Skip to main content
Log in

Algebraic Structure of Classical Field Theory: Kinematics and Linearized Dynamics for Real Scalar Fields

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We describe the elements of a novel structural approach to classical field theory, inspired by recent developments in perturbative algebraic quantum field theory. This approach is local and focuses mainly on the observables over field configurations, given by certain spaces of functionals which are studied here in depth. The analysis of such functionals is characterized by a combination of geometric, analytic and algebraic elements which (1) make our approach closer to quantum field theory, (2) allow for a rigorous analytic refinement of many computational formulae from the functional formulation of classical field theory and (3) provide a new pathway towards understanding dynamics. Particular attention will be paid to aspects related to nonlinear hyperbolic partial differential equations and their linearizations.

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. A causal (resp. timelike, null) curve \(\gamma :I\rightarrow {\mathscr {M}}\) is said to be inextendible if there is no causal (resp. timelike, null) curve \({\tilde{\gamma }}:{\tilde{I}}\rightarrow {\mathscr {M}}\) such that \({\tilde{I}}\supsetneqq I\) and \({\tilde{\gamma }} \vert _{I}=\gamma \).

  2. Some physics texts, such as [30], call \({\mathscr {C}}^\infty ({\mathscr {M}})\) the space of field histories on \({\mathscr {M}}\).

  3. In the context of field theory, such connections were formally introduced in [30]. They allow one to extend to higher orders the notion of fiber derivative employed in the calculus of variations [11]. For a precise, general concept of ultralocal lifts of connections on target spaces, see for instance Example 4.5.3, pp. 94 of [45].

  4. The restriction on \({\mathscr {U}}\) can be weakened in a certain sense. See Lemma 3.3.

  5. For simplicity, here we allow ourselves a slight abuse of notation—strictly speaking, the smooth density supported in \(\mathrm {supp}\ F\) representing \(F^{(1)}[\varphi ]\) for each \(\varphi \in {\mathscr {U}}\) is only defined up to an exact d-form, so when we write \(*_{\!g}F^{(1)}\) we apply \(*_{\!g}\) simultaneously to all representatives of \(F^{(1)}[\varphi ]\) for each \(\varphi \in {\mathscr {U}}\). In other words, we are dealing with all d-forms representing \(F^{(1)}[\varphi ]\) simultaneously. We shall be more precise with this from the proof of Proposition 2.2 onwards.

  6. This property is assumed a priori in Definition 6.1 of [15].

  7. Here “continuous” means sequentially continuous with respect to the (weak) Hörmander topology on the extended domain (see Sect. 4.1 below), as shown e.g. by Theorem 8.2.13, pp. 268–269 of [50] and, more precisely, by Theorems 8.2.9. (iii) and 8.2.10, pp. 515–520 of [23]. One can see indirectly from the arguments in [13] that one cannot hope to upgrade this result to full continuity, unless one uses instead the strong Hörmander topology (see also Remark 4.3 below).

  8. Recall that a completely regular topological space X is said to be compactly generated or a k-space if the topology of X coincides with the final topology induced by the inclusions of compact subsets of X. This is equivalent to the space of continuous real-valued functions on X being complete with respect to the topology of uniform convergence on compact subsets of X (see e.g. Theorem 3.6.4, pp. 70 of [54]).

  9. Nonetheless, in this case the \(c^\infty \)-topology coincides with the so-called Kelleyfication of \({\mathscr {F}}\), which is the final topology induced by all compact subsets of \({\mathscr {F}}\) through their respective inclusions (see e.g. Theorem 4.11 (3), pp. 39–40 of [63]). It is clear that the Kelleyfication of \({\mathscr {F}}\) coinciding with the original topology of \({\mathscr {F}}\) amounts to \({\mathscr {F}}\) being compactly generated (see footnote 8 above). This happens if e.g. \({\mathscr {F}}\) is metrizable.

  10. MB differentiability and MB smoothness are respectively listed in Keller’s treatise [55] as “\({\mathscr {C}}^k_c\)- and \({\mathscr {C}}^\infty _c\)-differentiability”. Here we avoid his nomenclature, for it clashes with the usual notation for differentiable and smooth functions with compact support.

  11. However, as argued e.g. in Proposition 2.7, pp. 17 of [63], if \(\gamma \) is Lipschitz (i.e. the subset \(\{(t-s)^{-1}(\gamma (t)-\gamma (s))\ |\ t\ne s,\,a\le t,s\le b\}\) is bounded) then it suffices to assume that \({\mathscr {F}}\) is convenient to get the Riemann integral of \(\gamma \) along [ab] with all the properties discussed in this Appendix.

References

  1. Abraham, R., Marsden, J.E.: Foundations of Mechanics, 2nd edn. Addison-Wesley, Boston (1978)

    MATH  Google Scholar 

  2. Adams, R.A., Fournier, J.J.F.: Sobolev Spaces, 2nd edn. Elsevier, New York (2002)

    MATH  Google Scholar 

  3. Anderson, I.M.: The Variational Bicomplex. Technical Report, Utah State University (1989)

  4. Bär, C., Ginoux, N., Pfäffle, F.: Wave Equations on Lorentzian Manifolds and Quantization. European Mathematical Society, Warsaw (2007)

    Book  MATH  Google Scholar 

  5. Bastiani, A.: Applications différentiables et varietés différentiables de dimension infinie. J. Anal. Math. 13, 1–114 (1964)

    Article  MathSciNet  MATH  Google Scholar 

  6. Benavides Navarro, J.J., Minguzzi, E.: Global hyperbolicity is stable in the interval topology. J. Math. Phys. 52, 112504 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Bernal, A.N., Sánchez, M.: On smooth Cauchy hypersurfaces and Geroch’s splitting theorem. Commun. Math. Phys. 243, 461–470 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Bernal, A.N., Sánchez, M.: Smoothness of time functions and the metric splitting of globally hyperbolic spacetimes. Commun. Math. Phys. 257, 43–50 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Bernal, A.N., Sánchez, M.: Further results on the smoothability of Cauchy hypersurfaces and Cauchy time functions. Lett. Math. Phys. 77, 183–197 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Bernal, A.N., Sánchez, M.: Globally hyperbolic spacetimes can be defined as "causal" instead of "strongly causal". Class. Quantum Grav. 24, 745–749 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Binz, E., Śniatycki, J., Fischer, H.: Geometry of Classical Fields. Elsevier, Amsterdam (1988)

    MATH  Google Scholar 

  12. Brennecke, F., Dütsch, M.: Removal of violations of the master Ward identity in perturbative QFT. Rev. Math. Phys. 20, 119–172 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  13. Brouder, C., Dang, N.V., Hélein, F.: Boundedness and continuity of the fundamental operations on distributions having a specified wave front set (with a counterexample by Semyon Alesker). Studia Math. 232, 201–226 (2016)

