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
applications.tex (26240B)
1 \section{Applications to the Calculus of Variations}\label{sec:aplications} 2 3 As promised, in this section we will apply our understanding of the structure 4 of \(H^1(I, M)\) to the calculus of variations, and in particular to the 5 geodesics problem. We also describe some further applications, such as the 6 Morse index theorem and the Jacobi-Darboux theorem. We start by defining\dots 7 8 \begin{definition}\label{def:variation} 9 Given \(\gamma \in H^1(I, M)\), a variation \(\{ \gamma_t \}_t\) of 10 \(\gamma\) is a smooth curve \(\gamma_\cdot : (-\epsilon, \epsilon) \to 11 H^1(I, M)\) with \(\gamma_0 = \gamma\). We call the vector 12 \(\left.\frac\dd\dt\right|_{t = 0} \gamma_t \in H^1(\gamma^* TM)\) \emph{the 13 variational vector field of \(\{ \gamma_t \}_t\)}. 14 \end{definition} 15 16 We should note that the previous definition encompasses the classical 17 definition of a variation of a curve, as defined in \cite[ch.~5]{gorodski} for 18 instance: any piece-wise smooth function \(H : I \times (-\epsilon, \epsilon) 19 \to M\) determines a variation \(\{ \gamma_t \}_t\) given by \(\gamma_t(s) = 20 H(s, t)\). This is representative of the theory that lies ahead, in the sense 21 that most of the results we'll discuss in the following are minor refinements 22 of the classical theory. Instead, the value of the theory we will develop in 23 here lies in its conceptual simplicity: instead of relying in ad-hoc methods we 24 can now use the standard tools of calculus to study the critical points of the 25 energy functional \(E\). 26 27 What we mean by this last statement is that by look at the energy functional as 28 a smooth function \(E \in C^\infty(H^1(I, M))\) we can study its classical 29 ``critical points'' -- i.e. curves \(\gamma\) with a variation \(\{ \gamma_t 30 \}_t\) such that \(\left.\frac\dd\dt\right|_{t = 0} E(\gamma_t) = 0\) -- by 31 looking at its derivative. The first variation of energy thus becomes a 32 particular case of a formula for \(d E\), and the second variation of energy 33 becomes a particular case of a formula for the Hessian of \(E\) at a critical 34 point. Without further ado, we prove\dots 35 36 \begin{theorem}\label{thm:energy-is-smooth} 37 The energy functional 38 \begin{align*} 39 E : H^1(I, M) & \to \mathbb{R} \\ 40 \gamma 41 & \mapsto \frac{1}{2} \norm{\partial \gamma}_0^2 42 = \frac{1}{2} \int_0^1 \norm{\dot\gamma(t)}^2 \; \dt 43 \end{align*} 44 is smooth and \(d E_\gamma X = \left\langle \partial \gamma, \frac\nabla\dt X 45 \right\rangle_0\). 46 \end{theorem} 47 48 \begin{proof} 49 The fact that \(E\) is smooth should be clear from the smoothness of 50 \(\partial\) and \(\norm\cdot_0\). Furthermore, from the definition of 51 \(\frac\nabla\dt\) we have 52 \[ 53 \begin{split} 54 \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 55 & = 56 \left\langle 57 \partial \gamma, (\nabla_X^0 \partial)_\gamma 58 \right\rangle_0 \\ 59 & = (\tilde X \langle \partial, \partial \rangle_0)(\gamma) - 60 \left\langle 61 (\nabla_X^0 \partial)_\gamma, \partial \gamma 62 \right\rangle_0 \\ 63 & = 2 \tilde X E(\gamma) - 64 \left\langle \frac\nabla\dt X, \partial \gamma \right\rangle_0 \\ 65 & = 2 d E_\gamma X - 66 \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 67 \end{split} 68 \] 69 where \(\tilde X \in \mathfrak{X}(H^1(I, M))\) is any vector field with 70 \(\tilde X_\gamma = X\). 