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

structure.tex (24271B)

  1 \section{The Structure of \(H^1(I, M)\)}\label{sec:structure}
  2 
  3 Throughout this section let \(M\) be a finite-dimensional Riemannian manifold.
  4 As promised, in this section we will highlight the differential and Riemannian
  5 structures of the space \(H^1(I, M)\) of class \(H^1\) curves in a \(M\). The
  6 first question we should ask ourselves is an obvious one: what is \(H^1(I,
  7 M)\)? Specifically, what is a class \(H^1\) curve in \(M\)?
  8 
  9 Given an interval \(I\), recall that a continuous curve \(\gamma : I \to
 10 \mathbb{R}^n\) is called \emph{a class \(H^1\)} curve if \(\gamma\) is
 11 absolutely continuous, \(\dot \gamma(t)\) exists for almost all \(t \in I\) and
 12 \(\dot\gamma \in H^0(I, \mathbb{R}^n) = L^2(I, \mathbb{R}^n)\). It is a well
 13 known fact that the so called \emph{Sobolev space \(H^1([0, 1],
 14 \mathbb{R}^n)\)} of all class \(H^1\) curves in \(\mathbb{R}^n\) is a Hilbert
 15 space under the inner product given by
 16 \[
 17   \langle \gamma, \eta \rangle_1
 18   = \int_0^1 \gamma(t) \cdot \eta(t) + \dot\gamma(t) \cdot \dot\eta(t) \; \dt
 19 \]
 20 
 21 Finally, we may define\dots
 22 
 23 \begin{definition}
 24   Given an \(n\)-dimensional manifold \(M\), a continuous curve \(\gamma : I
 25   \to M\) is called \emph{a class \(H^1\)} curve if \(\varphi_i \circ \gamma :
 26   J \to \mathbb{R}^n\) is a class \(H^1\) curve for any chart \(\varphi_i : U_i
 27   \subset M \to \mathbb{R}^n\) -- i.e. if \(\gamma\) can be locally expressed
 28   as a class \(H^1\) curve in terms of the charts of \(M\). We'll denote by
 29   \(H^1(I, M)\) the set of all class \(H^1\) curves \(I \to M\).
 30 \end{definition}
 31 
 32 \begin{note}
 33   From now on we fix \(I = [0, 1]\).
 34 \end{note}
 35 
 36 Notice in particular that every piece-wise smooth curve \(\gamma : I \to M\) is
 37 a class \(H^1\) curve. This answer raises and additional question though: why
 38 class \(H^1\) curves? The classical theory of the calculus of variations -- as
 39 described in \cite[ch.~5]{gorodski} for instance -- is usually exclusively
 40 concerned with the study of piece-wise smooth curves, so the fact that we are
 41 now interested a larger class of curves -- highly non-smooth curves, in fact --
 42 \emph{should} come as a surprise to the reader.
 43 
 44 To answer this second question we return to the case of \(M = \mathbb{R}^n\).
 45 Denote by \({C'}^\infty(I, \mathbb{R}^n)\) the space of piece-wise curves in
 46 \(\mathbb{R}^n\). As described in section~\ref{sec:introduction}, we would like
 47 \({C'}^\infty(I, \mathbb{R}^n)\) to be a Banach manifold under which both the
 48 energy functional and the length functional are smooth maps. As most function
 49 spaces, \({C'}^\infty(I, \mathbb{R}^n)\) admits several natural topologies.
 50 Some of the most obvious candidates are the uniform topology and the topology
 51 of the \(\norm\cdot_0\) norm, which are the topologies induced by the norms
 52 \begin{align*}
 53   \norm{\gamma}_\infty & = \sup_t \norm{\gamma(t)}                   \\
 54        \norm{\gamma}_0 & = \sqrt{\int_0^1 \norm{\gamma(t)}^2 \; \dt}
 55 \end{align*}
 56 respectively.
 57 
 58 The problem with the first candidate is that \(L : {C'}^\infty(I, \mathbb{R}^n)
 59 \to \mathbb{R}\) is not a continuous map under the uniform topology. This can
 60 be readily seen by approximating the curve
 61 \begin{align*}
 62   \gamma : I & \to     \mathbb{R}^2      \\
 63            t & \mapsto (t, 1 - t)
 64 \end{align*}
 65 with ``staircase curves'' \(\gamma_n : I \to \mathbb{R}^n\) for larger and
 66 larger values of \(n\), as shown in figure~\ref{fig:step-curves}: clearly
 67 \(\gamma_n \to \gamma\) in the uniform topology, but \(L(\gamma_n) = 2\) does
 68 not approach \(L(\gamma) = \sqrt 2\) as \(n\) approaches \(\infty\).