    MathSciNet  MATH  Google Scholar 

  14. Brouder, C., Dang, N.V., Laurent-Gengoux, C., Rejzner, K.: Properties of field functionals and characterization of local functionals. J. Math. Phys. 59, 023508 (2018)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Brunetti, R., Dütsch, M., Fredenhagen, K.: Perturbative algebraic quantum field theory and renormalization groups. Adv. Theor. Math. Phys. 13, 1541–1599 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  16. Brunetti, R., Fredenhagen, K.: Microlocal analysis and interacting quantum field theories: renormalization on physical backgrounds. Commun. Math. Phys. 208, 623–661 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Brunetti, R., Fredenhagen, K.: Quantum field theory on curved backgrounds. In: Bär, C., Fredenhagen, K. (eds.) Quantum Field Theory on Curved Spacetimes: Concepts and Mathematical Foundations. Lecture Notes in Physics, vol. 786, pp. 129–155. Springer, Berlin (2009)

    Chapter  Google Scholar 

  18. Brunetti, R., Fredenhagen, K., Köhler, M.: The microlocal spectrum condition and Wick polynomials of free fields on curved spacetimes. Commun. Math. Phys. 180, 633–652 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Brunetti, R., Fredenhagen, K., Rejzner, K.: Quantum gravity from the point of view of locally covariant quantum field theory. Commun. Math. Phys. 345, 741–779 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Brunetti, R., Fredenhagen, K., Ribeiro, P.L.: (in preparation)

  21. Bryant, R.L., Griffiths, P.A., Yang, D.: Characteristics and existence of isometric embeddings. Duke Math. J. 50, 893–994 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  22. Cariñena, J.F., Crampin, M., Ibort, L.A.: On the multisymplectic formalism for first order field theories. Differ. Geom. Appl. 1, 345–374 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  23. Chazarain, J., Piriou, A.: Introduction to the Theory of Linear Partial Differential Equations. North-Holland Publishing Company, Amsterdam (1982)

    MATH  Google Scholar 

  24. Christodoulou, D.: The Action Principle and Partial Differential Equations. Princeton University Press, Princeton (2000)

    MATH  Google Scholar 

  25. Crnković, C., Witten, E.: Covariant description of canonical formalism in geometrical theories. In: Hawking, S.W., Israel, W. (eds.) Three Hundred Years of Gravitation, pp. 676–684. Cambridge University Press, Cambridge (1987)

    Google Scholar 

  26. Dabrowski, Y.: Functional properties of generalized Hörmander spaces of distributions I: duality theory, completions and bornologifications. arXiv:1411.3012 [math-ph]

  27. Dabrowski, Y.: Functional properties of generalized Hörmander spaces of distributions II: multilinear maps and applications to spaces of functionals with wave front set conditions. arXiv:1412.1749 [math-ph]

  28. Dabrowski, Y., Brouder, C.: Functional properties of Hörmander’s space of distributions having a specified wavefront set. Commun. Math. Phys. 332, 1345–1380 (2014)

    Article  ADS  MATH  Google Scholar 

  29. de Donder, T.: Théorie Invariante du Calcul des Variations. Gauthier-Villars, Mexico (1935)

    MATH  Google Scholar 

  30. DeWitt, B.S.: The spacetime approach to quantum field theory. In: DeWitt, B.S., Stora, R. (eds.) Les Houches Session XL, Relativity, Groups and Topology II, pp. 382–738. Elsevier, Amsterdam (1983)

    Google Scholar 

  31. Dencker, N.: On the propagation of polarization sets for systems of real principal type. J. Funct. Anal. 46, 351–372 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  32. Dimock, J.: Algebras of local observables on a manifold. Commun. Math. Phys. 77, 219–228 (1980)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. Duistermaat, J.J.: Fourier Integral Operators. Birkhäuser, Basel (1996)

    MATH  Google Scholar 

  34. Dütsch, M., Fredenhagen, K.: The master Ward identity and generalized Schwinger–Dyson equation in classical field theory. Commun. Math. Phys. 243, 275–314 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Dütsch, M., Fredenhagen, K.: Causal perturbation theory in terms of retarded products, and a proof of the action Ward identity. Rev. Math. Phys. 16, 1291–1348 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  36. Engelking, R.: General Topology. Revised and Completed edn. Heldermann Verlag, Berlin (1989)

    MATH  Google Scholar 

  37. Forger, M., Romero, S.V.: Covariant Poisson brackets in geometric field theory. Commun. Math. Phys. 256, 375–410 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Fredenhagen, K., Rejzner, K.: Batalin-Vilkovisky formalism in the functional approach to classical field theory. Commun. Math. Phys. 314, 93–127 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Frölicher, A.: Smooth structures. In: Kamps, K.H., Pumplün, D., Tholen, W. (eds.) Category Theory—Applications to Algebra, Logic and Topology. Lecture Notes in Mathematics 962, pp. 69–81. Springer, Berlin (1982)

  40. Geroch, R.: Domain of dependence. J. Math. Phys. 11, 437–449 (1970)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. Glöckner, H.: Discontinuous non-linear mappings on locally convex direct limits. Publ. Math. Debr. 68, 1–13 (2006)

    MathSciNet  MATH  Google Scholar 

  42. Gotay, M.J.: A multisymplectic framework for classical field theory and the calculus of variations. In: Francaviglia, M. (ed.) Mechanics, Analysis and Geometry: 200 Years after Lagrange, pp. 203–235. Elsevier, Amsterdam (1991)

    Chapter  Google Scholar 

  43. Grafakos, L.: Classical Fourier Analysis, 2nd edn. Springer, Berlin (2008)

    MATH  Google Scholar 

  44. Haag, R.: Local Quantum Physics—Fields, Particles, Algebras, 2nd edn. Springer, Berlin (1996)

    MATH  Google Scholar 

  45. Hamilton, R.S.: The inverse function theorem of Nash and Moser. Bull. Am. Math. Soc. (N.S.) 7, 65–222 (1982)

  46. Hawking, S.W., Ellis, G.F.R.: The Large Scale Structure of Space-Time. Cambridge University Press, Cambridge (1973)

    Book  MATH  Google Scholar 

  47. Héléin, F.: Multisymplectic formalism and the covariant phase space. In: Bielawski, R., Houston, K., Speight, M. (eds.) Variational Problems in Differential Geometry, pp. 94–126. Cambridge University Press, Cambridge (2012)

    Google Scholar 

  48. Héléin, F.: First integrals for nonlinear dispersive equations. Trans. Am. Math. Soc. 368, 6939–6978 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  49. Hintz, P., Vasy, A.: Global analysis of quasilinear wave equations on asymptotically Kerr-de Sitter spaces. Int. Math. Res. Not. 2016, 5355–5426 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  50. Hörmander, L.: The Analysis of Linear Partial Differential Operators I—Distribution Theory and Fourier Analysis, 2nd edn. Springer, Berlin (1990)

    MATH  Google Scholar 

  51. Hörmander, L.: The Analysis of Linear Partial Differential Operators III—Pseudodifferential Operators, 2nd edn. Springer, Berlin (1994)