71 \end{proof} 72 73 As promised, by applying the chain rule and using the compatibility of 74 \(\nabla\) with the metric we arrive at the classical formula for the first 75 variation of energy \(E\). 76 77 \begin{corollary} 78 Given a piece-wise smooth curve \(\gamma : I \to M\) with 79 \(\gamma\!\restriction_{[t_i, t_{i + 1}]}\) smooth and a variation \(\{ 80 \gamma_t \}_t\) of \(\gamma\) with variational vector field \(X\) we have 81 \[ 82 \left.\frac\dd\dt\right|_{t = 0} E(\gamma_t) 83 = \sum_i 84 \left. \langle \dot\gamma(t), X_t \rangle\right|_{t = t_i}^{t_{i + 1}} 85 - \int_0^1 \left\langle \frac\nabla\dt \dot\gamma(t), X_t \right\rangle 86 \; \dt 87 \] 88 \end{corollary} 89 90 Another interesting consequence of theorem~\ref{thm:energy-is-smooth} is\dots 91 92 \begin{corollary} 93 The only critical points of \(E\) in \(H^1(I, M)\) are the constant curves. 94 \end{corollary} 95 96 \begin{proof} 97 Clearly every constant curve is a critical point. On the other hand, if 98 \(\gamma \in H^1(I, M)\) is such that \(\left\langle \partial \gamma, 99 \frac\nabla\dt X \right\rangle_0 = d E_\gamma X = 0\) for all \(X \in 100 H^1(\gamma^* TM)\) then \(\partial \gamma = 0\) and therefore \(\gamma\) is 101 constant. 102 \end{proof} 103 104 Another way to put is to say that the problem of characterizing the critical 105 points of \(E\) in \(H^1(I, M)\) is not interesting at all. This shouldn't 106 really come as a surprise, as most interesting results from the classical 107 theory are concerned with particular classes of variations of a curves, such as 108 variations with fixed endpoints or variations through loops. In the next 109 section we introduce two submanifolds of \(H^1(I, M)\), corresponding to the 110 classes of variations previously described, and classify the critical points of 111 the restrictions of \(E\) to such submanifolds. 112 113 \subsection{The Critical Points of \(E\)} 114 115 We begin with a technical lemma. 116 117 \begin{lemma} 118 The maps \(\sigma, \tau: H^1(I, M) \to M\) with \(\sigma(\gamma) = 119 \gamma(0)\) and \(\tau(\gamma) = \gamma(1)\) are submersions. 120 \end{lemma} 121 122 \begin{proof} 123 To see that \(\sigma\) and \(\tau\) are smooth it suffices to observe that 124 their local representation in \(U_\gamma\) for \(\gamma \in {C'}^\infty(I, 125 M)\) is given by the maps 126 \begin{align*} 127 U \subset H^1(W_\gamma) & \to T_{\gamma(0)} M & 128 U \subset H^1(W_\gamma) & \to T_{\gamma(1)} M \\ 129 X & \mapsto X_0 & 130 X & \mapsto X_1 131 \end{align*} 132 which are indeed smooth functions. This local representation also shows that 133 \begin{align*} 134 d\sigma_\gamma : H^1(\gamma^* TM) & \to T_{\gamma(0)} M & 135 d\tau_\gamma : H^1(\gamma^* TM) & \to T_{\gamma(1)} M \\ 136 X & \mapsto X_0 & 137 X & \mapsto X_1 138 \end{align*} 139 are surjective maps for all \(\gamma \in H^1(I, M)\). 140 \end{proof} 141 142 We can now show\dots 143 144 \begin{theorem} 145 The subspace \(\Omega_{p q} M \subset H^1(I, M)\) of curves joining \(p, q 146 \in M\) is a submanifold whose tangent space \(T_\gamma \Omega_{p q} M\) is 147 the subspace of \(H^1(\gamma^* TM)\) consisting of class \(H^1\) vector 148 fields \(X\) along \(\gamma\) with \(X_0 = X_1 = 0\). Likewise, the space 149 \(\Lambda M \subset H^1(I, M)\) of free loops is a submanifold whose tangent 150 at \(\gamma\) is given by all \(X \in H^1(\gamma^* TM)\) with \(X_0 = X_1\). 