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
Name | Size | Mode | |
.. | |||
sections/applications.tex | 26240B | -rw-r--r-- |
001 002 003 004 005 006 007 008 009 010 011 012 013 014 015 016 017 018 019 020 021 022 023 024 025 026 027 028 029 030 031 032 033 034 035 036 037 038 039 040 041 042 043 044 045 046 047 048 049 050 051 052 053 054 055 056 057 058 059 060 061 062 063 064 065 066 067 068 069 070 071 072 073 074 075 076 077 078 079 080 081 082 083 084 085 086 087 088 089 090 091 092 093 094 095 096 097 098 099 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564
\section{Applications to the Calculus of Variations}\label{sec:aplications} As promised, in this section we will apply our understanding of the structure of \(H^1(I, M)\) to the calculus of variations, and in particular to the geodesics problem. We also describe some further applications, such as the Morse index theorem and the Jacobi-Darboux theorem. We start by defining\dots \begin{definition}\label{def:variation} Given \(\gamma \in H^1(I, M)\), a variation \(\{ \gamma_t \}_t\) of \(\gamma\) is a smooth curve \(\gamma_\cdot : (-\epsilon, \epsilon) \to H^1(I, M)\) with \(\gamma_0 = \gamma\). We call the vector \(\left.\frac\dd\dt\right|_{t = 0} \gamma_t \in H^1(\gamma^* TM)\) \emph{the variational vector field of \(\{ \gamma_t \}_t\)}. \end{definition} We should note that the previous definition encompasses the classical definition of a variation of a curve, as defined in \cite[ch.~5]{gorodski} for instance: any piece-wise smooth function \(H : I \times (-\epsilon, \epsilon) \to M\) determines a variation \(\{ \gamma_t \}_t\) given by \(\gamma_t(s) = H(s, t)\). This is representative of the theory that lies ahead, in the sense that most of the results we'll discuss in the following are minor refinements of the classical theory. Instead, the value of the theory we will develop in here lies in its conceptual simplicity: instead of relying in ad-hoc methods we can now use the standard tools of calculus to study the critical points of the energy functional \(E\). What we mean by this last statement is that by look at the energy functional as a smooth function \(E \in C^\infty(H^1(I, M))\) we can study its classical ``critical points'' -- i.e. curves \(\gamma\) with a variation \(\{ \gamma_t \}_t\) such that \(\left.\frac\dd\dt\right|_{t = 0} E(\gamma_t) = 0\) -- by looking at its derivative. The first variation of energy thus becomes a particular case of a formula for \(d E\), and the second variation of energy becomes a particular case of a formula for the Hessian of \(E\) at a critical point. Without further ado, we prove\dots \begin{theorem}\label{thm:energy-is-smooth} The energy functional \begin{align*} E : H^1(I, M) & \to \mathbb{R} \\ \gamma & \mapsto \frac{1}{2} \norm{\partial \gamma}_0^2 = \frac{1}{2} \int_0^1 \norm{\dot\gamma(t)}^2 \; \dt \end{align*} is smooth and \(d E_\gamma X = \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0\). \end{theorem} \begin{proof} The fact that \(E\) is smooth should be clear from the smoothness of \(\partial\) and \(\norm\cdot_0\). Furthermore, from the definition of \(\frac\nabla\dt\) we have \[ \begin{split} \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 & = \left\langle \partial \gamma, (\nabla_X^0 \partial)_\gamma \right\rangle_0 \\ & = (\tilde