 69 
 70 \begin{figure}[h]
 71   \centering
 72   \begin{tikzpicture}
 73     \draw (4, 1) -- (1, 4);
 74     \draw (4, 1) -- (4, 2)
 75                  -- (3, 2)
 76                  -- (3, 3) node[right]{$\gamma_n$}
 77                  -- (2, 3) node[left]{$\gamma$}
 78                  -- (2, 4)
 79                  -- (1, 4);
 80     \draw[dotted] (4.5, .5) -- (4, 1);
 81     \draw[dotted] (.5, 4.5) -- (1, 4);
 82     \draw (1, 4.3) -- (2, 4.3);
 83     \draw (1, 4.2) -- (1, 4.4);
 84     \draw (2, 4.2) -- (2, 4.4);
 85     \node[above] at (1.5, 4.3) {$\sfrac{1}{n}$};
 86   \end{tikzpicture}
 87   \caption{A diagonal line representing the curve $\gamma$ overlaps a
 88   staircase-like curve $\gamma_n$, whose steps measure $\sfrac{1}{n}$ in
 89   width and height.}
 90   \label{fig:step-curves}
 91 \end{figure}
 92 
 93 The issue with this particular example is that while \(\gamma_n \to \gamma\)
 94 uniformly, \(\dot\gamma_n\) does not converge to \(\dot\gamma\) in the uniform
 95 topology. This hints at the fact that in order for \(E\) and \(L\) to be
 96 continuous maps we need to control both \(\gamma\) and \(\dot\gamma\). Hence a
 97 natural candidate for a norm in \({C'}^\infty(I, \mathbb{R}^n)\) is
 98 \[
 99   \norm{\gamma}_1^2 = \norm{\gamma}_0^2 + \norm{\dot\gamma}_0^2,
100 \]
101 which is, of course, the norm induced by the inner product \(\langle \, ,
102 \rangle_1\) -- here \(\norm{\cdot}_0\) denotes the norm of \(H^0(I,
103 \mathbb{R}^n) = L^2(I, \mathbb{R}^n)\).
104 
105 The other issue we face is one of completeness. Since \(\mathbb{R}^n\) has a
106 global chart, we expect \({C'}^\infty(I, \mathbb{R}^n)\) to be affine too. In
107 other words, it is natural to expect \({C'}^\infty(I, \mathbb{R}^n)\) to be
108 Banach space. In particular, \({C'}^\infty(I, \mathbb{R}^n)\) must be complete.
109 This is unfortunately not the case for \({C'}^\infty(I, \mathbb{R}^n)\) in the
110 \(\norm\cdot_1\) norm, but we can consider its completion. Lo and behold, a
111 classical result by Lebesgue establishes that this completion just so happens
112 to coincide with \(H^1(I, \mathbb{R}^n)\).
113 
114 It's also interesting to note that the completion of \({C'}^\infty(I,
115 \mathbb{R}^n)\) with respect to the norms \(\norm\cdot_\infty\) and
116 \(\norm\cdot_0\) are \(C^0(I, \mathbb{R}^n)\) and \(H^0(I, \mathbb{R}^n)\),
117 respectively, and that the natural inclusions
118 \begin{equation}\label{eq:continuous-inclusions-rn-curves}
119   H^1(I, \mathbb{R}^n)
120   \longhookrightarrow C^0(I, \mathbb{R}^n)
121   \longhookrightarrow H^0(I, \mathbb{R}^n)
122 \end{equation}
123 are continuous.