151 \end{theorem} 152 153 \begin{proof} 154 To see that these are submanifolds, it suffices to note that \(\Omega_{p q} 155 M\) and \(\Lambda M\) are the inverse images of the closed submanifolds 156 \(\{(p, q)\}, \{(p, p) : p \in M\} \subset M \times M\) under the submersion 157 \((\sigma, \tau) : H^1(I, M) \to M \times M\). 158 159 The characterization of their tangent bundles should also be clear: any curve 160 \((-\epsilon, \epsilon) \to H^1(I, M)\) passing through \(\gamma \in 161 \Omega_{p q} M\) whose image is contained in \(\Omega_{p q} M\) is a 162 variation of \(\gamma\) with fixed endpoints, so its variational vector field 163 \(X\) satisfies \(X_0 = X_1 = 0\). Likewise, any variation of a loop \(\gamma 164 \in \Lambda M\) trough loops -- i.e. a curve \((-\epsilon, \epsilon) \to 165 \Lambda M\) passing through \(\gamma\) -- satisfies \(X_0 = X_1\). 166 \end{proof} 167 168 Finally, as promised we will provide a characterization of the critical points 169 of \(E\!\restriction_{\Omega_{p q} M}\) and \(E\!\restriction_{\Lambda M}\). 170 171 \begin{theorem}\label{thm:critical-points-char-in-submanifolds} 172 The critical points of \(E\!\restriction_{\Omega_{p q} M}\) are precisely the 173 geodesics of \(M\) joining \(p\) and \(q\). The critical points of 174 \(E\!\restriction_{\Lambda M}\) are the closed geodesics of \(M\) -- 175 including the constant maps. 176 \end{theorem} 177 178 \begin{proof} 179 We start by supposing that \(\gamma\) is a geodesic. Since \(\gamma\) is 180 smooth, 181 \[ 182 d E_\gamma X 183 = \int_0^1 184 \left\langle \dot\gamma(t), \frac\nabla\dt X \right\rangle \; \dt 185 = \int_0^1 186 \frac\dd\dt \langle \dot\gamma(t), X_t \rangle - 187 \left\langle \frac\nabla\dt \dot\gamma(t), X \right\rangle \; \dt 188 = \langle \dot\gamma(1), X_1 \rangle - \langle \dot\gamma(0), X_0 \rangle 189 \] 190 191 Now if \(\gamma \in \Omega_{p q} M\) and \(X \in T_\gamma \Omega_{p q} M\) 192 then \(d E_\gamma X = \langle \dot\gamma(1), 0 \rangle - \langle 193 \dot\gamma(0), 0 \rangle = 0\). Likewise, if \(\gamma\) is a closed geodesic 194 and \(X \in T_\gamma \Lambda M\) we find \(d E_\gamma X = 0\) since 195 \(\dot\gamma(0) = \dot\gamma(1)\) and \(X_0 = X_1\). This establishes that 196 the geodesics are indeed critical points of the restrictions of \(E\). 197 198 Suppose \(\gamma \in \Omega_{p q} M\) is a critical point and let \(Y, Z \in 199 H^1(\gamma^* TM)\) be such that 200 \begin{align*} 201 \frac\nabla\dt Y & = \partial \gamma & 202 Y_0 & = 0 & 203 \frac\nabla\dt Z & = 0 & 204 Z_1 & = Y_1 205 \end{align*} 206 207 Let \(X_t = Y_t - t Z_t\). Then \(X_0 = X_1 = 0\) and \(\frac\nabla\dt X = 208 \partial \gamma - Z\). Furthermore, 209 \[ 210 \langle Z, \partial \gamma - Z \rangle_0 211 = \left\langle Z, \frac\nabla\dt X \right\rangle_0 212 = \int_0^1 \frac\dd\dt \langle Z_t, X_t \rangle \; \dt 213 = \langle Z_1, X_1 \rangle - \langle Z_0, X_0 \rangle 214 = 0 215 \] 216 and 217 \[ 218 \langle \partial \gamma, \partial \gamma - Z \rangle_0 219 = \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 220 = d E_\gamma X 221 = 0, 222 \] 223 which implies \(\norm{\partial \gamma - Z}_0^2 = 0\). In other words, 224 \(\partial \gamma = Z \in H^1(\gamma^* TM)\) and therefore \(\frac\nabla\dt 225 \dot\gamma(t) = \frac\nabla\dt Z = 0\) -- i.e. \(\gamma\) is a geodesic. 