X \langle \partial, \partial \rangle_0)(\gamma) - \left\langle (\nabla_X^0 \partial)_\gamma, \partial \gamma \right\rangle_0 \\ & = 2 \tilde X E(\gamma) - \left\langle \frac\nabla\dt X, \partial \gamma \right\rangle_0 \\ & = 2 d E_\gamma X - \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 \end{split} \] where \(\tilde X \in \mathfrak{X}(H^1(I, M))\) is any vector field with \(\tilde X_\gamma = X\). \end{proof} As promised, by applying the chain rule and using the compatibility of \(\nabla\) with the metric we arrive at the classical formula for the first variation of energy \(E\). \begin{corollary} Given a piece-wise smooth curve \(\gamma : I \to M\) with \(\gamma\!\restriction_{[t_i, t_{i + 1}]}\) smooth and a variation \(\{ \gamma_t \}_t\) of \(\gamma\) with variational vector field \(X\) we have \[ \left.\frac\dd\dt\right|_{t = 0} E(\gamma_t) = \sum_i \left. \langle \dot\gamma(t), X_t \rangle\right|_{t = t_i}^{t_{i + 1}} - \int_0^1 \left\langle \frac\nabla\dt \dot\gamma(t), X_t \right\rangle \; \dt \] \end{corollary} Another interesting consequence of theorem~\ref{thm:energy-is-smooth} is\dots \begin{corollary} The only critical points of \(E\) in \(H^1(I, M)\) are the constant curves. \end{corollary} \begin{proof} Clearly every constant curve is a critical point. On the other hand, if \(\gamma \in H^1(I, M)\) is such that \(\left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 = d E_\gamma X = 0\) for all \(X \in H^1(\gamma^* TM)\) then \(\partial \gamma = 0\) and therefore \(\gamma\) is constant. \end{proof} Another way to put is to say that the problem of characterizing the critical points of \(E\) in \(H^1(I, M)\) is not interesting at all. This shouldn't really come as a surprise, as most interesting results from the classical theory are concerned with particular classes of variations of a curves, such as variations with fixed endpoints or variations through loops. In the next section we introduce two submanifolds of \(H^1(I, M)\), corresponding to the classes of variations previously described, and classify the critical points of the restrictions of \(E\) to such submanifolds. \subsection{The Critical Points of \(E\)} We begin with a technical lemma. \begin{lemma} The maps \(\sigma, \tau: H^1(I, M) \to M\) with \(\sigma(\gamma) = \gamma(0)\) and \(\tau(\gamma) = \gamma(1)\) are submersions. \end{lemma} \begin{proof} To see that \(\sigma\) and \(\tau\) are smooth it suffices to observe that their local representation in \(U_\gamma\) for \(\gamma \in {C'}^\infty(I, M)\) is given by the maps \begin{align*} U \subset H^1(W_\gamma) & \to T_{\gamma(0)} M & U \subset H^1(W_\gamma) & \to T_{\gamma(1)} M \\ X & \mapsto X_0 & X & \mapsto X_1 \end{align*} which are indeed smooth functions. This local representation also shows that \begin{align*} d\sigma_\gamma : H^1(\gamma^* TM) & \to T_{\gamma(0)} M & d\tau_\gamma : H^1(\gamma^* TM) & \to T_{\gamma(1)} M \\ X & \mapsto X_0 & X & \mapsto X_1 \end{align*} are surjective maps for all \(\gamma \in H^1(I, M)\). \end{proof} We can now show\dots \begin{theorem} The subspace \(\Omega_{p q} M \subset H^1(I, M)\) of curves joining \(p, q \in M\) is a submanifold whose tangent space \(T_\gamma \Omega_{p q} M\) is the subspace of \(H^1(\gamma^* TM)\) consisting of class \(H^1\) vector fields \(X\) along \(\gamma\) with \(X_0 = X_1 = 0\). Likewise, the space \(\Lambda M \subset H^1(I, M)\) of free loops is a submanifold whose tangent at \(\gamma\) is given by all \(X \in H^1(\gamma^* TM)\) with \(X_0 = X_1\). \end{theorem} \begin{proof} To see that these are submanifolds, it suffices to note that \(\Omega_{p q} M\) and \(\Lambda M\) are the inverse images of the closed submanifolds \(\{(p, q)\}, \{(p, p) : p \in M\} \subset M \times M\) under the submersion \((\sigma, \tau) : H^1(I, M) \to M \times M\). The characterization of their tangent bundles should also be clear: any curve \((-\epsilon, \epsilon) \to H^1(I, M)\) passing through \(\gamma \in \Omega_{p q} M\) whose image is contained in \(\Omega_{p q} M\) is a variation of \(\gamma\) with fixed endpoints, so its variational vector field \(X\) satisfies \(X_0 = X_1 = 0\). Likewise, any variation of a loop \(\gamma \in \Lambda M\) trough loops -- i.e. a curve \((-\epsilon, \epsilon) \to \Lambda M\) passing through \(\gamma\) -- satisfies \(X_0 = X_1\). \end{proof} Finally, as promised we will provide a characterization of the critical points of \(E\!\restriction_{\Omega_{p q} M}\) and \(E\!\restriction_{\Lambda M}\). \begin{theorem}\label{thm:critical-points-char-in-submanifolds} The critical points of \(E\!\restriction_{\Omega_{p q} M}\) are precisely the geodesics of \(M\) joining \(p\) and \(q\). The critical points of \(E\!\restriction_{\Lambda M}\) are the closed geodesics of \(M\) -- including the constant maps. \end{theorem} \begin{proof} We start by supposing that \(\gamma\) is a geodesic. Since \(\gamma\) is smooth, \[ d E_\gamma X = \int_0^1 \left\langle \dot\gamma(t), \frac\nabla\dt X \right\rangle \; \dt = \int_0^1 \frac\dd\dt \langle \dot\gamma(t), X_t \rangle - \left\langle \frac\nabla\dt \dot\gamma(t), X \right\rangle \; \dt = \langle \dot\gamma(1), X_1 \rangle - \langle \dot\gamma(0), X_0 \rangle \] Now if \(\gamma \in \Omega_{p q} M\) and \(X \in T_\gamma \Omega_{p q} M\) then \(d E_\gamma X = \langle \dot\gamma(1), 0 \rangle - \langle \dot\gamma(0), 0 \rangle = 0\). Likewise, if \(\gamma\) is a closed geodesic and \(X \in T_\gamma \Lambda M\) we find \(d E_\gamma X = 0\) since \(\dot\gamma(0) = \dot\gamma(1)\) and \(X_0 = X_1\). This establishes that the geodesics are indeed critical points of the restrictions of \(E\). Suppose \(\gamma \in \Omega_{p q} M\) is a critical point and let \(Y, Z \in H^1(\gamma^* TM)\) be such that \begin{align*} \frac\nabla\dt Y & = \partial \gamma & Y_0 & = 0 & \frac\nabla\dt Z & = 0 & Z_1 & = Y_1 \end{align*} Let \(X_t = Y_t - t Z_t\). Then \(X_0 = X_1 = 0\) and \(\frac\nabla\dt X = \partial \gamma - Z\). Furthermore, \[ \langle Z, \partial \gamma - Z \rangle_0 = \left\langle Z, \frac\nabla\dt X \right\rangle_0 = \int_0^1 \frac\dd\dt \langle Z_t, X_t \rangle \; \dt = \langle Z_1, X_1 \rangle - \langle Z_0, X_0 \rangle = 0 \] and \[ \langle \partial \gamma, \partial \gamma - Z \rangle_0 = \left\langle \partial \gamma, \frac\nabla\dt X \right\rangle_0 = d E_\gamma X = 0, \] which implies \(\norm{\partial \gamma - Z}_0^2 = 0\). In other words, \(\partial \gamma = Z \in H^1(\gamma^* TM)\) and therefore \(\frac\nabla\dt \dot\gamma(t) = \frac\nabla\dt Z = 0\) -- i.e. \(\gamma\) is a geodesic. Finally, if \(\gamma \in \Lambda M\) with \(\gamma(0) = \gamma(1) = p\) we may apply the argument above