global-analysis-and-the-banach-manifold-of-class-h1-curvers

Riemannian Geometry course project on the manifold H¹(I, M) of class H¹ curves on a Riemannian manifold M and its applications to the geodesics problem

introduction.tex (15155B)

  1 \section{Introduction}\label{sec:introduction}
  2 
  3 Known as \emph{global analysis}, or sometimes \emph{non-linear functional
  4 analysis}, the field of study dedicated to the understanding of
  5 infinite-dimensional manifolds has seen remarkable progress in the past several
  6 decades. Among numerous discoveries, perhaps the greatest achievement in global
  7 analysis in the last century was the recognition of the fact that many
  8 interesting function spaces possess natural differentiable structures -- which
  9 are usually infinite-dimensional.
 10 
 11 As it turns out, many local problems regarding maps between finite-dimensional
 12 manifolds can be translated to global questions about the geometry of function
 13 spaces -- hence the name ``\emph{global} analysis''. More specifically, a
 14 remarkable number of interesting geometric objects can be characterized as
 15 ``critical points'' of functionals in functions spaces. The usual suspects are,
 16 of course, geodesics and minimal submanifolds in general, but there are many
 17 other interesting examples: harmonic functions, Einstein metrics, periodic
 18 solutions to Hamiltonian vector fields, etc. \cite[ch.~11]{palais}.
 19 
 20 Such objects are the domain of the so called \emph{calculus of variations},
 21 which is generally concerned with finding functions that minimize or maximize a
 22 given functional, known as the \emph{action functional}, by subjecting such
 23 functions to ``small variations'' -- which is known as \emph{the variational
 24 method}. The meaning of ``small variations'' have historically been a very
 25 dependent on the context of the problem at hand. Only recently, with the
 26 introduction of the tools of global analysis, the numerous ad-hoc methods under
 27 the umbrella of ``variational method'' have been unified into a coherent
 28 theory, which we describe in the following.
 29 
 30 By viewing the class of functions we're interested in as a -- most likely
 31 infinite-dimensional -- manifold \(\mathscr{F}\) and the action functional as a
 32 smooth functional \(f : \mathscr{F} \to \mathbb{R}\) we can find minimizing and
 33 maximizing functions by studying the critical points of \(f\). More generally,
 34 modern calculus of variations is concerned with the study of critical points of
 35 smooth functionals \(\Gamma(E) \to \mathbb{R}\), where \(E \to M\) is a smooth
 36 fiber bundle over a finite-dimensional manifold \(M\) and \(\Gamma\) is a given
 37 section functor, such as smooth sections, continuous sections or Sobolev
 38 sections -- notice that by taking \(E = M \times N\) the manifold \(\Gamma(E)\)
 39 is naturally identified with a space of functions \(M \to N\), which gets us
 40 back to the original case.
 41 
 42 In these notes we hope to provide a very brief introduction to modern theory
 43 the calculus of variations by exploring one of the simplest concrete examples
 44 of the previously described program. We study the differential structure of the
 45 Banach manifold \(H^1(I, M)\) of class \(H^1\) curves in a finite-dimensional
 46 Riemannian manifold \(M\), which encodes the solution to the \emph{classic}
 47 variational problem: that of geodesics. Hence the particular action functional
 48 we are interested is the infamous \emph{energy functional}
 49 \begin{align*}
 50   E : H^1(I, M) & \to     \mathbb{R}                                         \\
 51          \gamma & \mapsto \frac{1}{2} \int_0^1 \norm{\dot\gamma(t)}^2 \; \dt,
 52 \end{align*}
 53 as well as the \emph{length functional}
 54 \begin{align*}
 55   L : H^1(I, M) & \to \mathbb{R}                               \\
 56          \gamma & \mapsto \int_0^1 \norm{\dot\gamma(t)} \; \dt
 57 \end{align*}
 58 
 59 In section~\ref{sec:structure} we will describe the differential structure of
 60 \(H^1(I, M)\) and its canonical Riemannian metric. In
 61 section~\ref{sec:aplications} we study the critical points of the energy
 62 functional \(E\) and describe how the fundamental results of the classical
 63 theory of the calculus of variations in the context of Riemannian manifolds can
 64 be reproduced in our new setting. Other examples of function spaces are
 65 explored in detail in \cite[sec.~6]{eells}. The 11th chapter of \cite{palais}
 66 is also a great reference for the general theory of spaces of sections of fiber
 67 bundles.
