)

    MATH  Google Scholar 

  36. Deimling, K.: Nonlinear Functional Analysis. Springer, Berlin (1985)

    MATH  Google Scholar 

  37. DeTurck, D.M.: Deforming metrics in the direction of their Ricci tensors. J. Differ. Geom. 18, 157–162 (1983)

    MathSciNet  MATH  Google Scholar 

  38. Dierkes, U., Hildebrandt, S., Sauvigny, F.: Minimal Surfaces, 2nd edn., Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 339. Springer, Heidelberg. With assistance and contributions by A.Küster and R. Jakob (2010)

  39. Donaldson, S.K., Kronheimer, P.B.: The Geometry of Four-Manifolds. Oxford University Press, New York (1990)

    MATH  Google Scholar 

  40. Ecker, K.: Regularity Theory for Mean Curvature Flow, Progress in Nonlinear Differential Equations and their Applications, vol. 57. Birkhäuser Boston Inc., Boston (2004)

    Google Scholar 

  41. Eells, J., Lemaire, L.: Selected Topics in Harmonic Maps, CBMS Regional Conference Series in Mathematics, vol. 50. American Mathematical Society, Providence (1983)

    MATH  Google Scholar 

  42. Eichhorn, J.: The manifold structure of maps between open manifolds. Ann. Glob. Anal. Geom. 11, 253–300 (1993)

    MathSciNet  MATH  Google Scholar 

  43. Elĭasson, H.I.: Geometry of manifolds of maps. J. Differ. Geom. 1, 169–194 (1967)

    MathSciNet  MATH  Google Scholar 

  44. Feehan, P.M.N.: Global existence and convergence of solutions to gradient systems and applications to Yang–Mills gradient flow, xx+475 pages. arXiv:1409.1525v4

  45. Feehan, P.M.N.: Integrability of Jacobi vectors for analytic functions on Banach manifolds, preprint

  46. Feehan, Paul M. N.: Morse theory for the Yang–Mills energy function near flat connections, arXiv:1906.03954

  47. Feehan, Paul M. N.: Optimal Łojasiewicz–Simon inequalities and Morse–Bott Yang–Mills energy functions, arXiv:1706.09349

  48. Feehan, P.M.N.: Resolution of singularities and geometric proofs of the Łojasiewicz inequalities. Geom. Topol. 23(7), 3273–3313 (2019). arXiv:1708.09775

    MathSciNet  MATH  Google Scholar 

  49. Feehan, P.M.N., Maridakis, M.: Łojasiewicz–Simon gradient inequalities for analytic and Morse–Bott functions on Banach spaces. J. Reine Angew. Math. arXiv:1510.03817v8. (in press)

  50. Feehan, P.M.N., Maridakis, M.: Łojasiewicz–Simon gradient inequalities for analytic and Morse–Bott functions on Banach spaces and applications to harmonic maps. arXiv:1510.03817v5

  51. Feehan, P.M.N., Maridakis, M.: Łojasiewicz–Simon gradient inequalities for harmonic maps. arXiv:1903.01953

  52. Fernández, L.: The dimension and structure of the space of harmonic 2-spheres in the \(m\)-sphere. Ann. of Math. (2) 175(3), 1093–1125 (2012)

    MathSciNet  MATH  Google Scholar 

  53. Goldman, W.M., Millson, J.J.: The deformation theory of representations of fundamental groups of compact Kähler manifolds. Inst. Hautes Études Sci. Publ. Math. 67, 43–96 (1988)

    MATH  Google Scholar 

  54. Gromoll, D., Meyer, W.T.: On differentiable functions with isolated critical points. Topology 8, 361–369 (1969)

    MathSciNet  MATH  Google Scholar 

  55. Guillemin, V., Sternberg, S.: Geometric Asymptotics. American Mathematical Society, Providence (1977)

    MATH  Google Scholar 

  56. Guo, S., Jianhong, W.: Bifurcation Theory Of Functional Differential Equations, Applied Mathematical Sciences, vol. 184. Springer, New York (2013)

    MATH  Google Scholar 

  57. Haraux, A.: Positively homogeneous functions and the Łojasiewicz gradient inequality. Ann. Polon. Math. 87, 165–174 (2005)

    MathSciNet  MATH  Google Scholar 

  58. Haraux, A., Jendoubi, M.A.: The Łojasiewicz gradient inequality in the infinite-dimensional Hilbert space framework. J. Funct. Anal. 260, 2826–2842 (2011)

    MathSciNet  MATH  Google Scholar 

  59. Hélein, F.: Harmonic Maps, Conservation Laws and Moving Frames, 2nd Edition, Cambridge Tracts in Mathematics, vol. 150. Cambridge University Press, Cambridge (2002)

    Google Scholar 

  60. Henry, D.: Geometric Theory of Semilinear Parabolic Equations. Lecture Notes in Mathematics, vol. 840. Springer, New York (1981)

    Google Scholar 

  61. Hironaka, H.: Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II, Ann. Math. (2) 79 (1964), 109–203; ibid. (2) 79 (1964), 205–326

  62. Hofer, H.H.W.: A note on the topological degree at a critical point of mountainpass-type. Proc. Am. Math. Soc. 90(2), 309–315 (1984)

    MathSciNet  MATH  Google Scholar 

  63. Hofer, H.H.W.: The topological degree at a critical point of mountain-pass type, nonlinear functional analysis and its applications, Part 1 (Berkeley, California, 1963), Proceedings of Symposia in Pure Mathematics, vol. 45, American Mathematical Society, Providence, RI 1986, 501–509 (1983)

  64. Hörmander, L.: The Analysis of Linear Partial Differential Operators. III. Pseudo-Differential Operators. Springer, Berlin (2007)

    MATH  Google Scholar 

  65. Horn, R.A., Johnson, C.R.: Topics in Matrix Analysis. Cambridge University Press, Cambridge (1994). (corrected reprint of the 1991 original)

    MATH  Google Scholar 

  66. Huang, S.-Z.: Gradient Inequalities, Mathematical Surveys and Monographs, vol. 126. American Mathematical Society, Providence (2006)

    Google Scholar 

  67. Ilmanen, T.: Elliptic regularization and partial regularity for motion by mean curvature. Mem. Am. Math. Soc. 108(520) (1994)

  68. Jost, J.: Riemannian Geometry and Geometric Analysis, Universitext, 6th edn. Springer, Heidelberg (2011)

    MATH  Google Scholar 

  69. Krantz, S.G., Parks, H.R.: A Primer of Real Analytic Functions, Second Ed., Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks], Birkhäuser Boston, Inc., Boston (2002)

  70. Kuiper, N.H.: \(C^{1}\)-equivalence of functions near isolated critical points. Ann. Math. Stud. 69, 199–218 (1972)

    Google Scholar 

  71. Kuo, H.H.: The Morse–Palais lemma on Banach spaces. Bull. Am. Math. Soc. 80, 363–365 (1974)

    MathSciNet  MATH  Google Scholar 

  72. Kwon, H.: Asymptotic Convergence of Harmonic Map Heat Flow. Stanford University, Palo Alto, CA, Ph.D. thesis (2002)

  73. Lang, S.: Fundamentals of Differential Geometry, Graduate Texts in Mathematics, vol. 191. Springer, New York (1999)

    Google Scholar 

  74. Lang, S.: Introduction to Differentiable Manifolds, Universitext, 2nd edn. Springer, New York (2002)

    Google Scholar 

  75. Lawson Jr., H.B.: Lectures on Minimal Submanifolds. Vol. I, Second Ed., Mathematics Lecture Series, vol. 9. Publish or Perish Inc., Wilmington (1980)