    Google Scholar 

  52. Hörmander, L.: Lectures on Nonlinear Hyperbolic Differential Operators. Springer, Berlin (1997)

    MATH  Google Scholar 

  53. Jakobs, S.: Eichbrücken in der klassichen Feldtheorie. Diplomarbeit, Universität Hamburg. http://www-library.desy.de/preparch/desy/thesis/desy-thesis-09-009.pdf (2009). Accessed 25 Apr 2019

  54. Jarchow, H.: Locally Convex Spaces. B. G. Teubner, Stuttgart (1981)

    Book  MATH  Google Scholar 

  55. Keller, H.H.: Differential Calculus in Locally Convex Spaces. Lecture Notes in Mathematics, vol. 417. Springer, Berlin (1974)

  56. Keller, K.J.: Dimensional regularization in position space and a forest formula for regularized Epstein–Glaser renormalization. Ph.D. Thesis, Universität Hamburg (2010). arXiv:1006.2148 [math-ph]

  57. Kijowski, J.: A finite-dimensional canonical formalism in the classical field theory. Commun. Math. Phys. 30, 99–128 (1973)

    Article  ADS  MathSciNet  Google Scholar 

  58. Klainerman, S.: Global existence for nonlinear wave equations. Commun. Pure Appl. Math. 33, 43–101 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  59. Klainerman, S.: Long-time behavior of solutions to nonlinear evolution equations. Arch. Ration. Mech. Anal. 78, 73–98 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  60. Kolář, I.: A geometrical version of the higher order Hamilton formalism in fibered manifolds. J. Geom. Phys. 1, 127–137 (1984)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  61. Kolář, I., Michor, P.W., Slovák, J.: Natural Operations in Differential Geometry. Springer, Berlin (1993)

    Book  MATH  Google Scholar 

  62. Krasil’shchik, I.S., Lychagin, V.V., Vinogradov, A.M.: Geometry of Jet Spaces and Nonlinear Partial Differential Equations. Gordon and Breach, Washington (1986)

    MATH  Google Scholar 

  63. Kriegl, A., Michor, P.W.: The Convenient Setting of Global Analysis. American Mathematical Society, Providence (1997)

    Book  MATH  Google Scholar 

  64. Lerner, D.E.: The space of Lorentz metrics. Commun. Math. Phys. 32, 19–38 (1973)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  65. Leyland, P., Roberts, J.E.: The cohomology of nets over Minkowski space. Commun. Math. Phys. 62, 173–189 (1978)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  66. Liess, O.: Conical Refractions and Higher Microlocalization. Lecture Notes in Mathematics, vol. 1555. Springer, Berlin (1993)

  67. Majda, A.: Compressible Fluid Flow and Systems of Conservation Laws in Several Space Variables. Springer, Berlin (1984)

    Book  MATH  Google Scholar 

  68. Marolf, D.: The generalized Peierls bracket. Ann. Phys. (N.Y.) 236, 392–412 (1994)

  69. Meise, R.: Nicht-Nuklearität von Räumen beliebig oft differenzierbarer Funktionen. Arch. Math. 34, 143–148 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  70. Michal, A.D.: Differential calculus in linear topological spaces. Proc. Nat. Acad. Sci. USA 24, 340–342 (1938)

    Article  ADS  MATH  Google Scholar 

  71. Milnor, J.: Remarks on infinite-dimensional Lie groups. In: DeWitt, B., Stora, R. (eds.) Les Houches Session XL, Relativity, Groups and Topology II, pp. 1007–1057. Elsevier, Amsterdam (1984)

    Google Scholar 

  72. Moerdijk, I., Reyes, G.E.: Models for Smooth Infinitesimal Analysis. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  73. Müller, O., Sánchez, M.: Lorentzian manifolds isometrically embeddable in \({\mathbb{L}}^N\). Trans. Am. Math. Soc. 363, 5367–5379 (2011)

    Article  MATH  Google Scholar 

  74. Peetre, J.: Une charactérisation abstraite des opérateurs différentiels. Math. Scand. 7, 211–218 (1959). Erratum: ibid. 8, 116–120 (1960)

  75. Peierls, R.E.: The commutation laws of relativistic field theory. Proc. R. Soc. Lond. A214, 143–157 (1952)

    ADS  MathSciNet  MATH  Google Scholar 

  76. Pietsch, A.: Nuclear Locally Convex Spaces. Springer, Berlin (1972)

    Book  MATH  Google Scholar 

  77. Rao, M.M.: Local functionals. In: Kolzow, D. (ed.) Measure Theory, Oberwolfach 1979. Lecture Notes in Mathematics, vol. 794, pp. 484–496. Springer, Berlin (1980)

    Chapter  Google Scholar 

  78. Rejzner, K.: Fermionic fields in the functional approach to classical field theory. Rev. Math. Phys. 23, 1009–1033 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  79. Seiler, W.M.: Involution: The Formal Theory of Differential Equations and its Applications in Computer Algebra. Springer, Berlin (2010)

    Book  MATH  Google Scholar 

  80. Slovák, J.: Peetre theorem for nonlinear operators. Ann. Global Anal. Geom. 6, 273–283 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  81. Sogge, C.D.: Lectures on Non-Linear Wave Equations, 2nd edn. International Press, Boston (2008)

    MATH  Google Scholar 

  82. Stiefel, E.: Richtungsfelder and Fernparallelismus in Mannigfaltigkeiten. Commun. Math. Helv. 8, 3–51 (1936)

    MATH  Google Scholar 

  83. Tso, K.: Nonlinear symmetric positive systems. Ann. Inst. H. Poincaré Anal. Non Linéaire 9, 339–366 (1992)

  84. Vaisman, I.: Lectures on the Geometry of Poisson Manifolds. Birkhäuser, Boston (1994)

    Book  MATH  Google Scholar 

  85. Vinogradov, A.M.: On the algebro-geometric foundations of Lagrangian field theory. Dokl. Akad. Nauk SSSR 236, 284–287 (1977). English translation in Sov. Math. Dokl. 18, 1200–1204 (1977)

  86. Vinogradov, A.M.: A spectral sequence associated with a nonlinear differential equation, and algebro-geometric foundations of Lagrangian field theory with constraints. Dokl. Akad. Nauk SSSR 238, 1028–1031 (1978). English translation in Sov. Math. Dokl. 19, 144–148 (1978)

  87. Wald, R.M.: General Relativity. Chicago University Press, Chicago (1984)

    Book  MATH  Google Scholar 

  88. Wald, R.M.: On identically closed forms locally constructed from a field. J. Math. Phys. 31, 2378–2384 (1990)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  89. Weise, J.-C.: On the algebraic formulation of classical general relativity. Diplomarbeit, Universität Hamburg. http://www.desy.de/uni-th/theses/Dipl_Weise.pdf (2011). Accessed 25 Apr 2019