 68 
 69 We should point out that we will primarily focus on
 70 the broad strokes of the theory ahead and that we will leave many results
 71 unproved. The reasoning behind this is twofold. First, we don't want to bore
 72 the reader with the numerous technical details of some of the constructions
 73 we'll discuss in the following. Secondly, and this is more important, these
 74 notes are meant to be concise. Hence we do not have the necessary space to
 75 discuss neither technicalities nor more involved applications of the theory we
 76 will develop.
 77 
 78 In particular, we leave the intricacies of Palais' and Smale's discussion of
 79 condition (C) -- which can be seen as a substitute for the failure of a proper
 80 Hilbert space to be locally compact \cite[ch.~2]{klingenberg} -- and its
 81 applications to the study of closed geodesics out of these notes. As previously
 82 stated, many results are left unproved, but we will include references to other
 83 materials containing proofs. We'll assume basic knowledge of differential and
 84 Riemannian geometry, as well as some familiarity with the classical theory of
 85 the calculus of variations -- see \cite[ch.~5]{gorodski} for the classical
 86 approach.
 87 
 88 Before moving to the next section we would like to review the basics of the
 89 theory of real Banach manifolds.
 90 
 91 \subsection{Banach Manifolds}
 92 
 93 While it is certainly true that Banach spaces can look alien to someone who has
 94 never ventured outside of the realms of Euclidean space, Banach manifolds are
 95 surprisingly similar to their finite-dimensional counterparts. As we'll soon
 96 see, most of the usual tools of differential geometry can be quite easily
 97 translated to the context of Banach manifolds\footnote{The real difficulties
 98 with Banach manifolds only show up while proving certain results, and are
 99 mainly due to complications regarding the fact that not all closed subspaces of
100 a Banach space have a closed complement.}. The reason behind this is simple: it
101 turns out that calculus has nothing to do with \(\mathbb{R}^n\).
102 
103 What we mean by this last statement is that none of the fundamental ingredients
104 of calculus -- the ones necessary to define differentiable functions in
105 \(\mathbb{R}^n\), namely the fact that \(\mathbb{R}^n\) is a complete normed
106 space -- are specific to \(\mathbb{R}^n\). In fact, these ingredients are
107 precisely the features of a Banach space. Thus we may naturally generalize
108 calculus to arbitrary Banach spaces, and consequently generalize smooth
109 manifolds to spaces modeled after Banach spaces. We begin by the former.
110 
111 \begin{definition}
112   Let \(V\) and \(W\) be Banach spaces and \(U \subset V\) be an open subset. A
113   continuous map \(f : U \to W\) is called \emph{differentiable at \(p \in U\)}
114   if there exists a continuous linear operator \(d f_p \in \mathcal{L}(V, W)\)
115   such that
116   \[
117     \frac{\norm{f(p + h) - f(p) - d f_p h}}{\norm{h}} \to 0
118   \]
119   as \(h \to 0\) in \(V\).
120 \end{definition}
121 
122 \begin{definition}
123   Given Banach spaces \(V\) and \(W\) and an open subset \(U \subset V\), a
124   continuous map \(f : U \to W\) is called \emph{differentiable of class
125   \(C^1\)} if \(f\) is differentiable at \(p\) for all \(p \in U\) and the
126   \emph{derivative map}
127   \begin{align*}
128     df: U & \to     \mathcal{L}(V, W) \\
129         p & \mapsto d f_p
130   \end{align*}
131   is continuous. Since \(\mathcal{L}(V, W)\) is a Banach space under the
132   operator norm, we may recursively define functions of class \(C^n\) for \(n >
133   1\): a function \(f : U \to W\) of class \(C^{n - 1}\) is called
134   \emph{differentiable of class \(C^n\)} if the map\footnote{Here we consider
135   the \emph{projective tensor product} of Banach spaces. See
136   \cite[ch.~1]{klingenberg}.}
137   \[
138     d^{n - 1} f :
139     U \to \mathcal{L}(V, \mathcal{L}(V, \cdots \mathcal{L}(V, W)))
140     \cong \mathcal{L}(V^{\otimes n}, W)
141   \]
142   is of class \(C^1\). Finally, a map \(f : U \to W\) is called
143   \emph{differentiable of class \(C^\infty\)} or \emph{smooth} if \(f\) is of
144   class \(C^n\) for all \(n > 0\).