226 227 Finally, if \(\gamma \in \Lambda M\) with \(\gamma(0) = \gamma(1) = p\) we 228 may apply the argument above to conclude that \(\gamma\) is a geodesic 229 joining \(p\) to \(q = p\). To see that \(\gamma\) is a closed geodesic apply 230 the same argument again for \(\eta(t) = \gamma(1 + \sfrac{1}{2})\) to 231 conclude that \(\dot\gamma(0) = \dot\eta(\sfrac{1}{2}) = \dot\gamma(1)\). 232 \end{proof} 233 234 We should point out that the first part of 235 theorem~\ref{thm:critical-points-char-in-submanifolds} is a particular case of 236 a result regarding critical points of the restriction of \(E\) to the 237 submanifold \(H_{N_0, N_1}^1(I, M) \subset H^1(I, M)\) of curves joining 238 submanifolds \(N_0, N_1 \subset M\): the critical points of 239 \(E\!\restriction_{H_{N_0, N_1}^1(I, M)}\) are the geodesics \(\gamma\) joining 240 \(N_0\) to \(N_1\) with \(\dot\gamma(0) \in T_{\gamma(0)} N_0^\perp\) and 241 \(\dot\gamma(1) \in T_{\gamma(1)} N_1^\perp\). The proof of this result is 242 essentially the same as that of 243 theorem~\ref{thm:critical-points-char-in-submanifolds}, given that \(T_\gamma 244 H_{N_0, N_1}^1(I, M)\) is subspace of \(H^1\) vector fields \(X\) along 245 \(\gamma\) with \(X_0 \in T_{\gamma(0)} N_0\) and \(X_1 \in T_{\gamma(1)} 246 N_1\). 247 248 \subsection{Second Order Derivatives of \(E\)} 249 250 Having establish a clear connection between geodesics and critical points of 251 \(E\), the only thing we're missing to complete our goal of providing a modern 252 account of the classical theory is a refurnishing of the formula for second 253 variation of energy. Intuitively speaking, the second variation of energy 254 should be a particular case of a formula for the second derivative of \(E\). 255 The issue we face is, of course, that in general there is no such thing as 256 ``the second derivative'' of a smooth function between manifolds. 257 258 Nevertheless, the metric of \(H^1(I, M)\) allow us to discuss ``the second 259 derivative'' of \(E\) in a meaningful sense by looking at the Hessian form, 260 which we define in the following. 261 262 \begin{definition} 263 Given a -- possibly infinite-dimensional -- Riemannian manifold \(N\) and a 264 smooth functional \(f : N \to \mathbb{R}\), we call the symmetric tensor 265 \[ 266 d^2 f(X, Y) = \nabla d f (X, Y) = X Y f - df \nabla_X Y 267 \] 268 \emph{the Hessian of \(f\)}. 269 \end{definition} 270 271 We can now apply the classical formula for the second variation of energy to 272 compute the Hessian of \(E\) at a critical point. 273 274 \begin{theorem} 275 If \(\gamma\) is a critical point of \(E\!\restriction_{\Omega_{p q} M}\) 276 then 277 \begin{equation}\label{eq:second-variation-general} 278 (d^2 E\!\restriction_{\Omega_{p q} M})_\gamma(X, Y) 279 = \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle_0 280 - \langle R_\gamma X, Y \rangle_0, 281 \end{equation} 282 where \(R_\gamma : H^1(\gamma^* TM) \to H^1(\gamma^* TM)\) is given by 283 \((R_\gamma X)_t = R(X_t, \dot\gamma(t)) \dot\gamma(t)\). Formula 284 (\ref{eq:second-variation-general}) also holds for critical points of \(E\) 285 in \(\Lambda M\). 