    Google Scholar 

  76. Lemaire, L., Wood, J.C.: Jacobi fields along harmonic 2-spheres in \({\mathbb{C}}{{\rm P}}^{2}\) are integrable. J. Lond. Math. Soc. (2) 66(2), 468–486 (2002)

    MATH  Google Scholar 

  77. Lemaire, L., Wood, J.C.: Jacobi fields along harmonic 2-spheres in 3- and 4-spheres are not all integrable. Tohoku Math. J. (2) 61(2), 165–204 (2009)

    MathSciNet  MATH  Google Scholar 

  78. Leng, Y., Zhao, E., Zhao, H.: Notes on the extension of the mean curvature flow. Pac. J. Math. 269(2), 385–392 (2014)

    MathSciNet  MATH  Google Scholar 

  79. Lindenstrauss, J., Tzafriri, L.: On the complemented subspaces problem. Israel J. Math. 9, 263–269 (1971)

    MathSciNet  MATH  Google Scholar 

  80. Liu, Q., Yang, Y.: Rigidity of the harmonic map heat flow from the sphere to compact Kähler manifolds. Ark. Mat. 48, 121–130 (2010)

    MathSciNet  MATH  Google Scholar 

  81. Łojasiewicz, S.: Ensembles Semi-analytiques. Institut des Hautes Etudes Scientifiques, Bures-sur-Yvette (1965). (LaTeX version by M. Coste, August 29, 2006 based on mimeographed course notes by S. Łojasiewicz)

    MATH  Google Scholar 

  82. Math Stack Exchange, The dual of the direct sum, Internet, October 20 (2012). https://math.stackexchange.com/questions/217596/the-dual-of-the-direct-sum

  83. Mantegazza, C.: Lecture Notes on Mean Curvature Flow, Progress in Mathematics, vol. 290. Birkhäuser/Springer Basel AG, Basel (2011)

    MATH  Google Scholar 

  84. Mawhin, J., Willem, M.: On the generalized Morse lemma. Bull. Soc. Math. Belg. Sér. B 37(2), 23–29 (1985)

    MathSciNet  MATH  Google Scholar 

  85. Michor, P.W., Mumford, D.: Vanishing geodesic distance on spaces of submanifolds and diffeomorphisms. Doc. Math. 10, 217–245 (2005)

    MathSciNet  MATH  Google Scholar 

  86. Milnor, J.W.: Dynamics in One Complex Variable, Third Ed., Annals of Mathematics Studies, vol. 160. Princeton University Press, Princeton (2006)

    Google Scholar 

  87. Moore, J.D.: Introduction to Global Analysis, Graduate Studies in Mathematics, vol. 187. American Mathematical Society, Providence (2017)

    Google Scholar 

  88. Moser, J.K.: On the volume elements on a manifold. Trans. Am. Math. Soc. 120, 286–294 (1965)

    MathSciNet  MATH  Google Scholar 

  89. Nagura, T.: On the Jacobi differential operators associated to minimal isometric immersions of symmetric spaces into spheres. I. Osaka Math. J. 18(1), 115–145 (1981)

    MathSciNet  MATH  Google Scholar 

  90. Nagura, T.: On the Jacobi differential operators associated to minimal isometric immersions of symmetric spaces into spheres. II. Osaka Math. J. 19(1), 79–124 (1982)

    MathSciNet  MATH  Google Scholar 

  91. Nagura, T.: On the Jacobi differential operators associated to minimal isometric immersions of symmetric spaces into spheres III. Osaka Math. J. 19(2), 241–281 (1982)

    MathSciNet  MATH  Google Scholar 

  92. Nicolaescu, L.I.: An Invitation to Morse Theory, Universitext, 2nd edn. Springer, New York (2011)

    MATH  Google Scholar 

  93. Nirenberg, L.: Topics in Nonlinear Functional Analysis, Courant Lecture Notes in Mathematics, vol. 6, New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, Chapter 6 by E. Zehnder, Notes by R. A. Artino, Revised reprint of the 1974 original (2001)

  94. Palais, R.S.: Morse theory on Hilbert manifolds. Topology 2, 299–340 (1963)

    MathSciNet  MATH  Google Scholar 

  95. Palais, R.S.: The Morse lemma for Banach spaces. Bull. Am. Math. Soc. 75, 968–971 (1969)

    MathSciNet  MATH  Google Scholar 

  96. Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential Equations, Applied Mathematical Sciences, vol. 44. Springer, New York (1983)

    MATH  Google Scholar 

  97. Petro, M.: Moduli spaces of Riemann surfaces, Ph.D. thesis, The University of Wisconsin–Madison, Madison, WI, p. 119 (2008)

  98. Pietsch, A.: Operator Ideals, North-Holland Mathematical Library, vol. 20, North-Holland Publishing Co., Amsterdam-New York, Translated from German by the author (1980)

  99. Postnikov, M.M., Rudyak, Y.B.: Morse Lemma, Encyclopedia of Mathematics. Springer Verlag and the European Mathematical Society, http://www.encyclopediaofmath.org/index.php?title=Morse_lemma&oldid=32324

  100. Poston, T., Stewart, I.: Catastrophe Theory and Its Applications, Dover Publications, Inc., Mineola, NY, With an Appendix by D. R. Olsen, S. R. Carter and A. Rockwood, Reprint of the 1978 original (1996)

  101. Ritoré, M., Sinestrari, C.: Mean Curvature Flow and Isoperimetric Inequalities, Advanced Courses in Mathematics. CRM Barcelona, Birkhäuser Verlag, Basel, Edited by Vicente Miquel and Joan Porti (2010)

  102. Råde, J.: On the Yang–Mills heat equation in two and three dimensions. J. Reine Angew. Math. 431, 123–163 (1992)

    MathSciNet  MATH  Google Scholar 

  103. Rudin, W.: Functional Analysis, 2nd edn., International Series in Pure and Applied Mathematics, 2nd edn. McGraw-Hill Inc, New York (1991)

    Google Scholar 

  104. Schoen, R.M.: Topics in differential geometry minimal submanifolds, unpublished lecture notes

  105. Seidel, P.: A long exact sequence for symplectic Floer cohomology. Topology 42(5), 1003–1063 (2003)

    MathSciNet  MATH  Google Scholar 

  106. Sell, G.R., You, Y.: Dynamics of Evolutionary Equations, Applied Mathematical Sciences, vol. 143. Springer, New York (2002)

    MATH  Google Scholar 

  107. Shi, W., Vorotnikov, D.: Uniformly compressing mean curvature flow. J. Geom. Anal. 29(4), 3055–3097 (2019). arXiv:1711.03864

    MathSciNet  MATH  Google Scholar 

  108. Simon, L.: Asymptotics for a class of nonlinear evolution equations, with applications to geometric problems. Ann. Math. (2) 118, 525–571 (1983)

    MathSciNet  MATH  Google Scholar 

  109. Simon, L.: Isolated singularities of extrema of geometric variational problems. Lecture Notes in Mathematics, vol. 1161. Springer, Berlin (1985)

    Google Scholar 

  110. Simon, L.: Theorems on Regularity and Singularity of Energy Minimizing Maps. Lectures in Mathematics ETH Zürich. Birkhäuser, Basel (1996)

    Google Scholar 

  111. Simons, J.: Minimal varieties in riemannian manifolds. Ann. Math.(2) 88, 62–105 (1968)

    MathSciNet  MATH  Google Scholar 

  112. Simpson, C.T.: Higgs bundles and local systems. Inst. Hautes Études Sci. Publ. Math. 75, 5–95 (1992)

    MathSciNet  MATH  Google Scholar 

  113. Smale, S.J.: Morse theory and a non-linear generalization of the Dirichlet problem. Ann. Math. (2) 80, 382–396 (1964)