145 \end{definition}
146 
147 The following lemma is also of huge importance, and it is known as \emph{the
148 chain rule}.
149 
150 \begin{lemma}\label{thm:chain-rule}
151   Given Banach spaces \(V_1\), \(V_2\) and \(V_3\), open subsets \(U_1 \subset
152   V_1\) and \(U_2 \subset V_2\) and two smooth maps \(f : U_1 \to U_2\) and \(g
153   : U_2 \to V_3\), the composition map \(g \circ f : U_1 \to V_3\) is smooth
154   and its derivative is given by
155   \[
156     d (g \circ f)_p = d g_{f(p)} \circ d f_p
157   \]
158 \end{lemma}
159 
160 As promised, these simple definitions allows us to expand the usual tools of
161 differential geometry to the infinite-dimensional setting. In fact, in most
162 cases it suffices to simply copy the definition of the finite-dimensional case.
163 For instance, as in the finite-dimensional case we may call a map between
164 Banach manifolds \(M\) and \(N\) \emph{smooth} if it can be locally expressed
165 as a smooth function between open subsets of the model spaces. As such, we will
166 only provide the most important definitions: those of a Banach manifold and its
167 tangent space at a given point. Complete accounts of the subject can be found
168 in \cite[ch.~1]{klingenberg} and \cite[ch.~2]{lang}.
169 
170 \begin{definition}\label{def:banach-manifold}
171   A Banach manifold \(M\) is a Hausdorff topological space endowed with a
172   maximal atlas \(\{(U_i, \varphi_i)\}_i\), i.e. an open cover \(\{U_i\}_i\) of
173   \(M\) and homeomorphisms \(\varphi_i : U_i \to \varphi_i(U_i) \subset V_i\)
174   -- known as \emph{charts} -- where
175   \begin{enumerate}
176     \item Each \(V_i\) is a Banach space
177     \item For each \(i\) and \(j\), \(\varphi_i \circ \varphi_j^{-1} :
178       \varphi_j(U_i \cap U_j) \subset V_j \to \varphi_i(U_i \cap U_j) \subset
179       V_i\) is a smooth map
180     \item \(\{(U_i, \varphi_i)\}_i\) is maximal with respect to the items above
181   \end{enumerate}
182 \end{definition}
183 
184 \begin{definition}
185   Given a Banach manifold \(M\) with maximal atlas \(\{(U_i, \varphi_i)\}_i\)
186   and \(p \in M\), the tangent space \(T_p M\) of \(M\) at \(p\) is the
187   quotient of the space \(\{ \gamma : (- \epsilon, \epsilon) \to M \mid \gamma
188   \ \text{is smooth}, \gamma(0) = p \}\) by the equivalence relation that
189   identifies two curves \(\gamma\) and \(\eta\) such that
190   \((\varphi_i \circ \gamma)'(0) = (\varphi_i \circ \eta)'(0)\) for all \(i\)
191   with \(p \in U_i\).
192 \end{definition}
193 
194 \begin{definition}
195   Given \(p \in M\) and a chart \(\varphi_i : U_i \to V_i\) with \(p \in U_i\),
196   let
197   \begin{align*}
198     \phi_{p, i} :    T_p M & \to     V_i                          \\
199                   [\gamma] & \mapsto (\varphi_i \circ \gamma)'(0)
200   \end{align*}
201 \end{definition}
202 
203 \begin{proposition}\label{thm:tanget-space-topology}
204   Given \(p \in M\) and a chart \(\varphi_i\) with \(p \in U_i\), \(\phi_{p,
205   i}\) is a linear isomorphism. For any given charts \(\varphi_i, \varphi_j\),
206   the pullback of the norms of \(V_i\) and \(V_j\) by \(\varphi_i\) and
207   \(\varphi_j\) respectively define equivalent norms in \(T_p M\). In
208   particular, any choice chart gives \(T_p M\) the structure of a topological
209   vector space, and this topology is independent of this choice\footnote{In
210   general $T_p M$ is not a normed space, since the norms induced by two
211   distinct choices of chard need not to coincide. Nevertheless, the topology
212   induced by these norms is the same.}.