  90. Weyl, H.: Geodesic fields in the calculus of variations for multiple integrals. Ann. Math. 36, 607–629 (1935)

    Article  MathSciNet  MATH  Google Scholar 

  91. Zajtz, A.: Nonlinear Peetre-like theorems. Univ. Iagel. Acta Math. 37, 351–361 (1999)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank Prof. Frank Michael Forger for a critical reading of an early version of the Introduction, as well as for invaluable advice on the mathematical literature on classical field theory and general presentation details. We are specially grateful to him for discussions on physically relevant functionals, which led to most of the examples presented in Sect. 2.3. We would also like to thank Prof. Stefan Waldmann for pointing out a mistake in the proof of Corollary 4.1 in a previous version of the present paper, and also for his clarifying comments. Finally, we are much grateful to Prof. Christian Brouder for his several inquires about our work, particularly for pointing out a substantial gap in the previous proofs of of Proposition 3.4 and Corollary 3.3, and for numerous discussions, as well as to Prof. Peter Michor for the enlightening observations on MathOverflow, which led us to the crucial Lemma 2.6, and the anonymous referees for the valuable comments. The junior author (P.L.R.) would like to thank the hospitality of the II. Institut für theoretische Physik, Universität Hamburg, the Dipartimento di Matematica, Facoltà di Scienze della Università di Trento, and the Instituto de Matemática e Estatística, University of São Paulo, where most of the work presented in this paper was developed, and also the support from the Research Training Group 1670—“Mathematics Inspired by String Theory and Quantum Field Theory”, Universität Hamburg as well as from the Centro Italiano di Ricerca Matematica (CIRM) and the Bruno Kessler Foundation in the final stages of the writing.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pedro Lauridsen Ribeiro.

Additional information

Communicated by Y. Kawahigashi

Publisher's Note

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

A Short Review of Differential Calculus on Locally Convex Topological Vector Spaces

A Short Review of Differential Calculus on Locally Convex Topological Vector Spaces

In this Appendix we list the basic definitions and results of differential calculus we need. Our basic references are [45] and [63], to whom we refer for more details and proofs. The first reference works only with Fréchet spaces, but the proofs of the results quoted below work in the general case with little or no change.

The notion of differentiability of curves in locally convex topological vector spaces is straightforward.

Definition A.1

Let \(\gamma :(a,b)\rightarrow {\mathscr {F}}\), \(a<b\in {\mathbb {R}}\cup \{\pm \infty \}\) be a continuous curve into a locally convex topological vector space \({\mathscr {F}}\). We say that \(\gamma \) is a \({\mathscr {C}}^1\)curve if for all \(t\in (a,b)\) the limit

$$\begin{aligned} \gamma '(t)\doteq \lim _{s\rightarrow 0}\frac{1}{s}(\gamma (t+s)-\gamma (t)) \end{aligned}$$

exists and defines a continuous curve \(\gamma ':(a,b)\rightarrow {\mathscr {F}}\) (continuity of \(\gamma \) actually follows from these conditions alone, hence it does not hurt to assume it from the start). We also say that \(\gamma \) is a \({\mathscr {C}}^m\) curve, \(m\ge 1\), if \(\gamma ^{(k)}\doteq (\gamma ^{(k-1)})'\) exists and is continuous for all \(1\le k\le m\), where \(\gamma ^{(0)}\doteq \gamma \). If \(\gamma \) is a \({\mathscr {C}}^m\) curve for all m, we say that \(\gamma \) is a smooth curve.

We stress that there would be no loss of generality if we required the domain of smooth curves to be the whole real line: by the chain rule (A.3), \(\gamma :(a,b)\rightarrow {\mathscr {F}}\) is smooth if and only if \(\gamma \circ f:{\mathbb {R}}\rightarrow {\mathscr {F}}\) is smooth for any diffeomorphism \(f:{\mathbb {R}}\rightarrow (a,b)\) (e.g. \(f(\lambda )=\frac{b+a}{2} +\frac{b-a}{2}\tanh (\lambda )\)). Once this is said, let us see how Definition A.1 is realized in the concrete cases that interest us.

  • \({\mathscr {F}}={\mathscr {C}}^\infty ({\mathscr {M}})\) (endowed with the compact-open topology): \(\gamma :{\mathbb {R}}\rightarrow {\mathscr {F}}\) is smooth if and only if \(\gamma (\lambda )(p)={\varPhi }(\lambda ,p)\) for all \((\lambda ,p)\in {\mathbb {R}}\times {\mathscr {M}}\), where \({\varPhi }\in {\mathscr {C}}^\infty ({\mathbb {R}}\times {\mathscr {M}})\);

  • \({\mathscr {F}}={\mathscr {C}}^\infty _c({\mathscr {M}})\) (endowed with the usual inductive limit topology): \(\gamma :{\mathbb {R}}\rightarrow {\mathscr {F}}\) is smooth if and only if \(\gamma (\lambda )(p)={\varPhi }(\lambda ,p)\) for all \((\lambda ,p)\in {\mathbb {R}}\times {\mathscr {M}}\), where \({\varPhi }\in {\mathscr {C}}^\infty ({\mathbb {R}}\times {\mathscr {M}})\) is such that for any \(a<b\in {\mathbb {R}}\) there is a compact subset \(K\subset {\mathscr {M}}\) such that \({\varPhi }(\lambda ,p)={\varPhi }(a,p)\) for all \(p\not \in K\), \(\lambda \in [a,b]\).

The notion of smooth curves allows one to introduce another topology on \({\mathscr {F}}\), given by the final topology induced by \({\mathbb {R}}\) through all smooth curves \(\gamma : {\mathbb {R}}\rightarrow {\mathscr {F}}\). We call this topology the \(c^\infty \)-topology on \({\mathscr {F}}\). This topology is necessarily finer than the original one, but it is not in general a vector space topology—the finest locally convex vector space topology on \({\mathscr {F}}\) that is coarser then the \(c^\infty \)-topology is the bornologification of \({\mathscr {F}}\)’s original topology. The \(c^\infty \)- and the original locally convex vector space topologies coincide if \({\mathscr {F}}\) is e.g. metrizable (such as \({\mathscr {C}}^\infty ({\mathscr {M}})\)), but are distinct for \({\mathscr {F}}={\mathscr {C}}^\infty _c({\mathscr {M}})\) if \({\mathscr {M}}\) is non-compact since then the \(c^\infty \)-topology is not a vector space topology (see e.g. Proposition 4.26 (ii), pp. 45 of [63]).Footnote 9