    MathSciNet  MATH  Google Scholar 

  114. Smith, R.T.: Harmonic mappings of spheres. Am. J. Math. 97, 364–385 (1975)

    MathSciNet  MATH  Google Scholar 

  115. Smith, R.T.: The second variation formula for harmonic mappings. Proc. Am. Math. Soc. 47, 229–236 (1975)

    MathSciNet  MATH  Google Scholar 

  116. Smoczyk, K.: Mean Curvature Flow in Higher Codimension: Introduction and Survey, Global Differential Geometry, Springer Proceedings in Mathematics, vol. 17, pp. 231–274. Springer, Heidelberg (2012)

    MATH  Google Scholar 

  117. Struwe, M.: Variational Methods, 4th edn. Springer, Berlin (2008)

    MATH  Google Scholar 

  118. Tanabe, H.: Equations of Evolution, Monographs and Studies in Mathematics, vol. 6, Pitman (Advanced Publishing Program), Boston, MA (1979). Translated from the Japanese by N. Mugibayashi and H. Haneda. MR 533824 (82g:47032)

  119. Tanabe, H.: Functional Analytic Methods for Partial Differential Equations, Monographs and Textbooks in Pure and Applied Mathematics, vol. 204. Marcel Dekker Inc., New York (1997)

    Google Scholar 

  120. Topping, P.M.: Rigidity in the harmonic map heat flow. J. Differ. Geom. 45, 593–610 (1997)

    MathSciNet  MATH  Google Scholar 

  121. Tromba, A.J.: The Morse lemma on Banach spaces. Proc. Am. Math. Soc. 34, 396–402 (1972)

    MathSciNet  MATH  Google Scholar 

  122. Tuan, V.T., Ang, D.D.: A representation theorem for differentiable functions. Proc. Am. Math. Soc. 75(2), 343–350 (1979)

    MathSciNet  MATH  Google Scholar 

  123. Uhlenbeck, K.K.: Morse theory on Banach manifolds. Bull. Am. Math. Soc. 76, 105–106 (1970)

    MathSciNet  MATH  Google Scholar 

  124. Weinstein, A.D.: Symplectic structures on Banach manifolds. Bull. Am. Math. Soc. 75, 1040–1041 (1969)

    MathSciNet  MATH  Google Scholar 

  125. White, B.: The space of \(m\)-dimensional surfaces that are stationary for a parametric elliptic functional. Indiana Univ. Math. J. 36(3), 567–602 (1987)

    MathSciNet  MATH  Google Scholar 

  126. White, B.: The space of minimal submanifolds for varying Riemannian metrics. Indiana Univ. Math. J. 40(1), 161–200 (1991)

    MathSciNet  MATH  Google Scholar 

  127. White, B.: On the bumpy metrics theorem for minimal submanifolds. Am. J. Math. 139(4), 1149–1155 (2017)

    MathSciNet  MATH  Google Scholar 

  128. Whittlesey, E.F.: Analytic functions in Banach spaces. Proc. Am. Math. Soc. 16, 1077–1083 (1965)

    MathSciNet  MATH  Google Scholar 

  129. Wittmann, J.: The Banach manifold \(C^k(M, N)\). Differ. Geom. Appl. 63, 166–185 (2019)

    MATH  Google Scholar 

  130. Xin, Y.: Minimal Submanifolds and Related Topics, Nankai Tracts in Mathematics, vol. 8. World Scientific Publishing Co. Inc, River Edge (2003)

    Google Scholar 

  131. Yagi, A.: Abstract Parabolic Evolution Equations and Their Applications, Springer Monographs in Mathematics. Springer, Berlin (2010)

    Google Scholar 

  132. Zaal, M.M.: The gradient structure of the mean curvature flow. Adv. Calc. Var. 8(3), 183–202 (2015)

    MathSciNet  MATH  Google Scholar 

  133. Zeidler, E.: Nonlinear Functional Analysis and Its Applications. I. Fixed-Point Theorems. Springer, New York (1986)

    MATH  Google Scholar 

Download references

Acknowledgements

I am indebted to Michael Greenblatt and András Némethi for independently pointing out to me that, for functions on Euclidean space, the Morse Lemma for functions with degenerate critical points (also known as the Morse Lemma with parameters or Splitting Lemma) should be the key ingredient needed to prove the main result of this article in the finite-dimensional case (Corollary 3). I am extremely grateful to Brian White for explaining results of his [125,126,127] and others on minimal surfaces and integrability of Jacobi fields and to Leon Simon for explaining his results and results with Adams on integrability of Jacobi fields in [4, 108, 109]. I also thank Carles Bivià-Ausina, Otis Chodosh, Tristan Collins, Santiago Encinas, Luis Fernandez, Antonella Grassi, David Hurtubise, Johan de Jong, Daniel Ketover, Qingyue Liu, Doug Moore, Yanir Rubinstein, Siddhartha Sahi, Ovidiu Savin, Peter Topping, Graeme Wilkin, and Jarek Włodarczyk for helpful communications, discussions, or questions during the preparation of this article. I am grateful to the National Science Foundation for their support and the Dublin Institute for Advanced Studies and Yi-Jen Lee and the Institute of Mathematical Sciences at the Chinese University of Hong Kong for their hospitality and support. Lastly, I am most grateful to the anonymous referee for numerous comments and suggestions that helped improve this article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul M. N. Feehan.

Additional information

Communicated by A. Chang.

Publisher's Note

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

The author was partially supported by National Science Foundation Grant DMS-1510064 and the Dublin Institute for Advanced Studies.

Appendices

Appendix A: Rate of convergence of a gradient flow for a function obeying a Łojasiewicz gradient inequality

We recall the following enhancement of Huang [66, Theorem 3.4.8].

Theorem A.1

(Convergence rate under the validity of a Łojasiewicz gradient inequality) (See Feehan [44, Theorem 3].) Let \({\mathscr {U}}\) be an open subset of a real Banach space \({\mathscr {X}}\) that is continuously embedded and dense in a Hilbert space \({\mathscr {H}}\). Let \({\mathscr {E}}:{\mathscr {U}}\rightarrow \mathbb {R}\) be an analytic function with gradient map \({\mathscr {E}}':{\mathscr {U}}\rightarrow {\mathscr {H}}\) and \(x_\infty \in {\mathscr {U}}\) be a critical point, that is, \({\mathscr {E}}'(x_\infty )=0\). Assume that there are constants \(c \in (0,\infty )\), and \(\sigma \in (0,1]\), and \(\theta \in [1/2,1)\) such that

$$\begin{aligned} \Vert {\mathscr {E}}'(x)\Vert _{{\mathscr {H}}} \ge c|{\mathscr {E}}(x) - {\mathscr {E}}(x_\infty )|^\theta , \quad \text {for all } x \in {\mathscr {U}}_\sigma , \end{aligned}$$
(A.1)

where \({\mathscr {U}}_\sigma := \{x\in {\mathscr {X}}: \Vert x-x_\infty \Vert _{\mathscr {X}}< \sigma \}\). Let \(u \in C^\infty ([0,\infty ); {\mathscr {X}})\) be a solution to the gradient system

$$\begin{aligned} \dot{u}(t) = -{\mathscr {E}}'(u(t)), \quad t \in (0,\infty ), \end{aligned}$$
(A.2)

and assume that the orbit \(O(u) := \{u(t): t\ge 0\} \subset {\mathscr {X}}\) obeys \(O(u) \subset {\mathscr {U}}_\sigma \). Then there exists \(u_\infty \in {\mathscr {H}}\) such that

$$\begin{aligned} \Vert u(t) - u_\infty \Vert _{\mathscr {H}}\le \Psi (t), \quad t\ge 0, \end{aligned}$$
(A.3)

where

$$\begin{aligned} \Psi (t) := {\left\{ \begin{array}{ll} \displaystyle \frac{1}{c(1-\theta )}\left( c^2(2\theta -1)t + (\gamma -a)^{1-2\theta }\right) ^{-(1-\theta )/(2\theta -1)}, &{} 1/2< \theta < 1, \\ \displaystyle \frac{2}{c}\sqrt{\gamma -a}\exp (-c^2t/2), &{}\theta = 1/2, \end{array}\right. }\nonumber \\ \end{aligned}$$
(A.4)

and \(a, \gamma \) are constants such that \(\gamma > a\) and

$$\begin{aligned} a \le {\mathscr {E}}(v) \le \gamma , \quad \text {for all } v \in {\mathscr {U}}. \end{aligned}$$