286 \end{theorem} 287 288 \begin{proof} 289 Given the symmetry of \(d^2 E\), it suffices to take \(X \in T_\gamma 290 \Omega_{p q} M\) and show 291 \[ 292 d^2 E_\gamma(X, X) 293 = \norm{\frac\nabla\dt X}_0^2 - \langle R_\gamma X, X \rangle_0 294 \] 295 296 To that end, we fix a variation \(\{ \gamma_t \}_t\) of \(\gamma\) with fixed 297 endpoints and variational field \(X\) and compute 298 \[ 299 \begin{split} 300 d^2 E_\gamma(X, X) 301 & = \left.\frac{\dd^2}{\dt^2}\right|_{t = 0} E(\gamma_t) \\ 302 \text{(second variation of energy)} 303 & = \int_0^1 \norm{\frac\nabla\dt X}^2 304 - \langle R(X_t, \dot\gamma(t)) \dot\gamma(t), X_t \rangle \; \dt \\ 305 & = \norm{\frac\nabla\dt X}_0^2 - \langle R_\gamma X, X \rangle_0 306 \end{split} 307 \] 308 \end{proof} 309 310 Next we discuss some further applications of the theory we've developed so far. 311 In particular, we will work towards Morse's index theorem and and describe how 312 one can apply it to establish the Jacobi-Darboux theorem. We begin with a 313 technical lemma, whose proof amounts to an uninspiring exercise in analysis -- 314 see lemma 2.4.6 of \cite{klingenberg}. 315 316 \begin{lemma}\label{thm:inclusion-submnfds-is-compact} 317 Let \(\Omega_{p q}^0 M \subset C^0(I, M)\) be the space of continuous curves 318 joining \(p\) to \(q\). Then the inclusion \(\Omega_{p q} M 319 \longhookrightarrow \Omega_{p q}^0 M\) is continuous and compact. Likewise, 320 if \(M\) is compact and \(\Lambda^0 M \subset C^0(I, M)\) is the space of 321 continuous free loops then the inclusion \(\Lambda M \longhookrightarrow 322 \Lambda^0 M\) is continuous and compact. 323 \end{lemma} 324 325 As a first consequence, we prove\dots 326 327 \begin{proposition}\label{thm:morse-index-e-is-finite} 328 Given a critical point \(\gamma\) of \(E\) in \(\Omega_{p q} M\), the 329 self-adjoint operator \(A_\gamma : T_\gamma \Omega_{p q} M \to T_\gamma 330 \Omega_{p q} M\) given by 331 \[ 332 \langle A_\gamma X, Y \rangle_1 333 = \langle X, A_\gamma Y \rangle_1 334 = d^2 E_\gamma(X, Y) 335 \] 336 has the form \(A_\gamma = \operatorname{Id} + K_\gamma\) where \(K_\gamma : 337 T_\gamma \Omega_{p q} M \to T_\gamma \Omega_{p q} M\) is a compact operator. 338 The same holds for \(\Lambda M\) if \(M\) is compact. 339 \end{proposition} 340 341 \begin{proof} 342 Consider \(K_\gamma = - \left( \operatorname{Id} - \frac{\nabla^2}{\dt^2} 343 \right)^{-1} \circ (\operatorname{Id} + R_\gamma)\). We will show that 344 \(K_\gamma\) is compact and that \(A_\gamma = \operatorname{Id} + K_\gamma\) 345 for \(\gamma\) in both \(\Omega_{p q} M\) and \(\Lambda M\) -- in which case 346 assume \(M\) is compact. 347 348 Let \(\gamma \in \Omega_{p q} M\) be a critical point. By 349 theorem~\ref{thm:critical-points-char-in-submanifolds} we know that 350 \(\gamma\) is a geodesic. Let \(X, Y \in \Gamma(\gamma^* TM)\) with \(X_0 = 351 X_1 = Y_0 = Y_1 = 0\). Then 352 \begin{equation}\label{eq:compact-partial-result} 353 \begin{split} 354 \langle X, Y \rangle_1 355 & = \langle X, Y \rangle_0 356 + \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle \\ 357 & = \langle X, Y \rangle_0 358 + \int_0^1 359 \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle 360 \; \dt \\ 361 & = \langle X, Y \rangle_0 362 + \int_0^1 363 \frac\dd\dt \langle X_t, Y_t \rangle - 364 \left\langle \frac{\nabla^2}{\dt^2} X, Y \right\rangle \; \dt \\ 365 & = \langle X, Y \rangle_0 366 - \left\langle \frac{\nabla^2}{\dt^2} X, Y \right\rangle_0 367 + \left.