Given two locally convex vector spaces \({\mathscr {F}}_1\), \({\mathscr {F}}_2\), \({\mathscr {U}}\subset {\mathscr {F}}_1\)\(c^\infty \)-open, we say that a map \({\varPhi }:{\mathscr {U}}\rightarrow {\mathscr {F}}_2\) is conveniently smooth if \({\varPhi }\circ \gamma \) is a smooth curve on \({\mathscr {F}}_2\) for every smooth curve \(\gamma : {\mathbb {R}}\rightarrow {\mathscr {U}}\). We stress that conveniently smooth maps need not even be continuous (see [41] for a counterexample). A simple non-trivial example of a conveniently smooth map \({\varPhi }:{\mathscr {F}}\rightarrow {\mathscr {F}}\) is, of course, the translation \(\varphi \mapsto {\varPhi }(\varphi )=\varphi +\varphi _0\) by a fixed element \(\varphi _0\in {\mathscr {F}}\). In particular, the coordinate change maps \(\kappa _{\varphi _2}\circ \kappa _{\varphi _1}^{-1}:{\mathscr {C}}^\infty _c({\mathscr {M}}) \rightarrow {\mathscr {C}}^\infty _c({\mathscr {M}})\) in the affine flat manifold \({\mathscr {C}}^\infty ({\mathscr {M}})\) (endowed with the Whitney topology) are conveniently smooth for all \(\varphi _1,\varphi _2\in {\mathscr {C}}^\infty ({\mathscr {M}})\) such that \(\varphi _1-\varphi _2\in {\mathscr {C}}^\infty _c({\mathscr {M}})\). This shows that the atlas \({\mathfrak {U}}\) defined in (2.7) induces a smooth structure on \({\mathscr {C}}^\infty ({\mathscr {M}})\); the corresponding smooth manifold topology is, of course, the manifold topology generated by the \(c^\infty \)-open subsets of the modelling vector space \({\mathscr {C}}^\infty _c({\mathscr {M}})\), which is even finer than the Whitney topology. The connected components of this topology are, however, also of the form \({\mathscr {C}}^\infty _c({\mathscr {M}})+\varphi _0\), \(\varphi _0\in {\mathscr {C}}^\infty ({\mathscr {M}})\); therefore, the smooth curves in \({\mathscr {C}}^\infty ({\mathscr {M}})\) with respect to the smooth structure induced by the atlas \({\mathfrak {U}}\) must be of the form \({\mathbb {R}}\ni \lambda \mapsto \gamma (\lambda )=\varphi _0+\gamma _0(\lambda )\), where \(\gamma _0:{\mathbb {R}}\rightarrow {\mathscr {C}}^\infty _c({\mathscr {M}})\) is smooth. Hence, it is just fair to say that such \(\gamma \) is a smooth curve with respect to the Whitney topology, and the smooth structure induced by the atlas \({\mathfrak {U}}\), the smooth structure on\({\mathscr {C}}^\infty ({\mathscr {M}})\)induced by the Whitney topology.

Remark A.1

It can be shown [63] that, for \({\mathscr {C}}^\infty ({\mathscr {M}})\) endowed with the smooth structure induced by the Whitney topology, the bundles

$$\begin{aligned} T^{r,s}{\mathscr {C}}^\infty ({\mathscr {M}})=\left( \otimes ^sT^*{\mathscr {C}}^\infty ({\mathscr {M}})\right) \otimes \left( \otimes ^rT{\mathscr {C}}^\infty ({\mathscr {M}})\right) \end{aligned}$$

of tensors of contravariant rank r and covariant rank s are given at each \(\varphi \in {\mathscr {C}}^\infty ({\mathscr {M}})\) by the space of bounded linear mappings from \(\otimes ^s_\beta {\mathscr {C}}^\infty _c({\mathscr {M}})\) to \(\otimes ^r_\beta {\mathscr {C}}^\infty _c({\mathscr {M}})\). Here \(\otimes _\beta \) denotes the bornological tensor product, whose topology is the finest locally convex topology on the algebraic tensor product such that the canonical quotient map is bounded; this topology is finer than the projective tensor product topology. Nonetheless, \(T{\mathscr {C}}^\infty ({\mathscr {M}})\) and \(T^*{\mathscr {C}}^\infty ({\mathscr {M}})\) do assume the form given in Sect. 2.2 (see the proof of Theorem 42.17, pp. 447–448 of [63]). It also turns out that the particular structure of \({\mathscr {C}}^\infty _c({\mathscr {M}})\), together with Theorems 6.14, pp. 72–73 and 28.7, pp. 280–281 of [63], imply that every kinematical tangent vector on \({\mathscr {C}}^\infty ({\mathscr {M}})\) is also an operational one, i.e. it defines a point derivation on (conveniently) smooth maps \(F:{\mathscr {C}}^\infty ({\mathscr {M}})\rightarrow {\mathbb {R}}\).

In principle, we could develop essentially all tools of differential calculus by using convenient smoothness. However, for the purposes of this paper, it is often preferrable to use a stronger concept of smoothness. Such a notion is provided, for instance, by Michal [70] and Bastiani [5]. This is also the notion employed in the accounts of infinite dimensional differential calculus done by Milnor [71] and Hamilton [45], and all the basic results of Calculus we present in the remainder of this Appendix are formulated in this context (see, however, Remark A.2 below). The basic definition is as follows (See also Definition 2.3 for the special case of real-valued maps):

Definition A.2

Let \({\mathscr {F}}_1,{\mathscr {F}}_2\) be locally convex topological vector spaces, \({\mathscr {U}}\subset {\mathscr {F}}_1\) open, and \(F:{\mathscr {U}}\rightarrow {\mathscr {F}}_2\) a continuous map. We say that F is (MB-)differentiable of order\(m(>0)\) (“MB” stands for the names of Michal and Bastiani) if for all \(k=1, \ldots ,m\) the k-th order directional (Gâteaux) derivatives

$$\begin{aligned} F^{(k)}[\varphi ](\mathbf {\varphi }_1,\ldots ,\mathbf {\varphi }_k)\doteq \frac{\partial ^k}{\partial \lambda _1 \cdots \partial \lambda _k}\left. \phantom {\frac{}{}}\!\right| _{\lambda _1=\cdots =\lambda _k=0}F \left( \varphi +\sum ^k_{j=1}\lambda _j \mathbf {\varphi }_j\right) \end{aligned}$$
(A.1)

exist as jointly continuous maps from \({\mathscr {U}}\times {\mathscr {F}}^k_1\ni (\varphi ,\mathbf {\varphi }_1,\ldots , \mathbf {\varphi }_k)\) to \({\mathscr {F}}_2\). If F is differentiable of order m for all \(m\in {\mathbb {N}}\), we say that F is (MB-)smooth.Footnote 10

The right-hand side of formula (A.1) should be understood as the differentiation of a k-parameter curve taking values in \({\mathscr {F}}_2\), for fixed \(\varphi ,\mathbf {\varphi }_1,\ldots , \mathbf {\varphi }_k\). The argument of F inside the limit is guaranteed to lie inside \({\mathscr {U}}\) for sufficiently small \(\lambda _1,\ldots ,\lambda _k\).