124 
125 This can be seen as a particular case of a more general result regarding spaces
126 of sections of vector bundles over the unit interval \(I\). Explicitly, we
127 find\dots
128 
129 \begin{proposition}
130   Given an Euclidean bundle \(E \to I\) -- i.e. a vector bundle endowed with a
131   Riemannian metric -- the space \(C^0(E)\) of all continuous sections of \(E\)
132   is the completion of \({C'}^\infty(E)\) under the norm given by
133   \[
134     \norm{\xi}_\infty = \sup_t \norm{\xi_t}
135   \]
136 \end{proposition}
137 
138 \begin{proposition}\label{thm:h0-bundle-is-complete}
139   Given an Euclidean bundle \(E \to I\), the space \(H^0(E)\) of all square
140   integrable sections of \(E\) is the completion of \({C'}^\infty(E)\) under
141   the inner product given by
142   \[
143     \langle \xi, \eta \rangle_0 = \int_0^1 \langle \xi_t, \eta_t \rangle \; \dt
144   \]
145 \end{proposition}
146 
147 \begin{proposition}
148   Given an Euclidean bundle \(E \to I\), the space \(H^1(E)\) of all class
149   \(H^1\) sections of \(E\) is the completion of the space \({C'}^\infty(E)\)
150   of piece-wise smooth sections of \(E\) under the inner product given by
151   \[
152     \langle \xi, \eta \rangle_1
153     = \langle \xi, \eta \rangle_0 +
154     \left\langle
155       \nabla_{\frac\dd\dt} \xi, \nabla_{\frac\dd\dt} \eta
156     \right\rangle_0
157   \]
158 \end{proposition}
159 
160 \begin{proposition}\label{thm:continuous-inclusions-sections}
161   Given an Euclidean bundle \(E \to I\), the inclusions
162   \[
163     H^1(E) \longhookrightarrow C^0(E) \longhookrightarrow H^0(E)
164   \]
165   are continuous. More precisely, \(\norm{\xi}_\infty \le \sqrt 2
166   \norm{\xi}_1\) and \(\norm{\xi}_0 \le \norm{\xi}_\infty\).
167 \end{proposition}
168 
169 \begin{proof}
170   Given \(\xi \in H^0(E)\) we have
171   \[
172     \norm{\xi}_0^2
173     = \int_0^1 \norm{\xi_t}^2 \; \dt
174     \le \int_0^1 \norm{\xi}_\infty^2 \; \dt
175     = \norm{\xi}_\infty^2
176   \]
177 
178   Now given \(\xi \in H^1(E)\) fix \(t_0, t_1 \in I\) with \(\norm{\xi}_\infty
179   = \norm{\xi_{t_1}}\) and \(\norm{\xi_{t_0}} \le \norm{\xi}_0\). If \(t_0 <
180   t_1\) then
181   \[
182     \begin{split}
183       \norm{\xi}_\infty^2
184       & = \norm{\xi_{t_0}}^2
185         + \int_{t_0}^{t_1} \frac{\dd}\dt \norm{\xi_t}^2 \; \dt \\
186       & \le \norm{\xi}_0^2
187         + \int_{t_0}^{t_1} \frac{\dd}\dt \norm{\xi_t}^2 \; \dt \\
188       \text{(\(\nabla\) is compatible with the metric)}
189       & = \norm{\xi}_0^2 + \int_{t_0}^{t_1}
190         2 \left\langle \xi_t, \nabla_{\frac\dd\dt} \xi_t \right\rangle
191         \; \dt \\
192       \text{(Cauchy-Schwarz)}
193       & \le \norm{\xi}_0^2 + \int_0^1
194         2 \norm{\xi_t} \cdot \norm{\nabla_{\frac\dd\dt} \xi_t} \; \dt \\
195       & \le \norm{\xi}_0^2
196         + \int_0^1 \norm{\xi_t}^2 + \norm{\nabla_{\frac\dd\dt} \xi_t}^2
197         \; \dt \\
198       & = \norm{\xi}_0^2
199         + \norm{\xi}_0^2 + \norm{\nabla_{\frac\dd\dt} \xi}_0^2 \\
200       & \le 2 \norm{\xi}_1^2
201     \end{split}
202   \]
203 
204   Similarly, if \(t_0 > t_1\) then
205   \[
206     \norm{\xi}_\infty^2
207     = \norm{\xi_{t_0}}^2 + \int_{t_0}^{t_1} \frac{\dd}\dt \norm{\xi_t}^2 \; \dt
208     = \norm{\xi_{t_0}}^2 + \int_{1 - t_0}^{1 - t_1}
209       \frac{\dd}\dt \norm{\xi_{1 - t}}^2 \; \dt
210     \le 2 \norm{\xi}_1^2
211   \]
212 \end{proof}
213 
214 \begin{note}
215   Apply proposition~\ref{thm:continuous-inclusions-sections} to the trivial
216   bundle \(I \times \mathbb{R}^n \to I\) to get the continuity of the maps in
217   (\ref{eq:continuous-inclusions-rn-curves}).