If in addition u obeys Hypothesis A.2, then \(u_\infty \in {\mathscr {X}}\) and

$$\begin{aligned} \Vert u(t+1) - u_\infty \Vert _{\mathscr {X}}\le 2C_1\Psi (t), \quad t\ge 0, \end{aligned}$$
(A.5)

where \(C_1 \in [1,\infty )\) is the constant in Hypothesis A.2 for \(\delta =1\).

We recall the

Hypothesis A.2

(A priori interior estimate for a trajectory) (See Feehan [44, Hypothesis 2.1].) Let \({\mathscr {X}}\) be a Banach space that is continuously embedded in a Hilbert space \({\mathscr {H}}\). If \(\delta \in (0,\infty )\) is a constant, then there is a constant \(C_1 = C_1(\delta ) \in [1,\infty )\) with the following significance. If \(S, T \in \mathbb {R}\) are constants obeying \(S+\delta \le T\) and \(u \in C^\infty ([S,T); {\mathscr {X}})\), we say that \(\dot{u} \in C^\infty ([S,T); {\mathscr {X}})\) obeys an a priori interior estimate on (0, T] if

$$\begin{aligned} \int _{S+\delta }^T \Vert \dot{u}(t)\Vert _{\mathscr {X}}\,dt \le C_1\int _S^T \Vert \dot{u}(t)\Vert _{\mathscr {H}}\,dt. \end{aligned}$$
(4.6)

In applications, \(u \in C^\infty ([S,T); {\mathscr {X}})\) in Hypothesis A.2 will often be a solution to a quasi-linear parabolic partial differential system, from which an a priori estimate (4.6) may be deduced. For example, Hypothesis A.2 is verified by Feehan [44, Lemma 17.12] for a nonlinear evolution equation on a Banach space \({\mathcal {V}}\) of the form (see Caps [26], Henry [60], Pazy [96], Sell and You [106], Tanabe [118, 119] or Yagi [131])

$$\begin{aligned} \frac{du}{dt} + {\mathcal {A}}u = {\mathcal {F}}(t,u(t)), \quad t\ge 0, \quad u(0) = u_0, \end{aligned}$$
(4.7)

where \({\mathcal {A}}\) is a positive, sectorial, unbounded operator on a Banach space \({\mathcal {W}}\) with domain \({\mathcal {V}}^2 \subset {\mathcal {W}}\) and the nonlinearity \({\mathcal {F}}\) has suitable properties.

Results on the rate of convergence of a gradient flow defined by a function obeying a Łojasiewicz gradient inequality in specific examples have been proved earlier—see Simon [108] and Adams and Simon [4] for a restricted class of analytic energy functions arising in geometric analysis and Råde [102, Proposition 7.4] for the Yang-Mills energy function on connections on principal bundles over a closed smooth manifold of dimension two or three. For a recent example, see Carlotto, Chodosh, and Rubinstein [27, Theorem 1] for the Yamabe function on Riemannian metrics over closed smooth manifolds of dimension greater than or equal to three.

Appendix B: Morse–Bott functions and quadratic simple normal crossing functions

We are often asked about the relationship between Morse–Bott functions and quadratic simple normal crossing functions as in (1.5), so we explain the relationship in this section for \(\mathbb {K}=\mathbb {R}\); the analogous discussion applies for \(\mathbb {K}=\mathbb {C}\).

For an integer \(p \ge 1\) and writing \(\mathbb {R}^* = \mathbb {R}{\setminus }\{0\}\), we let \(\mathbb {R}^{p-1} = (\mathbb {R}^p{\setminus }\{0\})/\mathbb {R}^* = S^{p-1}/\{\pm 1\}\) denote real projective space, so \(\mathbb {R}\mathbb {P}^0 \cong \{1\}\) and \(\mathbb {R}\mathbb {P}^1 \cong S^1\) while \(\mathbb {R}\mathbb {P}^{p-1}\) with \(p\ge 3\) is obtained by identifying antipodal points of the sphere \(S^{p-1}\).

Definition B.1

(Blowup at a point and exceptional divisor) (See Krantz and Parks [69, Definition 6.2.2].) Let \(p\ge 1\) be an integer and W be an open neighborhood of the origin in \(\mathbb {R}^p\). The blowup of W at the origin is the set

$$\begin{aligned} \widetilde{W} := \{(x,\ell ) \in W \times \mathbb {R}^{p-1} : x \in \ell \}, \end{aligned}$$

where \(\pi : \widetilde{W} \ni (x,\ell ) \mapsto x \in W\) is the blowup map and \(E := \pi ^{-1}(0) \subset \widetilde{W}\) is the exceptional divisor.

The set \(\widetilde{W}\) is a real analytic manifold and the quotient map \(\pi : \widetilde{W} \rightarrow W\) is real analytic and restricts to a real analytic diffeomorphism \(\pi : \widetilde{W} {\setminus }E \cong W {\setminus }\{0\}\). By viewing \(\mathbb {R}^{p-1} = S^{p-1}/\{\pm 1\}\) and \(\mathbb {R}^p {\setminus }\{0\} = \mathbb {R}_+\times S^{p-1}\), we may also write

$$\begin{aligned} \widetilde{W}&= \{(x,\ell ) \in W \times \mathbb {R}^{p-1} : x \in \ell \} \\&= \{(x,[u]) \in W \times S^{p-1}/\{\pm 1\}: x \in \mathbb {R}u\} \\&= \{(x,[u]) \in W \times S^{p-1}/\{\pm 1\}: x = \pm |x| u\} \\&= \{(x,u) \in W \times S^{p-1}: (x,u) \sim (y,v) \text { if } |x|=|y| \text { and } u = \pm v \} \\&= \{(s,u) \in \mathbb {R}\times S^{p-1}: su \in W \text { and } (s,u) \sim (t,v) \text { if } (t,v) = \pm (s,u) \}, \end{aligned}$$

where \([u] = \{\pm u\}\) and, in the last line, the blowup map is \(\pi : \widetilde{W} \ni [s,u] \mapsto su \in W\) and

$$\begin{aligned} \pi ^{-1}(0) = \{\{0\} \times S^{p-1}\}/\{\pm 1\} \cong S^{p-1}/\{\pm 1\} = \mathbb {R}^{p-1} \end{aligned}$$

is the exceptional divisor.

If W is an open neighborhood of the origin in \(\mathbb {R}^p\) or the half-space \(\mathbb {H}^p = \{x\in \mathbb {R}^p:x_p\ge 0\}\), then we could alternatively define the blowup of W at the origin to be the real analytic manifold with boundary,

$$\begin{aligned} \widehat{W} := \{(r,u) \in [0,\infty )\times S^{p-1}: ru \in W\}, \end{aligned}$$

following the usual definition of polar coordinates on \(\mathbb {R}^p{\setminus }\{0\}\). The map \(\pi : \widehat{W} \ni (r,u) \mapsto ru \in W\) is the blowup map and \(\pi ^{-1}(0) = \{0\}\times S^{p-1} \cong S^{p-1}\) is now the exceptional divisor.