213 \end{proposition}
214 
215 \begin{proof}
216   The first statement about \(\phi_{p, i}\) being a linear isomorphism should
217   be clear from the definition of \(T_p M\). The second statement about the
218   equivalence of the norms is equivalent to checking that \(\phi_{p, i} \circ
219   \phi_{p, j}^{-1} : V_j \to V_i\) is continuous for each \(i\) and \(j\) with
220   \(p \in U_i\) and \(p \in U_j\).
221 
222   But this follows immediately from the identity
223   \[
224     \begin{split}
225       (\phi_{p, i} \circ \phi_{p, j}^{-1}) v
226       & = (\varphi_i \circ \varphi_j^{-1} \circ \gamma_v)'(0) \\
227       \text{(chain rule)}
228       & = d (\varphi_i \circ \varphi_j^{-1})_{\varphi_j(p)} \dot\gamma_v(0) \\
229       & = d (\varphi_i \circ \varphi_j^{-1})_{\varphi_j(p)} v
230     \end{split}
231   \]
232   where \(v \in V_j\) and \(\gamma_v : (-\epsilon, \epsilon) \to V_j\) is any
233   smooth curve with \(\gamma_v(0) = \varphi_j(p)\) and \(\dot\gamma_v(0) = v\):
234   \(\phi_{p, i} \circ \phi_{p, j}^{-1} = d (\varphi_i \circ
235   \varphi_j^{-1})_{\varphi_j(p)}\) is continuous by definition.
236 \end{proof}
237 
238 Notice that a single Banach manifold may be ``modeled after'' multiple Banach
239 spaces, in the sense that the \(V_i\)'s of definition~\ref{def:banach-manifold}
240 may vary with \(i\). Lemma~\ref{thm:chain-rule} implies that for each \(i\) and
241 \(j\) with \(p \in U_i \cap U_j\), \((d \varphi_i \circ
242 \varphi_j^{-1})_{\varphi_j(p)} : V_j \to V_i\) is a continuous linear
243 isomorphism, so that we may assume that each connected component of \(M\) is
244 modeled after a single Banach space \(V\). It is sometimes convenient, however,
245 to allow ourselves the more lenient notion of Banach manifold afforded by
246 definition~\ref{def:banach-manifold}.
247 
248 We should also note that some authors assume that both the \(V_i\)'s and \(M\)
249 itself are \emph{separable}, in which case the assumption that \(M\) is
250 Hausdorff is redundant. Although we are primarily interested in manifolds
251 modeled after separable spaces, in the interest of affording ourselves a
252 greater number of examples we will \emph{not} assume separability -- unless
253 explicitly stated otherwise. Speaking of examples\dots
254 
255 \begin{example}
256   Any Banach space \(V\) can be seen as a Banach manifold with atlas given by
257   \(\{(V, \operatorname{id} : V \to V)\}\) -- sometimes called \emph{an affine
258   Banach manifold}. In fact, any open subset \(U \subset V\) of a Banach space
259   \(V\) is a Banach manifold under a global chart \(\operatorname{id} : U \to
260   V\).
261 \end{example}
262 
263 \begin{example}
264   The group of units \(A^\times\) of a Banach algebra \(A\) is an open subset,
265   so that it constitutes a Banach manifold modeled after \(A\)
266   \cite[sec.~3]{eells}. In particular, given a Banach space \(V\) the group
267   \(\operatorname{GL}(V)\) of continuous linear isomorphisms \(V \to V\) is a
268   -- possibly non-separable -- Banach manifold modeled after the space
269   \(\mathcal{L}(V) = \mathcal{L}(V, V)\) under the operator norm:
270   \(\operatorname{GL}(V) = \mathcal{L}(V)^\times\).
271 \end{example}
272 
273 \begin{example}
274   Given a complex Hilbert space \(H\), the space \(\operatorname{U}(H)\) of
275   unitary operators \(H \to H\) -- endowed with the topology of the operator
276   norm -- is a Banach manifold modeled after the closed subspace
277   \(\mathfrak{u}(H) \subset \mathcal{L}(H)\) of continuous skew-symmetric
278   operators \(H \to H\) \cite[p.~4]{unitary-group-strong-topology}.
279 \end{example}
280 
281 These last two examples are examples of Banach Lie groups -- i.e. Banach
282 manifolds endowed with a group structure whose group operations are smooth.
283 Perhaps more interesting to us is the fact that these are both examples of
284 function spaces. Having reviewed the basics of the theory of Banach manifolds
285 we can proceed to our in-depth exploration of a particular example, that of the
286 space \(H^1(I, M)\).
287