218 \end{note}
219 
220 We are particularly interested in the case of the pullback bundle \(E =
221 \gamma^* TM \to I\), where \(\gamma : I \to M\) is a piece-wise smooth curve.
222 \begin{center}
223   \begin{tikzcd}
224     \gamma^* TM \arrow{r} \arrow[swap]{d}{\pi} & TM \arrow{d}{\pi} \\
225     I \arrow[swap]{r}{\gamma} & M
226   \end{tikzcd}
227 \end{center}
228 
229 We now have all the necessary tools to describe the differential structure of
230 \(H^1(I, M)\).
231 
232 \subsection{The Charts of \(H^1(I, M)\)}
233 
234 We begin with a technical lemma.
235 
236 \begin{lemma}\label{thm:section-in-open-is-open}
237   Let \(W \subset TM\) be an open neighborhood of the zero section in \(TM\).
238   Given \(\gamma \in {C'}^\infty(I, M)\), denote by \(W_{\gamma, t}\) the set
239   \(W \cap T_{\gamma(t)} M\) and let \(W_\gamma = \bigcup_t W_{\gamma, t}\).
240   Then \(H^1(W_\gamma) = \{ X \in H^1(\gamma^* TM) : X_t \in W_{\gamma, t} \;
241   \forall t \}\) is an open subset of \(H^1(\gamma^* TM)\).
242 \end{lemma}
243 
244 \begin{proof}
245   Let \(C^0(W_\gamma) = \{ X \in C^0(\gamma^* TM) : X_t \in W_{\gamma, t} \;
246   \forall t \}\). We claim \(C^0(W_\gamma)\) is open in \(C^0(\gamma^* TM)\).
247   Indeed, given \(X \in C^0(W_\gamma)\) there exists \(\delta > 0\) such
248   that
249   \[
250     \begin{split}
251       \norm{X - Y}_\infty < \delta
252       & \implies \norm{X_t - Y_t} < \delta \; \forall t \\
253       & \implies Y_t \in W_{\gamma, t} \; \forall t \\
254       & \implies Y \in C^0(W_\gamma)
255     \end{split}
256   \]
257 
258   Finally, notice that \(H^1(W_\gamma)\) is the inverse image of
259   \(C^0(W_\gamma)\) under the continuous inclusion \(H^1(\gamma^* TM)
260   \longhookrightarrow C^0(\gamma^* TM)\) and is therefore open.
261 \end{proof}
262 
263 Let \(W \subset TM\) be an open neighborhood of the zero section in \(TM\) such
264 that \(\exp\!\restriction_W : W \to \exp(W)\) is invertible -- whose existence
265 follows from the fact that the injectivity radius depends continuously on \(p
266 \in M\).
267 
268 \begin{definition}
269   Given \(\gamma \in {C'}^\infty(I, M)\) let \(W_\gamma, W_{\gamma, t} \subset
270   \gamma^* TM\) be as in lemma~\ref{thm:section-in-open-is-open}, define
271   \[
272     \arraycolsep=1pt
273     \begin{array}{rl}
274       \exp_\gamma : H^1(W_\gamma) & \to H^1(I, M) \\
275       X &
276       \begin{array}[t]{rl}
277         \mapsto \exp \circ X : I & \to M \\
278         t & \mapsto \exp_{\gamma(t)}(X_t)
279       \end{array}
280     \end{array}
281   \]
282   and let \(U_\gamma = \exp_\gamma(H^1(W_\gamma))\).
283 \end{definition}
284 
285 Finally, we find\dots
286 
287 \begin{theorem}
288   Given \(\gamma \in {C'}^\infty(I, M)\), the map \(\exp_\gamma : H^1(W_\gamma)
289   \to U_\gamma\) is bijective. The collection \(\{(U_\gamma, \exp_\gamma^{-1} :
290   U_\gamma \to H^1(\gamma^* TM))\}_{\gamma \in {C'}^\infty(I, M)}\) is an atlas
291   for \(H^1(I, M)\) under the final topology of the maps \(\exp_\gamma\) --
292   i.e. the coarsest topology such that such maps are continuous. This atlas
293   gives \(H^1(I, M)\) the structure of a \emph{separable} Banach manifold
294   modeled after separable Hilbert spaces, with typical
295   representatives\footnote{Any trivialization of $\gamma^* TM$ induces an
296   isomorphism $H^1(\gamma^* TM) \isoto H^1(I \times \mathbb{R}^n) \cong H^1(I,
297   \mathbb{R}^n)$.} \(H^1(\gamma^* TM) \cong H^1(I, \mathbb{R}^n)\).