Suppose now that \(U\subset \mathbb {R}^d\) is an open neighborhood of the origin and \(f:U \rightarrow \mathbb {R}\) is a \(C^2\) function with \(f(0)=0\) and \(f'(0)=0\) and that is Morse–Bott at the origin in the sense of Definition 1.5 (1). Thus, after possibly shrinking U, we have that \({{\,\mathrm{Crit}\,}}f\) is a \(C^2\) submanifold of U of dimension \(c = \dim {\text {Ker}}f''(0)\). Moreover, we may further assume that U is connected and so \({{\,\mathrm{Crit}\,}}f \subset f^{-1}(0)\).

Theorem 2.10 and Remark 2.12 (the Morse–Bott Lemma) imply, after possibly shrinking U, that one can find an neighborhood V of the origin in \(\mathbb {R}^d\) and a \(C^2\) diffeomorphismFootnote 16\(\Phi :\mathbb {R}^d \supset V \ni y\mapsto x\in U \subset \mathbb {R}^d\) such that \(\Phi (0)=0\) and

$$\begin{aligned} f\circ \Phi (y) = \sum _{i=1}^{p} y_i^2 - \sum _{i=p+1}^{p+n} y_i^2, \quad \text {for all } y \in V. \end{aligned}$$

Note that \(p+n = d-c\) and

$$\begin{aligned} {{\,\mathrm{Crit}\,}}f\circ \Phi = V\cap \bigcap _{i=1}^{d-c}\{y_i=0\}. \end{aligned}$$

If \(n=0\) and thus \(1 \le p = d-c\), we may write \((y_1,\ldots ,y_p) = su\), for \(s \in [0,\infty )\) and \(u \in S^{p-1} \subset \mathbb {R}^p\), so that

$$\begin{aligned} f\circ \varpi (s,u,y_{p+1},\ldots ,y_d) = s^2, \quad \text {for all } (su,y_{p+1},\ldots ,y_d) \in U, \end{aligned}$$

where we define

$$\begin{aligned} \varpi (s,u,y_{p+1},\ldots ,y_d) := \Phi (su,y_{p+1},\ldots ,y_d). \end{aligned}$$

We see that \(\varpi \) gives a \(C^2\) map from an open neighborhood V of the origin in \([0,\infty )\times S^{p-1}\times \mathbb {R}^{d-p}\) onto \(U\subset \mathbb {R}^d\) such that \(\varpi (0)=0\) and

$$\begin{aligned} \varpi (\{s=0\}\cap V) = U\cap \bigcap _{i=1}^p\{y_i=0\} = U\cap \bigcap _{i=1}^{d-c}\{y_i=0\}, \end{aligned}$$

and \(\varpi \) is a diffeomorphism from \(V{\setminus }\{s=0\}\) onto its image.

Similarly, if \(p=0\) and thus \(1 \le n = d-c\), we may write \((y_1,\ldots ,y_n) = tv\), for \(t \in [0,\infty )\) and \(v \in S^{n-1} \subset \mathbb {R}^n\), so that

$$\begin{aligned} f\circ \varpi (t,v,y_{n+1},\ldots ,y_d) = -t^2, \quad \text {for all } (tv,y_{n+1},\ldots ,y_d) \in U, \end{aligned}$$

where we define \(\varpi (t,v,y_{n+1},\ldots ,y_d) = \Phi (tv,y_{n+1},\ldots ,y_d)\). We see that \(\varpi \) gives a \(C^2\) map from an open neighborhood of the origin in \([0,\infty )\times S^{p-1}\times \mathbb {R}^{d-p}\) into \(\mathbb {R}^d\) such that \(\varpi (0)=0\) and

$$\begin{aligned} \varpi (\{t=0\}\cap V) = U\cap \bigcap _{i=1}^n\{y_i=0\} = U\cap \bigcap _{i=1}^{d-c}\{y_i=0\}, \end{aligned}$$

and \(\varpi \) is a diffeomorphism from \(V{\setminus }\{t=0\}\) onto its image.

Finally, if \(n \ge 1\) and \(p \ge 1\), we may write \((y_1,\ldots ,y_p) = su\) and \((y_{p+1},\ldots ,y_{p+n}) = tv\), for \(s, t \in [0,\infty )\) and \(u \in S^{p-1}\) and \(v \in S^{n-1}\), so that

$$\begin{aligned} f\circ \varpi (s,t,u,v,y_{p+n+1},\ldots ,y_d) = s^2-t^2, \quad \text {for all } (su,tv,y_{p+n+1},\ldots ,y_d) \in U, \end{aligned}$$

where we define

$$\begin{aligned} \varpi (s,t,u,v,y_{p+n+1},\ldots ,y_d) := \Phi (su,tv,y_{p+n+1},\ldots ,y_d). \end{aligned}$$

We see that \(\varpi \) gives a \(C^2\) map from an open neighborhood of the origin in

$$\begin{aligned}{}[0,\infty )\times [0,\infty )\times S^{p-1}\times S^{n-1}\times \mathbb {R}^{d-n-p} \end{aligned}$$

into \(\mathbb {R}^d\) such that \(\varpi (0)=0\) and

$$\begin{aligned} \varpi (\{s=0\}\cap \{t=0\}\cap W) = V\cap \bigcap _{i=1}^{n+p}\{y_i=0\} = V\cap \bigcap _{i=1}^{d-c}\{y_i=0\}, \end{aligned}$$

after possibly shrinking V and \(\varpi \) is a diffeomorphism from \(W{\setminus }(\{s=0\}\cup \{t=0\})\) onto its image.

Define a diffeomorphism of \(\mathbb {R}^2\) by \((t_1,t_2) \mapsto (s,t) = \varphi (t_1,t_2)\) where \(t_1=s+t\) and \(t_2=s-t\), so that \(s = \frac{1}{2}(t_1+t_2)\) and \(t = \frac{1}{2}(t_1-t_2)\). Hence, we obtain

$$\begin{aligned}&f\circ \Pi (t_1,t_2,u,v,y_{p+n+1},\ldots ,y_d) = t_1t_2, \quad \\&\quad \text {for all } (\varphi _1(t_1,t_2)u,\varphi _2(t_1,t_2)v,y_{p+n+1},\ldots ,y_d) \in U, \end{aligned}$$

where we define

$$\begin{aligned} \Pi (t_1,t_2,u,v,y_{p+n+1},\ldots ,y_d) := \Phi (\varphi _1(t_1,t_2)u,\varphi _2(t_1,t_2)v,y_{p+n+1},\ldots ,y_d). \end{aligned}$$

We see that \(\Pi \) gives a \(C^2\) map from an open neighborhood of the origin in

$$\begin{aligned} \{(t_1,t_2)\in [0,\infty )\times \mathbb {R}: |t_2|\le t_1\} \times S^{p-1}\times S^{n-1}\times \mathbb {R}^{d-n-p} \end{aligned}$$

into \(\mathbb {R}^d\) such that \(\Pi (0)=0\) and

$$\begin{aligned} \Pi (\{t_1=0\}\cap \{t_2=0\}\cap V) = U\cap \bigcap _{i=1}^{d-c}\{y_i=0\}, \end{aligned}$$

and \(\Pi \) is a diffeomorphism from \(V{\setminus }(\{t_1=0\}\cup \{t_2=0\})\) onto its image.