It follows from Definition A.2 that if \(F:{\mathscr {U}}\subset {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\) is MB-differentiable of order \(m>0\) then the maps \({\mathscr {U}}\ni \varphi \mapsto F^{(k)}[\varphi ]\in {\mathscr {L}}^k({\mathscr {F}}_1,{\mathscr {F}}_2)\) are continuous for all \(1\le k\le m\), where \({\mathscr {L}}^k({\mathscr {F}}_1,{\mathscr {F}}_2)\) is the locally convex topological vector space of all k-linear maps from \({\mathscr {F}}_1^k\) to \({\mathscr {F}}_2\) endowed with the compact-open topology. If \({\mathscr {F}}_1\) is semi-Montel (i.e. closed and bounded subsets of \({\mathscr {F}}_1\) are compact), such topology amounts to uniform convergence in bounded subsets of \({\mathscr {F}}_1^k\). If \({\mathscr {F}}_1^k\) is compactly generated (e.g. when \({\mathscr {F}}_1\) is metrizable, see e.g. Proposition 3.3.20, pp. 152 of [36] and footnote 9 above) and \({\mathscr {F}}_2\) is complete, then by Proposition 16.6.2, pp. 361 of [54] \({\mathscr {L}}^k({\mathscr {F}}_1,{\mathscr {F}}_2)\) is also complete.

Given \({\mathscr {U}}\) an arbitrary (i.e. not necessarily open) subset of \({\mathscr {F}}_1\), we say that a continuous map \(F:{\mathscr {U}}\rightarrow {\mathscr {F}}_2\) is differentiable of order m (resp. smooth) if there is \({\mathscr {V}}\supset {\mathscr {U}}\) open in the compact-open topology and a functional \({\tilde{F}}:{\mathscr {V}}\rightarrow {\mathscr {F}}_2\) extending F (i.e. \({\tilde{F}}\vert _{{\mathscr {U}}}=F\)) such that \({\tilde{F}}\) is differentiable of order m (resp. smooth). For completely arbitrary \({\mathscr {U}}\), the derivatives of F on \({\mathscr {U}}\) depend on the choice of extension \({\tilde{F}}\) (take for instance \({\mathscr {U}}=\{\varphi \}\) for some \(\varphi \in {\mathscr {F}}_1\)). However, if \({\mathscr {U}}\) happens to have a nonvoid interior, then it is easily shown that the derivatives of F on \({\mathscr {U}}\) do not depend on the choice of extension. Under certain conditions on F, one can weaken this condition (see, for instance, Remark 2.4).

Remark A.2

For Mackey-complete locally convex topological vector spaces (also called \(c^\infty \)-complete or convenient topological vector spaces), convenient smoothness enjoys essentially all the rules of Calculus presented in the remainder of this Appendix assuming MB differentiability (see e.g. footnote 11 below). Moreover, for Fréchet spaces (which are convenient and whose topology coincides with the corresponding \(c^\infty \)-topology) convenient and MB smoothness coincide (see e.g. Theorem 1, pp. 77 of [39] together with Theorem 2.14, pp. 20–21 of [63]).

Let \(\gamma :[a,b]\rightarrow {\mathscr {F}}\), \(a<b\in {\mathbb {R}}\), be a continuous curve segment in the complete locally convex topological vector space \({\mathscr {F}}\). We can define the (Riemann) integral of \(\gamma \) along [ab]

$$\begin{aligned} \int ^b_{a}\gamma (\lambda )\mathrm {d}\lambda \in {\mathscr {F}}\end{aligned}$$

as the unique linear map from the space \({\mathscr {C}}([a,b],{\mathscr {F}})\) of continuous curves from [ab] to \({\mathscr {F}}\) into the space \({\mathscr {F}}\) such that:Footnote 11

  1. 1.

    For any continuous linear functional \(u:{\mathscr {F}}\rightarrow {\mathbb {R}}\), we have that \(u\left( \int ^b_{a} \gamma (\lambda )\mathrm {d}\lambda \right) =\int ^b_{a}u(\gamma (\lambda ))\mathrm {d}\lambda \);

  2. 2.

    For any continuous seminorm \(\Vert \cdot \Vert \) on \({\mathscr {F}}\), we have that \(\left\| \int ^b_{a} \gamma (\lambda )\mathrm {d}\lambda \right\| \le \int ^b_{a}\Vert \gamma (\lambda )\Vert \mathrm {d}\lambda \);

  3. 3.

    If \(a<c<b\in {\mathbb {R}}\), then \(\int ^b_{a}\gamma (\lambda )\mathrm {d}\lambda =\int ^c_{a}\gamma (\lambda ) \mathrm {d}\lambda +\int ^b_{c}\gamma (\lambda )\mathrm {d}\lambda \).

The Fundamental Theorem of Calculus holds for the Riemann integral of curves taking values in \({\mathscr {F}}\):

Theorem A.1

([45], Theorems 2.2.3 and 2.2.2). Let \(\gamma _0:[a,b]\rightarrow {\mathscr {F}}\) be a continuous curve, \(a\le t\le b\), and define \(\gamma _1(t)\doteq \int ^t_{a}\gamma _0(\lambda )\mathrm {d}\lambda \). Then \(\gamma _1:[a,b]\rightarrow {\mathscr {F}}\) is a \({\mathscr {C}}^1\) curve, and \(\gamma '_1(t)=\gamma _0(t)\). Conversely, if \(\gamma _1:[a,b]\rightarrow {\mathscr {F}}\) is a \({\mathscr {C}}^1\) curve, then \(\gamma _1(b)-\gamma _1(a)=\int ^b_{a}\gamma '_1(\lambda )\mathrm {d}\lambda \)\(\quad \square \)

Corollary A.1

([45], Theorem 3.2.2). Let \(F:{\mathscr {U}}\subset {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\) be a continuous map with \({\mathscr {F}}_2\) complete, \(\varphi _0 \in {\mathscr {U}}\), and \(\mathbf {\varphi }\in {\mathscr {U}}-\varphi _0\doteq \{\varphi -\varphi _0\in {\mathscr {F}}_1\ |\ \varphi \in {\mathscr {U}}\}\). Assume that \({\mathscr {U}}\) is convex for simplicity. If F is differentiable of order one in the sense of Definition A.2, then

$$\begin{aligned} F(\varphi _0+\mathbf {\varphi })-F(\varphi _0)=\int ^1_0F^{(1)} [\varphi _0+\lambda \mathbf {\varphi }] (\mathbf {\varphi })\mathrm {d}\lambda . \end{aligned}$$
(A.2)

With the aid of the fundamental theorem of Calculus A.1, the following key results can be proven. First, the usual linearity property for first-order derivatives holds:

Lemma A.1

([45], Lemma 3.2.3 and Theorem 3.2.5). Let \(F:{\mathscr {U}}\subset {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\) be a continuous map with \({\mathscr {F}}_2\) complete, \(\varphi \in {\mathscr {U}}\). If F is differentiable of order one in the sense of Definition A.2, then for all scalars \(\lambda ,\mu \) and all \(\mathbf {\varphi },\mathbf {\varphi }'\in {\mathscr {F}}_1\) we have that

$$\begin{aligned} F^{(1)}[\varphi ](\lambda \mathbf {\varphi }+\mu \mathbf {\varphi }')=\lambda F^{(1)}[\varphi ](\mathbf {\varphi }) +\mu F^{(1)}[\varphi ](\mathbf {\varphi }'). \end{aligned}$$

\(\square \)

Next, the chain rule holds:

Theorem A.2

([45], Theorem 3.3.4). Let \(F:{\mathscr {U}}\subset {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\), \(G:{\mathscr {V}}\subset {\mathscr {F}}_2\rightarrow {\mathscr {F}}_3\) be respectively continuous maps from open subsets \({\mathscr {U}},{\mathscr {V}}\) of locally convex topological vector spaces \({\mathscr {F}}_1,{\mathscr {F}}_2\) into \({\mathscr {F}}_2\) and the locally convex topological vector space \({\mathscr {F}}_3\), such that \(F({\mathscr {U}})\subset {\mathscr {V}}\). Suppose that \({\mathscr {F}}_2\) and \({\mathscr {F}}_3\) are complete. If F (resp. G) is once differentiable on \({\mathscr {U}}\) (resp. \({\mathscr {V}}\)) in the sense of Definition A.2, then for all \(\varphi \in {\mathscr {U}}\), \(\mathbf {\varphi }\in {\mathscr {F}}_1\) we have that

$$\begin{aligned} (G\circ F)^{(1)}(\varphi )(\mathbf {\varphi })= G^{(1)}[F(\varphi )](F^{(1)}[\varphi ](\mathbf {\varphi })). \end{aligned}$$
(A.3)

\(\square \)

The chain rule (A.3) yields, after taking direct sums, the Leibniz’s rule for derivatives of composition of n-tuples of maps \(F_1,\ldots ,F_n\) with a continuous n-linear map \(\psi \)

$$\begin{aligned} (\psi (F_1,\ldots ,F_n))^{(1)}[\varphi ](\mathbf {\varphi })= \sum ^n_{j=1}\psi (F_1[\varphi ], \ldots ,F^{(1)}_j[\varphi ](\mathbf {\varphi }),\ldots ,F_n[\varphi ]). \end{aligned}$$
(A.4)

This, together with the fundamental theorem of Calculus (A.2), yields the integration by parts formula and, even more importantly, Taylor’s formula with (integral) remainder

$$\begin{aligned} F(\varphi _0+\mathbf {\varphi })= & {} \sum ^k_{j=0}\frac{1}{j!}F^{(j)}[\varphi _0] (\mathbf {\varphi },\ldots ,\mathbf {\varphi }) \nonumber \\&+\int ^1_0\frac{(1-\lambda )^k}{k!} F^{(k+1)} [\varphi _0+\lambda \mathbf {\varphi }](\mathbf {\varphi },\ldots ,\mathbf {\varphi }) \mathrm {d}\lambda . \end{aligned}$$
(A.5)

To see this, note that Leibniz’s rule implies the following key formula:

$$\begin{aligned} \begin{aligned}&\frac{(1-\lambda )^{k-1}}{(k-1)!}F^{(k)}[\varphi _0+\lambda \mathbf {\varphi }](\mathbf {\varphi }, \ldots ,\mathbf {\varphi }) \\&\quad =\frac{(1-\lambda )^k}{k!}F^{(k+1)}[\varphi _0+\lambda \mathbf {\varphi }] (\mathbf {\varphi },\ldots ,\mathbf {\varphi })\\&\qquad -\frac{\mathrm {d}}{\mathrm {d}\lambda } \left[ \frac{(1-\lambda )^k}{k!}F^{(k)} [\varphi _0+\lambda \mathbf {\varphi }](\mathbf {\varphi },\ldots , \mathbf {\varphi })\right] . \end{aligned} \end{aligned}$$
(A.6)

Integrating both sides of formula (A.6) from \(\lambda =0\) to \(\lambda =1\) by means of the fundamental theorem of Calculus (A.2) yields the fundamental induction step from \(k-1\) to k. Since the case \(k=0\) of (A.5) is settled by the fundamental theorem of Calculus itself, we are done.

For the convenience of the reader, we prove the generalization of the chain rule (A.3) for higher derivatives, since this proof is not easy to find in the literature at the present level of generality. We follow the argument employed in [56].

Corollary A.2

(Faà di Bruno’s formula). Let \(F:{\mathscr {U}}\subset {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\), \(G:{\mathscr {V}}\subset {\mathscr {F}}_2\rightarrow {\mathscr {F}}_3\) satisfy the hypotheses of Theorem A.2. If F (resp. G) is m-times differentiable on \({\mathscr {U}}\) (resp. \({\mathscr {V}}\)), then \(G\circ F\) is also m-times differentiable on \({\mathscr {U}}\), and for all \(1\le k\le m\),

$$\begin{aligned} (G\circ F)^{(k)}[\varphi ](\mathbf {\varphi }_1,\ldots , \mathbf {\varphi }_k)=\sum _{\pi \in P_k} G^{(|\pi |)}[F(\varphi )]\left( \bigotimes _{I\in \pi } F^{(|I|)}[\varphi ](\otimes _{j\in I} \mathbf {\varphi }_j)\right) ,\qquad \end{aligned}$$
(A.7)

where \(P_k\) is the set of all partitions \(\pi =\{I_1,\ldots ,I_l\}\) of \(\{1,\ldots ,k\}\), that is, \(I_j\ne \varnothing \), \(I_j\cap I_{j'}=\varnothing \) for \(j\ne j'\) and \(\cup ^l_{j=1}I_j=\{1,\ldots ,k\}\).

Proof

We proceed by induction on k. The case \(k=1\) is just the usual chain rule (A.3). Assume that the formula is valid up to order \(k-1\) along \(\mathbf {\varphi }_1,\ldots , \mathbf {\varphi }_{k-1}\). Then for each partition \(\pi \) of \(\{1,\ldots ,k-1\}\) in the above sum we have, by Leibniz’s rule (A.4),

$$\begin{aligned} \begin{aligned}&\Bigg [G^{(|\pi |)}\left. \circ \,F\left( \bigotimes _{I\in \pi } F^{(|I|)}(\otimes _{j\in I}\mathbf {\varphi }_j) \right) \right] ^{(1)}[\varphi ](\mathbf {\varphi }_k) \\&\quad =G^{(|\pi |+1)}[F(\varphi )]\left( F^{(1)}[\varphi ] (\mathbf {\varphi }_k)\otimes \bigotimes _{I\in \pi } F^{(|I|)}[\varphi ](\otimes _{j\in I}\mathbf {\varphi }_j)\right) \\&\qquad +\sum _{I'\in \pi }G^{(|\pi |)}[F(\varphi )]\Bigg (F^{(|I'|+1)} [\varphi ] \bigg (\mathbf {\varphi }_k\otimes \bigotimes _{j\in I'}\mathbf {\varphi }_j\bigg )\\&\qquad \otimes \bigotimes _{I\in \pi {\smallsetminus }\{I'\}} F^{(|I|)}[\varphi ](\otimes _{l\in I}\mathbf {\varphi }_l)\Bigg ). \end{aligned} \end{aligned}$$