298 \end{theorem}
299 
300 The fact that \(\exp_\gamma\) is bijective should be clear from the definition
301 of \(U_\gamma\) and \(W_\gamma\). That each \(\exp_\gamma^{-1}\) is a
302 homeomorphism is also clear from the definition of the topology of \(H^1(I,
303 M)\). Moreover, since \({C'}^\infty(I, M)\) is dense, \(\{U_\gamma\}_{\gamma
304 \in {C'}^\infty(I, M)}\) is an open cover of \(H^1(I, M)\). The real difficulty
305 of this proof is showing that the transition maps
306 \[
307   \exp_\eta^{-1} \circ \exp_\gamma :
308   \exp_\gamma^{-1}(U_\gamma \cap U_\eta) \subset H^1(\gamma^* TM)
309   \to H^1(\eta^* TM)
310 \]
311 are diffeomorphisms, as well as showing that \(H^1(I, M)\) is separable. We
312 leave these details as an exercise to the reader -- see theorem 2.3.12 of
313 \cite{klingenberg} for a full proof.
314 
315 It's interesting to note that this construction is functorial. More
316 precisely\dots
317 
318 \begin{theorem}
319   Given finite-dimensional Riemannian manifolds \(M\) and \(N\) and a smooth
320   map \(f : M \to N\), the map
321   \begin{align*}
322     H^1(I, f) : H^1(I, M) & \to     H^1(I, N) \\
323                    \gamma & \mapsto f \circ \gamma
324   \end{align*}
325   is smooth. In addition, \(H^1(I, f \circ g) = H^1(I, f) \circ H^1(I, g)\) and
326   \(H^1(I, \operatorname{id}) = \operatorname{id}\) for any composable smooth
327   maps \(f\) and \(g\). We thus have a functor \(H^1(I, -) : \mathbf{Rmnn} \to
328   \mathbf{BMnfd}\) from the category \(\mathbf{Rmnn}\) of finite-dimensional
329   Riemannian manifolds and smooth maps onto the category \(\mathbf{BMnfd}\) of
330   Banach manifolds and smooth maps.
331 \end{theorem}
332 
333 We would also like to point out that this is a particular case of a more
334 general construction: that of the Banach manifold \(H^1(E)\) of class \(H^1\)
335 sections of a smooth fiber bundle \(E \to I\) -- not necessarily a vector
336 bundle. Our construction of \(H^1(I, M)\) is equivalent to that of the manifold
337 \(H^1(I \times M)\), in the sense that the canonical map
338 \[
339   \arraycolsep=1pt
340   \begin{array}{rl}
341     \tilde{\cdot} : H^1(I, M) \to & H^1(I \times M) \\
342     \gamma \mapsto &
343     \begin{array}[t]{rl}
344       \tilde\gamma : I & \to I \times M \\
345       t & \mapsto (t, \gamma(t))
346     \end{array}
347   \end{array}
348 \]
349 can be easily checked to be a diffeomorphism.
350 
351 The space \(H^1(E)\) is modeled after the Hilbert spaces \(H^1(F)\) of class
352 \(H^1\) sections of open sub-bundles \(F \subset E\) which have the structure
353 of a vector bundle -- the so called \emph{vector bundle neighborhoods of
354 \(E\)}. This construction is highlighted in great detail and generality in the
355 first section of \cite[ch.~11]{palais}, but unfortunately we cannot afford such
356 a diversion in these short notes. Having said that, we are now finally ready to
357 layout the Riemannian structure of \(H^1(I, M)\).