In the preceding discussion we could have replaced the roles of the blowups \([0,\infty )\times S^{p-1}\) or \([0,\infty )\times S^{n-1}\) by \((\mathbb {R}\times S^{p-1})/\{\pm 1\}\) or \((\mathbb {R}\times S^{n-1})/\{\pm 1\}\) and the roles of the exceptional divisors, \(S^{p-1}\) or \(S^{n-1}\) by \(\mathbb {R}\mathbb {P}^{p-1}\) or \(\mathbb {R}\mathbb {P}^{n-1}\), the only difference being an increase in notational complexity. In summary, we have proved the

Proposition B.2

(Pull-back of a Morse–Bott function to a quadratic simple normal crossing function) Let \(d \ge 2\) be an integer, \(U\subset \mathbb {R}^d\) be an open neighborhood of the origin, and \(f:U \rightarrow \mathbb {R}\) be a \(C^2\) function that is Morse–Bott at the origin and obeys \(f(0)=0\). Then, after possibly shrinking U, there are an open neighborhood V of the origin in \(\mathbb {R}^d\) and a \(C^2\) map \(\pi :V \rightarrow U\) such that \(\pi \) restricts to a diffeomorphism from \(V{\setminus }\{y_1=0\}\) or \(V{\setminus }(\{y_1=0\}\cup \{y_2=0\})\) onto its image and

$$\begin{aligned} \pi ^*f(y) = \pm y_1^2 \quad \text {or}\quad y_1y_2, \quad \text {for all } y=(y_1,\ldots ,y_d) \in V, \end{aligned}$$

and \(\pi ({{\,\mathrm{Crit}\,}}f\circ \pi ) = {{\,\mathrm{Crit}\,}}f\), where \({{\,\mathrm{Crit}\,}}f\circ \pi = \{y_1=0\}\cap V\) or \((\{y_1=y_2=0\})\cap V\).

Appendix C: Integrability and Morse–Bott conditions for the harmonic map energy and the area functions

In Sect. 1.5 we defined the concepts of Jacobi vector, integrable Jacobi vector, and integrable critical point (see Definition 1.17). We noted (see Lemma 1.18) that if a function is Morse–Bott at a critical point, then that critical point is integrable. Theorem 8 has been proved by Simon for a specific class of analytic functions on certain Banach spaces (given by \(C^{2,\alpha }\) sections of a Riemannian vector bundle over a closed Riemannian manifold) that includes the harmonic map energy and the area functions. We shall give a proof of a more general version of Theorem 8 elsewhere [45], but we outline here how Theorem 8 may be proved; in addition to the references cited below, we also refer the reader to Simon [110, Sects. 3.11–3.14 and 3.13.16] for further expository details.

Proof

(Outline of proof of Theorem 8) In order to avoid notational conflict with the remainder of this section, we let \({\mathscr {E}}\) denote the analytic function considered in Theorem 8. As in Simon’s proof of his infinite-dimensional version [108, Theorem 3] of the Łojasiewicz gradient inequality, one first applies Lyapunov–Schmidt reduction as in [108, Sect. 2] to the function \({\mathscr {E}}:{\mathscr {U}}\rightarrow \mathbb {K}\). This step yields an analytic embedding \(\Psi :{\mathscr {V}}\cap {\mathscr {K}}\rightarrow {\mathscr {X}}\) of the intersection with the kernel \({\mathscr {K}}:={\text {Ker}}{\mathscr {E}}''(x_0)\) and an open neighborhood \({\mathscr {V}}\) of the origin in \(\tilde{{\mathscr {X}}}\), together with an analytic function \(\Gamma = {\mathscr {E}}\circ \Psi :{\mathscr {V}}\cap {\mathscr {K}}\rightarrow \mathbb {K}\) (see Adams and Simon [4, p. 230], Feehan and Maridakis [49, Lemmas 2.3 and 2.5], Simon [108, pp. 538–539], or Simon [109, Part II, Sect. 6]). By hypothesis, \(x_0\) is an integrable critical point in the sense of Definition 1.17 and so, after possibly shrinking \({\mathscr {V}}\), the function \(\Gamma \) is constant on \({\mathscr {V}}\cap {\mathscr {K}}\) by Adams and Simon [4, Lemma 1, p. 231].

One can now show that \({\mathscr {E}}'(x)=0\) for all \(x \in \Psi ({\mathscr {V}}\cap {\mathscr {K}})\), essentially by reversing our proof of [49, Lemma 2.5] or arguing as in Simon [108, p. 539], and thus

$$\begin{aligned} \Psi ({\mathscr {V}}\cap {\mathscr {K}}) \subseteq {{\,\mathrm{Crit}\,}}{\mathscr {E}}. \end{aligned}$$

By hypothesis, \({\mathscr {E}}''(x_0) \in {\mathscr {L}}({\mathscr {X}},\tilde{{\mathscr {X}}})\) is Fredholm with index zero and thus we have \({\mathscr {X}}={\mathscr {X}}_0\oplus {\mathscr {K}}\) and \(\tilde{{\mathscr {X}}}\cong {\mathscr {X}}_0\oplus {\mathscr {K}}\) (see Lemma 2.4 (2)). In particular,

$$\begin{aligned} {\text {Ran}}{\mathscr {E}}''(x_0) + {\mathscr {K}}= \tilde{{\mathscr {X}}} \end{aligned}$$

and so, after possibly shrinking \({\mathscr {U}}\), the analytic gradient map \({\mathscr {E}}':{\mathscr {U}}\rightarrow \tilde{{\mathscr {X}}}\) is transverse to the (linear) submanifold \({\mathscr {K}}\subset \tilde{{\mathscr {X}}}\) and hence the preimage \(({\mathscr {E}}')^{-1}({\mathscr {K}})\) is an open analytic submanifold of \({\mathscr {X}}\) by the Preimage Theorem from differential topology [73, Proposition II.2.4]. We thus have inclusions

$$\begin{aligned} \Psi ({\mathscr {V}}\cap {\mathscr {K}}) \subseteq {{\,\mathrm{Crit}\,}}{\mathscr {E}}\subseteq ({\mathscr {E}}')^{-1}({\mathscr {K}}), \end{aligned}$$

noting that \({{\,\mathrm{Crit}\,}}{\mathscr {E}}\equiv ({\mathscr {E}}')^{-1}(0)\). Furthermore, we have

$$\begin{aligned} T_{x_0}\Psi ({\mathscr {V}}\cap {\mathscr {K}}) = {\mathscr {K}}= T_{x_0}({\mathscr {E}}')^{-1}({\mathscr {K}}), \end{aligned}$$

where the first equality follows from the construction of \(\Psi \) (see [49, Lemma 2.3]) and the second from the observations below:

$$\begin{aligned} T_{x_0}({\mathscr {E}}')^{-1}({\mathscr {K}})&= ({\mathscr {E}}''(x_0))^{-1}(T_{{\mathscr {E}}'(x_0)}{\mathscr {K}}) = ({\mathscr {E}}''(x_0))^{-1}({\mathscr {K}}) \\&= ({\mathscr {E}}''(x_0))^{-1}(0) \quad \text {(since } \tilde{{\mathscr {X}}}={\text {Ran}}{\mathscr {E}}''(x_0)\oplus {\mathscr {K}}) \\&= {\mathscr {K}}\quad \text {(by definition)}. \end{aligned}$$

Hence, after possibly shrinking \({\mathscr {U}}\) or \({\mathscr {V}}\), we have \(\Psi ({\mathscr {V}}\cap {\mathscr {K}}) = ({\mathscr {E}}')^{-1}({\mathscr {K}})\) and consequently

$$\begin{aligned} \Psi ({\mathscr {V}}\cap {\mathscr {K}}) = {{\,\mathrm{Crit}\,}}{\mathscr {E}}= ({\mathscr {E}}')^{-1}({\mathscr {K}}). \end{aligned}$$