\langle X_t, Y_t \rangle\right|_{t = 0}^1 \\ 368 & = \left\langle 369 \left(\operatorname{Id} - \frac{\nabla^2}{\dt^2}\right) X, Y 370 \right\rangle_0 371 \end{split} 372 \end{equation} 373 374 Since \(\Gamma(\gamma^* TM) \subset H^1(\gamma^* TM)\) is dense, 375 (\ref{eq:compact-partial-result}) extends to all of \(T_\gamma \Omega_{p q} 376 M\). Hence given \(X, Y \in T_\gamma \Omega_{p q} M\) we have 377 \[ 378 \begin{split} 379 \langle A_\gamma X, Y \rangle_1 380 & = \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle_0 381 - \langle R_\gamma X, Y \rangle_0 \\ 382 & = \langle X, Y \rangle_1 - \langle X, Y \rangle_0 383 - \langle R_\gamma X, Y \rangle_0 \\ 384 & = \langle X, Y \rangle_1 385 - \langle (\operatorname{Id} + R_\gamma) X, Y \rangle_0 \\ 386 & = \langle X, Y \rangle_1 387 - \left\langle 388 \left( \operatorname{Id} - \frac{\nabla^2}{\dt^2} \right)^{-1} 389 \circ (\operatorname{Id} + R_\gamma) X, Y 390 \right\rangle_1 \\ 391 & = \langle X, Y \rangle_1 + \langle K_\gamma X, Y \rangle_1 \\ 392 \end{split} 393 \] 394 395 Now consider a critical point \(\gamma \in \Lambda M\) -- i.e. a closed 396 geodesic. Equation (\ref{eq:compact-partial-result}) also holds for \(X, Y 397 \in \Gamma(\gamma^* TM)\) with \(X_0 = X_1\) and \(Y_0 = Y_1\), so it holds 398 for all \(X, Y \in T_\gamma \Lambda M\). Hence by applying the same argument 399 we get \(\langle A_\gamma X, Y \rangle_1 = \langle (\operatorname{Id} + 400 K_\gamma) X, Y \rangle_1\). 401 402 As for the compactness of \(K_\gamma\) in the case of \(\Omega_{p q} M\), 403 from (\ref{eq:compact-partial-result}) we get \(\norm{K_\gamma X}_1^2 = - 404 \langle (\operatorname{Id} + R_\gamma) X, K_\gamma X \rangle_0\), so that 405 proposition~\ref{thm:continuous-inclusions-sections} implies 406 \begin{equation}\label{eq:compact-operator-quota} 407 \norm{K_\gamma X}_1^2 408 \le \norm{\operatorname{Id} + R_\gamma} 409 \cdot \norm{K_\gamma X}_\infty \cdot \norm{X}_0 410 \le \sqrt{2} \norm{\operatorname{Id} + R_\gamma} \cdot \norm{K_\gamma X}_1 411 \cdot \norm{X}_0 412 \end{equation} 413 414 Given a bounded sequence \((X_n)_n \subset T_\gamma \Omega_{p q} M\), it 415 follow from lemma~\ref{thm:inclusion-submnfds-is-compact} that \((X_n)_n\) is 416 relatively compact as a \(C^0\)-sequence. From 417 (\ref{eq:compact-operator-quota}) we then get that \((K_\gamma X_n)_n\) is 418 relatively compact as an \(H^1\)-sequence, as desired. The same argument 419 holds for \(\Lambda M\) if \(M\) is compact -- so that we can once more apply 420 lemma~\ref{thm:inclusion-submnfds-is-compact}. 421 \end{proof} 422 423 Once again, the first part of this proposition is a particular case of a 424 broader result regarding the space of curves joining submanifolds of \(M\): if 425 \(N \subset M\) is a totally geodesic manifold of codimension \(1\) and 426 \(\gamma \in H_{N, \{q\}}^1(I, M)\) is a critical point of the restriction of 427 \(E\) then \(A_\gamma = \operatorname{Id} + K_\gamma\). These results aren't 428 that appealing on their own, but they allow us to establish the following 429 result, which is essential for stating Morse's index theorem. 430 431 \begin{corollary} 432 Given a critical point \(\gamma\) of \(E\!