358 
359 \subsection{The Metric of \(H^1(I, M)\)}
360 
361 We begin our discussion of the Riemannian structure of \(H^1(I, M)\) by looking
362 at its tangent bundle. Notice that for each \(\gamma \in {C'}^\infty(I, M)\)
363 the chart \(\exp_\gamma^{-1} : U_\gamma \to H^1(\gamma^* TM)\) induces a
364 canonical isomorphism \(\phi_\gamma = \phi_{\gamma, \gamma} : T_\gamma H^1(I,
365 M) \isoto H^1(\gamma^* TM)\), as described in
366 proposition~\ref{thm:tanget-space-topology}. In fact, these isomorphisms may be
367 extended to a canonical isomorphism of vector bundles, as seen in\dots
368 
369 \begin{lemma}\label{thm:alpha-fiber-bundles-definition}
370   Given \(i = 0, 1\), the collection \(\{(\psi_{i, \gamma}(H^1(W_\gamma) \times
371   H^i(\gamma^* TM)), \psi_{i, \gamma}^{-1})\}_{\gamma \in {C'}^\infty(I,
372   M)}\) with
373   \[
374     \arraycolsep=1pt
375     \begin{array}{rl}
376       \psi_{i, \gamma} : H^1(W_\gamma) \times H^i(\gamma^* TM) \to
377       & \coprod_{\eta \in H^1(I, M)} H^i(\eta^* TM) \\
378       (X, Y) \mapsto &
379       \begin{array}[t]{rl}
380         \psi_{i, \gamma}(X) : I & \to \exp_\gamma(X)^* TM \\
381         t & \mapsto (d \exp)_{X_t} Y_t
382       \end{array}
383     \end{array}
384   \]
385   gives \(\coprod_{\gamma \in {C'}^\infty(I, M)} H^i(\gamma^* TM) \to H^1(I,
386   M)\) the structure of a smooth vector bundle\footnote{Here we use the
387   canonical identification $T_{\gamma(t)} M \cong T_{X_t} TM$ to apply the
388   vector $Y_t \in T_{\gamma(t)} M$ to the map $(d \exp)_{X_t} : T_{X_t} TM \to
389   T_{\exp_{\gamma(t)}(X_t)} M$.}.
390 \end{lemma}
391 
392 \begin{proposition}
393   There is a canonical isomorphism of vector bundles
394   \[
395     T H^1(I, M) \isoto \coprod_{\gamma \in H^1(I, M)} H^1(\gamma^* TM)
396   \]
397   whose restriction \(T_\gamma H^1(I, M) \isoto H^1(\gamma^* TM)\) is given by
398   \(\phi_\gamma\) for all \(\gamma \in {C'}^\infty(I, M)\).
399 \end{proposition}
400 
401 \begin{proof}
402   Note that the sets \(H^1(W_\gamma) \times T_\gamma H^1(I, M)\) are precisely
403   the images of the charts
404   \[
405     \varphi_\gamma^{-1} :
406     \varphi_\gamma(H^1(W_\gamma) \times T_\gamma H^1(I, M))
407     \subset T H^1(I, M)
408     \to H^1(W_\gamma) \times T_\gamma H^1(I, M)
409   \]
410   of \(T H^1(I, M)\) given by\footnote{Once more, we use the
411   canonical identification $T_X H^1(W_\gamma) \cong H^1(\gamma^* TM)$ to apply
412   the vector $\phi_\gamma(Y) \in H^1(\gamma^* TM)$ to $(d \exp_\gamma)_X : T_X
413   H^1(W_\gamma) \to T_{\exp_\gamma(X)} H^1(I, M)$.}
414   \begin{align*}
415     \varphi_\gamma : H^1(W_\gamma) \times T_\gamma H^1(I, M)
416     & \to T H^1(I, M) \\
417     (X, Y) & \mapsto (d \exp_\gamma)_X \phi_\gamma(Y)
418   \end{align*}
419 
420   By composing charts we get a fiber-preserving, fiber-wise linear
421   diffeomorphism
422   \[
423     \varphi_\gamma(H^1(W_\gamma) \times T_\gamma H^1(I, M))
424     \subset T H^1(I, M)
425     \isoto
426     \psi_{1, \gamma}(H^1(W_\gamma) \times H^1(\gamma^* TM)),
427   \]
428   which takes \(\varphi_\gamma(X, Y) \in T_{\exp_\gamma(X)} H^1(I, M)\) to
429   \(\psi_{1, \gamma}(X, \phi_\gamma(Y)) \in H^1(\exp_\gamma(X)^* TM)\). With
430   enough patience, one can deduce from the fact that \(\varphi_\gamma^{-1}\)
431   and \(\psi_{1, \gamma}^{-1}\) are charts that these maps agree in the
432   intersections of the open subsets \(\varphi_\gamma(H^1(W_\gamma) \times
433   T_\gamma H^1(I, M))\), so that they may be glued together into a global
434   smooth map \(\Phi : T H^1(I, M) \to \coprod_{\eta \in H^1(I, M)} H^1(\eta^*
435   TM)\).
436 
437   Since this map is a fiber-preserving, fiber-wise linear local diffeomorphism,
438   this is an isomorphism of vector bundles.