In particular, \({{\,\mathrm{Crit}\,}}{\mathscr {E}}\) is an open analytic submanifold of \({\mathscr {X}}\) with tangent space \(T_{x_0}{{\,\mathrm{Crit}\,}}{\mathscr {E}}= {\text {Ker}}{\mathscr {E}}''(x_0)\) and so \({\mathscr {E}}\) is Morse–Bott at \(x_0\) in the sense of Definition 1.9 (1). \(\square \)

1.1 Appendix C.1: Integrability and Morse–Bott conditions for the harmonic map energy function

Following Lemaire and Wood [76, Sect. 1], we review the concept of integrability of a Jacobi field along a harmonic map and describe the relation between integrability and the Morse–Bott condition for the harmonic map energy function at a harmonic map. We then list a few examples where integrability is known for harmonic maps.Footnote 17

We begin by recalling the second variation of the energy for the harmonic energy function \({\mathscr {E}}\) discussed in Sect. 1.7.2. For a smooth two-parameter variation \(f_{t,s}:M\rightarrow N\) of a map \(f:M\rightarrow N\) with \(f_{0,0} = f\) and \(\partial f_{t,s}/\partial t|_{(0,0)} = v\) and \(\partial ^2 f_{t,s}/\partial s|_{(0,0)} = w\), the Hessian of \({\mathscr {E}}\) at f is defined by

$$\begin{aligned} {\mathscr {E}}''(f)(v,w) := \left. \frac{\partial ^2{\mathscr {E}}(f_{t,s})}{\partial t\partial s}\right| _{(0,0)}, \end{aligned}$$

where \({\mathscr {E}}\) is as in (1.19). One has

$$\begin{aligned} {\mathscr {E}}''(f)(v,w) = (J_f(v),w)_{L^2(M,g)}, \end{aligned}$$

where

$$\begin{aligned} J_f(v) := \Delta v - {\text {tr}}R^N(df,v)df \end{aligned}$$

is called the Jacobi operator, a self-adjoint linear elliptic differential operator. Here, \(\Delta \) denotes the Laplacian induced on \(f^{-1}TN\) and the sign conventions for \(\Delta \) and the curvature \(R^N\) are those of Eells and Lemaire [41].

Let v be a vector field along f, that is, a smooth section of \(f^{-1}TN\), where \(f:M\rightarrow N\) is a smooth map. Then v is called a Jacobi field (for the energy) if \(J_f(v) = 0\). The space of Jacobi fields \({\text {Ker}}J_f\) is finite-dimensional and its dimension is called the (\({\mathscr {E}}\))-nullity of f.

Definition C.1

(Integrability of a Jacobi field along a harmonic map) (See Lemaire and Wood [76, Definition 1.2].) A Jacobi field v along a harmonic map \(f_0:M\rightarrow N\) is said to be integrable if there is a smooth family of harmonic maps, \(f_t:M\rightarrow N\) for \(t\in (-\varepsilon ,\varepsilon )\), such that \(f_t|_{t=0} = f_0\) and \(v = \partial f_t/\partial t|_{t=0}\).

The following result is stated by Kwon in her Ph.D. thesis [72] (directed by Simon); it can be deduced from Theorem 8 by applying, for example, methods of Feehan and Maridakis [51].

Theorem C.2

(Integrability of Jacobi fields and manifolds of harmonic maps) (See Kwon [72, Proposition 4.1].) Let \(d\ge 2\) be an integer and \(\alpha \in (0,1)\) be a constant. Let (Mg) and (Nh) be closed, smooth Riemannian manifolds, with M of dimension d, and assume that there is a smooth isometric embedding \(N\subset \mathbb {R}^n\) for some integer n. If \(f_0 \in C^\infty (M;N)\) is a harmonic map, so \({\mathscr {E}}'(f_0)=0\), then the following hold:

  1. (1)

    If there is a constant \(\delta =\delta (f_0,g,h,n,\alpha ) \in (0,1]\) such that

    $$\begin{aligned} U_{f_0,\delta } := \left\{ f \in C^{2,\alpha }(M;N): \Vert f-f_0\Vert _{C^{2,\alpha }(M;\mathbb {R}^n)} < \delta \text { and } {\mathscr {E}}'(f)=0\right\} \end{aligned}$$
    (C.1)

    is an open smooth manifold with tangent space \(T_{f_0}U_{f_0,\delta } = {\text {Ker}}{\mathscr {E}}''(f_0)\) at \(f_0\), then every Jacobi vector field in \({\text {Ker}}{\mathscr {E}}''(f_0)\) is integrable.

  2. (2)

    If (Nh) is real analytic, the isometric embedding \(N\subset \mathbb {R}^n\) is real analytic, and every Jacobi vector field in \({\text {Ker}}{\mathscr {E}}''(f_0)\) is integrable, then there is a constant \(\delta =\delta (f_0,g,h,n,\alpha ) \in (0,1]\) such that the set \(U_{f_0,\delta }\) in (C.1) is an open smooth manifold with tangent space \(T_{f_0}U_{f_0,\delta } = {\text {Ker}}{\mathscr {E}}''(f_0)\) at \(f_0\).

It follows that for real-analytic target manifolds, all Jacobi fields along all harmonic maps are integrable if and only if the space of harmonic maps is a manifold whose tangent bundle is given by the Jacobi fields [76, p. 470]. By Definition 1.5, the conclusion of Theorem C.2 (2) is equivalent to the assertion that all Jacobi fields along \(f_0\) are integrable if and only if the harmonic map energy function \({\mathscr {E}}\) is Morse–Bott at \(f_0\).

For a further discussion of integrability and additional references, see Adams and Simon [4, Sect. 1], Kwon [72, Sect. 4.1], and Simon [109, pp. 270–272].

According to [76, Theorem 1.3], any Jacobi field along a harmonic map from \(S^2\) to \(\mathbb {C}\mathbb {P}^2\) is integrable, where the two-sphere \(S^2\) has its unique conformal structure and the complex projective space \(\mathbb {C}\mathbb {P}^2\) has its standard Fubini-Study metric of holomorphic sectional curvature 1; see Crawford [34] for additional results.

From the list of examples provided by Lemaire and Wood [76, p. 471], there are few other examples of families of harmonic maps that are guaranteed to be integrable, with the list including harmonic maps from \(S^2\) to \(S^2\) but excluding harmonic maps from \(S^2\) to \(S^3\) or \(S^4\) [77].

Fernández [52] has proved that the space \({\mathrm {Harm}}_d(S^2,S^{2n})\) of degree-d harmonic maps from \(S^2\) into \(S^{2n}\) has dimension \(2d+n^2\). However, thus far, integrability for such maps is known only when \(n=1\). Bolton and Fernandez [18] provide a nice survey of what is known regarding regularity of \({\mathrm {Harm}}_d(S^2,S^{2n})\): they recall that \({\mathrm {Harm}}_d(S^2,S^2)\) is known to be a smooth manifold, outline a proof that \({\mathrm {Harm}}_d(S^2,S^6)\) is also a smooth manifold, and survey results on the structure of \({\mathrm {Harm}}_d(S^2,S^4)\) and why that space is not a smooth manifold.