\restriction_{\Omega_{p q} M}\), 433 there is an orthogonal decomposition 434 \[ 435 T_\gamma \Omega_{p q} M 436 = T_\gamma^- \Omega_{p q} M 437 \oplus T_\gamma^0 \Omega_{p q} M 438 \oplus T_\gamma^+ \Omega_{p q} M, 439 \] 440 where \(T_\gamma^- \Omega_{p q} M\) is the finite-dimensional subspace 441 spanned by eigenvectors with negative eigenvalues, \(T_\gamma^0 \Omega_{p q} 442 M = \ker A_\gamma\) and \(T_\gamma^+ \Omega_{p q} M\) is the proper Hilbert 443 subspace given by the closure of the subspace spanned by eigenvectors with 444 positive eigenvalues. The same holds for critical points \(\gamma\) of 445 \(E\!\restriction_{\Lambda M}\) and \(T_\gamma \Lambda M\) if \(M\) is 446 compact. 447 \end{corollary} 448 449 \begin{definition}\label{def:morse-index} 450 Given a critical point \(\gamma\) of \(E\!\restriction_{\Omega_{p q} M}\) we 451 call the number \(\dim T_\gamma^- \Omega_{p q} M\) \emph{the \(\Omega\)-index 452 of \(\gamma\)}. Likewise, we call \(\dim T_\gamma^- \Lambda M\) for a 453 critical point \(\gamma\) of \(E\!\restriction_{\Lambda M}\) \emph{the 454 \(\Lambda\)-index of \(\gamma\)}. Whenever the submanifold \(\gamma\) lies in 455 is clear from context we refer to the \(\Omega\)-index or the 456 \(\Lambda\)-index of \(\gamma\) simply by \emph{the index of \(\gamma\)}. 457 \end{definition} 458 459 This definition highlights one of the greatest strengths of our approach: while 460 the index of a geodesic \(\gamma\) can be defined without the aid of the tools 461 developed in here, by using of the Hessian form \(d^2 E_\gamma\) we can place 462 definition~\ref{def:morse-index} in the broader context of Morse theory. In 463 fact, the geodesics problems and the energy functional where among Morse's 464 original proposed applications. Proposition~\ref{thm:morse-index-e-is-finite} 465 and definition~\ref{def:morse-index} amount to a proof that the Morse index of 466 \(E\) at a critical point \(\gamma\) is finite. 467 468 We are now ready to state Morse's index theorem. 469 470 \begin{theorem}[Morse] 471 Let \(\gamma \in \Omega_{p q} M\) be a critical point of \(E\). Then the 472 index of \(\gamma\) is given of the sum of the multiplicities of the proper 473 conjugate points\footnote{By ``conjugate points of a geodesic $\gamma$'' we 474 of course mean points conjugate to $\gamma(0) = p$ along $\gamma$.} of 475 \(\gamma\) in the interior of \(I\). 476 \end{theorem} 477 478 Unfortunately we do not have the space to include the proof of Morse's theorem 479 in here, but see theorem 2.5.9 of \cite{klingenberg}. The index theorem can be 480 generalized for \(H_{N, \{q\}}^1(I, M)\) by replacing the notion of conjugate 481 point with the notion of focal points of \(N\) -- see theorem 7.5.4 of 482 \cite{gorodski} for the classical approach. What we are really interested in, 483 however, is the following consequence of Morse's theorem. 484 485 \begin{theorem}[Jacobi-Darboux]\label{thm:jacobi-darboux} 486 Let \(\gamma \in \Omega_{p q} M\) be a critical point of \(E\). 487 \begin{enumerate} 488 \item If there are no conjugate points of \(\gamma\) then there exists a 489 neighborhood \(U \subset \Omega_{p q} M\) of \(\gamma\) such that 490 \(E(\eta) > E(\gamma)\) for all \(\eta \in U\) with \(\eta \ne \gamma\). 491 492 \item Let \(k > 0\) be the sum of the multiplicities of the conjugate 493 points of \(\gamma\) in the interior of \(I\). Then there exists an 494 immersion 495 \[ 496 i : B^k \to \Omega_{p q} M 497 \] 498 of the unit ball \(B^k = \{v \in \mathbb{R}^k : \norm{v} < 1\}\) with 499 \(i(0) = \gamma\), \(E(i(v)) < E(\gamma)\) and \(L(i(v)) < L(\gamma)\) 500 for all nonzero \(v \in B^k\). 