439   Furthermore, by construction
440   \[
441     \Phi(X)_t
442     = \psi_{1, \gamma}(0, \phi_\gamma(X))_t
443     = (d \exp)_{0_{\gamma(t)}} \phi_\gamma(X)_t
444     = \phi_\gamma(X)_t
445   \]
446   for each \(\gamma \in {C'}^\infty(I, M)\) and \(X \in T_\gamma H^1(I, M)\).
447   In other words, \(\Phi\!\restriction_{T_\gamma H^1(I, M)} = \phi_\gamma\) as
448   required.
449 \end{proof}
450 
451 At this point it may be tempting to think that we could now define the metric
452 of \(H^1(I, M)\) in a fiber-wise basis via the identification \(T_\gamma H^1(I,
453 M) \cong H^1(\gamma^* TM)\). In a very real sense this is what we are about to
454 do, but unfortunately there are still technicalities in our way. The issue we
455 face is that proposition~\ref{thm:h0-bundle-is-complete} only applies for
456 \emph{smooth} vector bundles \(E \to I\), which may not be the case for \(E =
457 \gamma^* TM\) if \(\gamma \in H^1(I, M)\) lies outside of \({C'}^\infty(I,
458 M)\). In fact, neither \(\langle X , Y \rangle_0\) nor \(\langle \, ,
459 \rangle_1\) are defined \emph{a priori} for \(X, Y \in H^0(\gamma^* TM)\) with
460 \(\gamma \notin {C'}^\infty(I, M)\).
461 
462 Nevertheless, we can get around this limitation by extending the metric
463 \(\langle \, , \rangle_0\) and the covariant derivative \(\frac\nabla\dt =
464 \nabla_{\frac\dd\dt}\) to
465 those \(H^0(\gamma^* TM)\) with \(\gamma \notin {C'}^\infty(I, M)\). In other
466 words, we'll show\dots
467 
468 \begin{theorem}\label{thm:h0-has-metric-extension}
469   The vector bundle \(\coprod_{\gamma \in H^1(I, M)} H^0(\gamma^* TM) \to
470   H^1(I, M)\) admits a canonical Riemannian metric whose restriction to the
471   fibers \(H^0(\gamma^* TM) = \left.\coprod_{\eta \in H^1(I, M)} H^0(\eta^*
472   TM)\right|_\gamma\) for \(\gamma \in {C'}^\infty(I, M)\) is given by
473   \(\langle \, , \rangle_0\) as defined in
474   proposition~\ref{thm:h0-bundle-is-complete}.
475 \end{theorem}
476 
477 \begin{proof}
478   Given \(\gamma \in {C'}^\infty(I, M)\) and \(X \in H^1(\gamma^* TM)\), let
479   \begin{align*}
480     g_X^\gamma : H^0(\gamma^* TM) \times H^0(\gamma^* TM) & \to \mathbb{R} \\
481     (Y, Z) &
482     \mapsto \int_0^1
483     \langle (d\exp)_{X_t} Y_t, (d\exp)_{X_t} Z_t \rangle \; \dt
484   \end{align*}
485 
486   This is clearly a Riemannian metric in the bundle \(H^1(W_\gamma) \times
487   H^0(\gamma^* TM) \to H^1(W_\gamma)\). Now by composing with the chart
488   \(\psi_{0, \gamma}^{-1}\) as in
489   lemma~\ref{thm:alpha-fiber-bundles-definition} we get a Riemannian metric
490   \(g_\gamma\) in the bundle \(\coprod_{\eta \in U_\gamma} H^0(\eta^* TM) =
491   \left. \coprod_{\eta \in H^1(I, M)} H^0(\eta^* TM) \right|_{U_\gamma} \to
492   U_\gamma\). One can then quickly verify that the \(g^\gamma\)'s agree in the
493   intersection of the \(U_\gamma\)'s, so that they define a global Riemannian
494   metric \(g\) in \(\coprod_{\eta \in H^1(I, M)} H^0(\eta^* TM)\).