1.2 Appendix C.2: Integrability and Morse–Bott conditions for the area function

Suppose that \(m,n \ge 1\) and \(r\ge 2\) are integers and M is a closed, connected, oriented, smooth manifold of dimension m. We let \(C^{r,\alpha }(M;\mathbb {R}^n)\) denote the Banach space of \(C^{r,\alpha }\) maps from M into \(\mathbb {R}^n\), where \(\alpha \in [0,1]\), and let \({{\,\mathrm{Imm}\,}}^{r,\alpha }(M;\mathbb {R}^n) \subset C^{r,\alpha }(M;\mathbb {R}^n)\) denote the open subset of \(C^{r,\alpha }\) immersions, and let \({{\,\mathrm{Emb}\,}}^{r,\alpha }(M;\mathbb {R}^n) \subset C^{r,\alpha }(M;\mathbb {R}^n)\) denote the open subset of \(C^{r,\alpha }\) embeddings. If \(\Phi \in C^{r,\alpha }(M;\mathbb {R}^d)\) is an embedding, then \(g_\Phi := \Phi ^*g\) is a Riemannian metric on M while if \(\Phi \) is an immersion, then \(g_\Phi \) may be singular. We now consider the area or volume function,

$$\begin{aligned} {{\,\mathrm{Imm}\,}}^{r,\alpha }(M;\mathbb {R}^n) \ni \Phi \mapsto {\mathscr {E}}(\Phi ) := {\text {Vol}}(M, g_\Phi ) \in [0,\infty ). \end{aligned}$$

Then \(\Phi (M)\) is called a critical immersed submanifold or (as customary) a minimal immersed submanifold if \({\mathscr {E}}'(\Phi )=0\), where

$$\begin{aligned} {\mathscr {E}}'(\Phi )\eta = \left. \frac{d}{dt}{\text {Vol}}(M, g_{\Phi +t\Phi _\eta })\right| _{t=0} \end{aligned}$$

for all vector fields \(\eta \in C^{r,\alpha }(TM)\). One can show that

$$\begin{aligned} {\mathscr {E}}'(\Phi )\eta = \left( \eta ,{\mathscr {E}}'(\Phi )\right) _{L^2(M)}, \end{aligned}$$

with an explicit expression for the gradient \({\mathscr {E}}'(\Phi )\) provided by the first variation formula—see Calegari [24, Proposition 2.1], Colding and Minicozzi [30, pp. 154–155], Dajczer and Tojeiro [35, Proposition 3.1], Lawson [75], Schoen [104, Sect. 2.1], or Xin [130, Theorem 1.2.2 and Remark 1.2.5].

An explicit expression for the Hessian \({\mathscr {E}}''(\Phi )\) at a critical point \(\Phi \) is provided by the second variation formula—see Calegari [24, Proposition 3.1], Colding and Minicozzi [30, pp. 154–155], Lawson [75], Schoen [104, Sect. 2.1], and Xin [130, Theorem 6.1.1].

More generally, if we replace \(\mathbb {R}^n\) in the preceding discussion by a connected, smooth manifold N without boundary, then it is known that \(C^r(M;N)\) is a smooth Banach manifold—see Abraham [1], Bruveris [23], Eichhorn [42], Eliasson [43], or Wittmann [129]. (It is highly likely that published proofs of this result extend to show that \(C^{r,\alpha }(M;N)\) is a Banach manifold when \(\alpha \in [0,1]\) and, furthermore, that \(W^{k,p}(M;N)\) is Banach manifold for \(k\in \mathbb {N}\) and \(p\in [1,\infty )\), at least for \(k\ge 2\) and \(kp>m\), taking note of the Sobolev Embedding Theorem [5, Theorem 4.12].) We refer to Michor and Mumford [85, Sect. 2.1] for their analysis of these spaces in the \(C^\infty \) category.

Recall from Dajczer and Tojeiro [35, Corollary 3.7] or Xin [130, Corollary 1.3.4] that there exists no minimal isometric immersion \(\Phi : M^m \rightarrow \mathbb {R}^n\) of a compact Riemannian manifold without boundary. Hence, we restrict our attention to cases where M and N are closed or M and N are complete or M is compact with boundary and N is complete.

One could again derive an analogue of Theorem C.2 (giving the relationship between integrability of Jacobi vector fields and the Morse–Bott property of an immersed minimal submanifold) from Theorem 8 or derive an analogue of Theorem C.2 for immersed minimal submanifolds from prior, more general results of Simon [108, 109] and Adams and Simon [4].

Adams and Simon list examples of minimal submanifolds whose Jacobi vector fields are all integrable as well as examples that have nontrivial Jacobi vector fields that are not integrable [4, pp. 249–252]. See also Allard and Almgren [6, Sect. 6], Nagura [89,90,91], Simons [111], and Smith [114, 115] (via Remark C.3) for related examples.

White [126, 127] has shown that for generic \(C^r\) Riemannian metrics on a manifold N, there are no closed, immersed, minimal submanifolds \(M\subset N\) with nontrivial Jacobi fields; the case of geodesics, including immersed geodesics, was proved earlier by Abraham [2].

Remark C.3

(On the relationship between harmonic maps and minimal surfaces) It is useful to recall the relationship between harmonic maps f from a closed, smooth Riemann surface \((\Sigma ,g)\) into a closed Riemannian manifold (Nh), and immersed minimal surfaces in (Nh), since that relationship enriches our supply of examples. Chern and Goldberg [28, Proposition 5.1] show that if \(\Sigma =S^2\) and f is a harmonic immersion, then f is a minimal immersion. More generally, though they assume \((N,h)=\mathbb {R}^3\) with its standard metric and allow \((\Sigma ,g)\) to be a Riemann surface with boundary, Dierkes, Hildebrandt, and Sauvigny prove [38, Theorem 2.6.1] that a conformal map f is minimal if and only if it is harmonic. According to their [38, Definition 2.6.1], they may replace \(\mathbb {R}^3\) by \(\mathbb {R}^n\) for any \(n\ge 2\) and more generally, by any Riemannian manifold (Nh) of dimension \(n\ge 2\). Moore [87, Theorem 4.2.2] proves a similar result, namely that (the image of) a (weakly) conformal harmonic map \(f:(\Sigma ,g) \rightarrow (N,h)\) is a minimal surface. By restricting to \(\Sigma =S^2\), Moore [87, Proposition 4.2.3] recovers the result of Chern and Goldberg: a harmonic two-sphere, \(f:(S^2,g_{\mathrm {round}}) \rightarrow (N,h)\), is automatically weakly conformal and hence a parametrized minimal surface. See also [38, pp. 36, 77, 249–250, and 309–311] and their discussion of Lichtenstein’s Theorem on reparameterizing maps of the disk and [38, pp. 249–250] for the relationship between area and energy integrals and the minimization problem.

Remark C.4

(On the interpretation of mean curvature flow as a gradient flow) While there is a wealth of references on mean curvature flow, relatively few treat it as gradient flow for the area (volume) function, thus making it less accessible to gradient flow methods pioneered by Simon [108, 109]. For interpretations of mean curvature flow as a gradient system, we refer the reader to Bellettini [14, Remark 2.8 and Sect. 2.3], Colding, Minicozzi, and Pedersen [33, Sect. 1], Ilmanen [67], Mantegazza [83, p. 7, second paragraph], Ritoré and Sinestrari [101, Equation (4.3)], Smoczyk [116], and Zaal [132]. Shi and Vorotnikov [107] provide a useful recent reference, with a view to applications. For introductions to mean curvature flow, we refer to Ecker [40], Mantegazza [83], Ritoré and Sinestrari [101].

For applications of the DeTurck trick [37] to convert mean curvature flow to a nonlinear parabolic partial differential equation and establish short-time existence, we refer to Andrews and Baker [7], Baker [12], and Leng, Zhao, and Zhao [78].

As in the case of Ricci flow, the interpretation of mean curvature flow as a gradient system can lead to the introduction of a time-varying family of Hilbert spaces—a family of \(L^2\) spaces defined by a measure that depends on the time-varying family of immersions [83, Sect. 1.2, page 7].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Feehan, P.M.N. On the Morse–Bott property of analytic functions on Banach spaces with Łojasiewicz exponent one half. Calc. Var. 59, 87 (2020). https://doi.org/10.1007/s00526-020-01734-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-020-01734-4

Mathematics Subject Classification

Navigation