501 \end{enumerate} 502 \end{theorem} 503 504 \begin{proof} 505 First of all notice that given \(\eta \in U_\gamma\) with \(\eta = 506 \exp_\gamma(X)\), \(X \in H^1(W_\gamma)\), the Taylor series for \(E(\eta)\) 507 is given by \(E(\eta) = E(\gamma) + \frac{1}{2} d^2 E_\gamma(X, X) + 508 \cdots\). More precisely, 509 \begin{equation}\label{eq:energy-taylor-series} 510 \frac 511 {\abs{E(\exp_\gamma(X)) - E(\gamma) - \frac{1}{2} d^2 E_\gamma(X, X)}} 512 {\norm{X}_1^2} 513 \to 0 514 \end{equation} 515 as \(X \to 0\). 516 517 Let \(\gamma\) be as in \textbf{(i)}. Since \(\gamma\) has no conjugate 518 points, it follows from Morse's index theorem that \(T_\gamma^- \Omega_{p q} 519 M = 0\). Furthermore, by noticing that any piece-wise smooth \(X \in \ker 520 A_\gamma\) is Jacobi field vanishing at \(p\) and \(q\) one can also show 521 \(T_\gamma^0 \Omega_{p q} M = 0\). Hence \(T_\gamma \Omega_{p q} M = 522 T_\gamma^+ \Omega_{p q} M\) and therefore \(d^2 E\!\restriction_{\Omega_{p q} 523 M}\) is positive-definite. Now given \(\eta = \exp_\gamma(X)\) as before, 524 (\ref{eq:energy-taylor-series}) implies that \(E(\eta) > E(\gamma)\), 525 provided \(\norm{X}_1\) is sufficiently small. 526 527 As for part \textbf{(ii)}, fix an orthonormal basis 528 \(\{X_j : 1 \le j \le k\}\) of \(T_\gamma^- \Omega_{p q} M\) consisting of 529 eigenvectors of \(A_\gamma\) with negative eigenvalues \(- \lambda_i\). Let 530 \(\delta > 0\) and define 531 \begin{align*} 532 i : B^k & \to \Omega_{p q} M \\ 533 v & \mapsto \exp_\gamma(\delta (v_1 \cdot X_1 + \cdots + v_k \cdot X_k)) 534 \end{align*} 535 536 Clearly \(i\) is an immersion for small enough \(\delta\). Moreover, from 537 (\ref{eq:energy-taylor-series}) and 538 \[ 539 E(i(v)) 540 = E(\gamma) - \frac{1}{2} \delta^2 \sum_j \lambda_j \cdot v_j + \cdots 541 \] 542 we find that \(E(i(v)) < E(\gamma)\) for sufficiently small \(\delta\). In 543 particular, \(L(i(v))^2 \le E(i(v)) < E(\gamma) = L(\gamma)^2\). 544 \end{proof} 545 546 We should point out that part \textbf{(i)} of theorem~\ref{thm:jacobi-darboux} 547 is weaker than the classical formulation of the Jacobi-Darboux theorem -- such 548 as in theorem 5.5.3 of \cite{gorodski} for example -- in two aspects. First, we 549 do not compare the length of curves \(\gamma\) and \(\eta \in U\). This could 550 be amended by showing that the length functional \(L : H^1(I, M) \to 551 \mathbb{R}\) is smooth and that its Hessian \(d^2 L_\gamma\) is given by \(C 552 \cdot d^2 E_\gamma\) for some \(C > 0\). Secondly, unlike the classical 553 formulation we only consider curves in an \(H^1\)-neighborhood of \(\gamma\) -- 554 instead of a neighborhood of \(\gamma\) in \(\Omega_{p q} M\) in the uniform 555 topology. On the other hand, part \textbf{(ii)} is definitively an improvement 556 of the classical formulation: we can find curves \(\eta = i(v)\) shorter than 557 \(\gamma\) already in an \(H^1\)-neighborhood of \(\gamma\). 558 559 This concludes our discussion of the applications of our theory to the 560 geodesics problem. We hope that these short notes could provide the reader with 561 a glimpse of the rich theory of the calculus of variations and global analysis 562 at large. We once again refer the reader to \cite[ch.~2]{klingenberg}, 563 \cite[ch.~11]{palais} and \cite[sec.~6]{eells} for further insight on modern 564 variational methods.