495 
496   Furthermore, given \(\gamma \in {C'}^\infty(I, M)\) and \(X, Y \in
497   H^0(\gamma^* TM)\) by construction we have
498   \[
499     g_\gamma(X, Y)
500     = g_0^\gamma(X, Y)
501     = \int_0^1
502       \langle
503       (d \exp)_{0_{\gamma(t)}} X_t, (d \exp)_{0_{\gamma(t)}} Y_t
504       \rangle \; \dt
505     = \int_0^1 \langle X_t, Y_t \rangle \; \dt
506     = \langle X, Y \rangle_0
507   \]
508 \end{proof}
509 
510 \begin{proposition}\label{thm:partial-is-smooth-sec}
511   The map
512   \begin{align*}
513     \partial : H^1(I, M) & \to     \coprod_{\gamma} H^0(\gamma^* TM) \\
514                   \gamma & \mapsto \dot\gamma
515   \end{align*}
516   is a smooth section of \(\coprod_{\gamma \in H^1(I, M)} H^0(\gamma^* TM) \to
517   H^1(I, M)\).
518 \end{proposition}
519 
520 \begin{proposition}\label{thm:covariant-derivative-h0}
521   Denote by \(\nabla^0 : \mathfrak{X}(H^1(I, M)) \times
522   \Gamma\left(\coprod_\gamma H^0(\gamma^* TM)\right) \to
523   \Gamma\left(\coprod_{\gamma} H^0(\gamma^* TM)\right)\) the Levi-Civita
524   connection of \(\coprod_\gamma H^0(\gamma^* TM)\). The map
525   \begin{align*}
526     \mathfrak{X}(H^1(I, M))
527     & \to \Gamma\left(\coprod_\gamma H^0(\gamma^* TM)\right) \\
528     \tilde X
529     & \mapsto \nabla_{\tilde X}^0 \partial
530   \end{align*}
531   is such that
532   \[
533     (\nabla_X^0 \partial)_\gamma
534     = \nabla_{\frac\dd\dt} X
535     = \frac\nabla\dt X
536   \]
537   for all \(\gamma \in {C'}^\infty(I, M)\) and \(X \in H^1(\gamma^* TM) \cong
538   T_\gamma H^1(I, M)\). Given some arbitrary \(\gamma \in H^1(I, M)\) and \(X
539   \in H^1(\gamma^* TM)\) we therefore denote \((\nabla_X^0 \partial)_\gamma\)
540   simply by \(\frac\nabla\dt X\).
541 \end{proposition}
542 
543 The proofs of these last two propositions were deemed too technical to be
544 included in here, but see proposition 2.3.16 and 2.3.18 of \cite{klingenberg}.
545 We may now finally describe the canonical Riemannian metric of \(H^1(I, M)\).
546 
547 \begin{definition}\label{def:h1-metric}
548   Given \(\gamma \in H^1(I, M)\) and \(X, Y \in H^1(\gamma^* TM)\), let
549   \[
550     \langle X, Y \rangle_1
551     = \langle X, Y \rangle_0
552     + \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle_0
553   \]
554 \end{definition}
555 
556 At this point it should be obvious that definition~\ref{def:h1-metric} does
557 indeed endow \(H^1(I, M)\) with the structure of a Riemannian manifold: the
558 inner products \(\langle \, , \rangle_1 : H^1(\gamma^* TM) \times H^1(\gamma^*
559 TM) \to \mathbb{R}\) may be glued together into a single positive-definite
560 section \(\langle \, , \rangle_1 \in \Gamma\left(\operatorname{Sym}^2
561 \coprod_\gamma H^1(\gamma^* TM)\right)\) -- whose smoothness follows from
562 theorem~\ref{thm:h0-has-metric-extension},
563 proposition~\ref{thm:partial-is-smooth-sec} and
564 proposition~\ref{thm:covariant-derivative-h0} -- which is then mapped to a
565 positive-definite section of \(\operatorname{Sym}^2 T H^1(I, M)\) by the
566 induced isomorphism
567 \[
568   \Gamma\left(\operatorname{Sym}^2 \coprod_\gamma H^1(\gamma^* TM)\right)
569   \isoto \Gamma(\operatorname{Sym}^2 T H^1(I, M))
570 \]
571 
572 We are finally ready to discuss some applications.