However, any partition \(\pi '\) of \(\{1,\ldots ,k\}\) is either of the form \(\pi '=\{\{k\}\}\cup \pi \) or \(\pi '=(\pi {\smallsetminus }\{I'\})\cup \{I'\cup \{k\}\}\) for some \(I'\in \pi \), \(\pi \in P_{k-1}\). Hence, summing the above identities over all such \(\pi \) gives the desired result. \(\quad \square \)

A consequence of Faà di Bruno’s formula (A.7) is the generalization of Leibniz’s rule (A.4) for higher order derivatives of composition of l-tuples of maps \(F_1,\ldots ,F_l\) with a continuous l-linear map \(\psi \)

$$\begin{aligned} \begin{aligned}&(\psi (F_1,\ldots ,F_l))^{(k)}[\varphi ]\left( \mathbf {\varphi }_1, \ldots ,\mathbf {\varphi }_k\right) \\&\quad =\sum _{\{I_1,\ldots ,I_l\}\in {\tilde{P}}_{k,l}}\psi \left( F_1^{(|I_1|)}[\varphi ](\otimes _{j\in I_1} \mathbf {\varphi }_j),\ldots ,F_l^{(|I_l|)}[\varphi ] (\otimes _{j\in I_l}\mathbf {\varphi }_j)\right) , \end{aligned} \end{aligned}$$
(A.8)

where \({\tilde{P}}_{k,l}\) is the set of all partitions \(\pi =\{I_1,\ldots ,I_l\}\) of \(\{1,\ldots ,k\}\) in lpossibly (but not all) empty subsets, i.e. \(I_j\cap I_{j'}=\varnothing \) for \(j\ne j'\) and \(\cup ^l_{j=1}I_j=\{1,\ldots ,k\}\). As another application, we obtain the so-called k-th order resolvent formula (A.12) below which shall often be useful. Consider two MB-differentiable maps \(F:{\mathscr {U}}\times {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\), \(G:{\mathscr {U}}\times {\mathscr {F}}_2\rightarrow {\mathscr {F}}_1\) of order one, where \({\mathscr {F}}_1,{\mathscr {F}}_2\) are locally convex topological vector spaces and \({\mathscr {U}}\subset {\mathscr {F}}\) is a nonvoid open subset of the locally convex topological vector space \({\mathscr {F}}\). For notational convenience, we also occasionally write \(F(\varphi ,\mathbf {\varphi })\doteq F[\varphi ] \mathbf {\varphi }\), \(G(\varphi ,\mathbf {\psi })\doteq G[\varphi ]\mathbf {\psi }\). Suppose that both F and G are linear in their second arguments and satisfy

$$\begin{aligned} \begin{aligned} F[\varphi ]G[\varphi ]\mathbf {\psi }&=\mathbf {\psi }\ ,\quad \forall \varphi \in {\mathscr {U}},\, \mathbf {\psi }\in {\mathscr {F}}_2,\\ G[\varphi ]F[\varphi ]\mathbf {\varphi }&=\mathbf {\varphi }\ ,\quad \forall \varphi \in {\mathscr {U}},\, \mathbf {\varphi }\in {\mathscr {F}}_1. \end{aligned} \end{aligned}$$
(A.9)

If we define

$$\begin{aligned}&D^k_1F[\varphi ](\mathbf {\varphi }_1,\ldots ,\mathbf {\varphi }_k) \mathbf {\varphi }=F^{(k)}[\varphi , \mathbf {\varphi }] ((\mathbf {\varphi }_1,0),\ldots ,(\mathbf {\varphi }_k,0))\ , \nonumber \\&D^1_1\doteq D_1,\, D^0_1=\mathbb {1}, \end{aligned}$$
(A.10)

then by the chain rule (A.3) applied to the pair of maps \(F,(\mathbb {1},G)\) and (A.9) we have the (first-order) resolvent formula

$$\begin{aligned} D_1G[\varphi ](\mathbf {\varphi }_1)\mathbf {\psi }= -G[\varphi ]D_1F[\varphi ](\mathbf {\varphi }_1)G[\varphi ] \mathbf {\psi }. \end{aligned}$$
(A.11)

It follows from the above formula that if in addition F is MB-smooth, then so is G. More precisely, in this case we obtain the following (not so pleasant) higher-order generalization of (A.11), obtained by induction on \(k\ge 1\) from (A.11) and an argument analogous to the one used in the proof of Corollary A.2:

$$\begin{aligned} \begin{aligned}&D^k_1G[\varphi ]\left( \mathbf {\varphi }_1,\ldots , \mathbf {\varphi }_k\right) \mathbf {\psi }\\&\quad =\sum ^k_{l=1}(-1)^l\sum _{\{I_1,\ldots ,I_l\} \in P_k}\sum _{\sigma \in S_l}\left( \prod ^l_{j=1}G[\varphi ] D^{|I_{\sigma (j)}|}F[\varphi ](\otimes _{i\in I_{\sigma (j)}}\mathbf {\varphi }_i)\right) G[\varphi ] \mathbf {\psi }.\qquad \end{aligned} \end{aligned}$$
(A.12)

Here, \(P_k\) is again the set of all partitions of \(\{1,\ldots ,k\}\) as in the statement of Corollary A.2, whereas \(S_l\) is the set of all permutations of \(\{1,\ldots ,l\}\).

Finally, one can show that the order of differentiation for higher order derivatives is irrelevant:

Theorem A.3

([45], Theorem 3.6.2). Let \(F:{\mathscr {U}}\subset {\mathscr {F}}_1\rightarrow {\mathscr {F}}_2\) be a continuous map with \({\mathscr {F}}_2\) complete. If F is differentiable of order \(m>1\) in the sense of Definition A.2, then \(F^{(k)} [\varphi ]:{\mathscr {F}}^k_1\ni (\mathbf {\varphi }_1,\ldots ,\mathbf {\varphi }_k)\mapsto F^{(k)}[\varphi ] (\mathbf {\varphi }_1,\ldots ,\mathbf {\varphi }_k)\in {\mathscr {F}}_2\) is a symmetric, k-linear map for all fixed \(\varphi \in {\mathscr {U}}\), \(2\le k\le m\)\(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Brunetti, R., Fredenhagen, K. & Ribeiro, P.L. Algebraic Structure of Classical Field Theory: Kinematics and Linearized Dynamics for Real Scalar Fields. Commun. Math. Phys. 368, 519–584 (2019). https://doi.org/10.1007/s00220-019-03454-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03454-z

Navigation