to conclude that \(\gamma\) is a geodesic joining \(p\) to \(q = p\). To see that \(\gamma\) is a closed geodesic apply the same argument again for \(\eta(t) = \gamma(1 + \sfrac{1}{2})\) to conclude that \(\dot\gamma(0) = \dot\eta(\sfrac{1}{2}) = \dot\gamma(1)\). \end{proof} We should point out that the first part of theorem~\ref{thm:critical-points-char-in-submanifolds} is a particular case of a result regarding critical points of the restriction of \(E\) to the submanifold \(H_{N_0, N_1}^1(I, M) \subset H^1(I, M)\) of curves joining submanifolds \(N_0, N_1 \subset M\): the critical points of \(E\!\restriction_{H_{N_0, N_1}^1(I, M)}\) are the geodesics \(\gamma\) joining \(N_0\) to \(N_1\) with \(\dot\gamma(0) \in T_{\gamma(0)} N_0^\perp\) and \(\dot\gamma(1) \in T_{\gamma(1)} N_1^\perp\). The proof of this result is essentially the same as that of theorem~\ref{thm:critical-points-char-in-submanifolds}, given that \(T_\gamma H_{N_0, N_1}^1(I, M)\) is subspace of \(H^1\) vector fields \(X\) along \(\gamma\) with \(X_0 \in T_{\gamma(0)} N_0\) and \(X_1 \in T_{\gamma(1)} N_1\). \subsection{Second Order Derivatives of \(E\)} Having establish a clear connection between geodesics and critical points of \(E\), the only thing we're missing to complete our goal of providing a modern account of the classical theory is a refurnishing of the formula for second variation of energy. Intuitively speaking, the second variation of energy should be a particular case of a formula for the second derivative of \(E\). The issue we face is, of course, that in general there is no such thing as ``the second derivative'' of a smooth function between manifolds. Nevertheless, the metric of \(H^1(I, M)\) allow us to discuss ``the second derivative'' of \(E\) in a meaningful sense by looking at the Hessian form, which we define in the following. \begin{definition} Given a -- possibly infinite-dimensional -- Riemannian manifold \(N\) and a smooth functional \(f : N \to \mathbb{R}\), we call the symmetric tensor \[ d^2 f(X, Y) = \nabla d f (X, Y) = X Y f - df \nabla_X Y \] \emph{the Hessian of \(f\)}. \end{definition} We can now apply the classical formula for the second variation of energy to compute the Hessian of \(E\) at a critical point. \begin{theorem} If \(\gamma\) is a critical point of \(E\!\restriction_{\Omega_{p q} M}\) then \begin{equation}\label{eq:second-variation-general} (d^2 E\!\restriction_{\Omega_{p q} M})_\gamma(X, Y) = \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle_0 - \langle R_\gamma X, Y \rangle_0, \end{equation} where \(R_\gamma : H^1(\gamma^* TM) \to H^1(\gamma^* TM)\) is given by \((R_\gamma X)_t = R(X_t, \dot\gamma(t)) \dot\gamma(t)\). Formula (\ref{eq:second-variation-general}) also holds for critical points of \(E\) in \(\Lambda M\). \end{theorem} \begin{proof} Given the symmetry of \(d^2 E\), it suffices to take \(X \in T_\gamma \Omega_{p q} M\) and show \[ d^2 E_\gamma(X, X) = \norm{\frac\nabla\dt X}_0^2 - \langle R_\gamma X, X \rangle_0 \] To that end, we fix a variation \(\{ \gamma_t \}_t\) of \(\gamma\) with fixed endpoints and variational field \(X\) and compute \[ \begin{split} d^2 E_\gamma(X, X) & = \left.\frac{\dd^2}{\dt^2}\right|_{t = 0} E(\gamma_t) \\ \text{(second variation of energy)} & = \int_0^1 \norm{\frac\nabla\dt X}^2 - \langle R(X_t, \dot\gamma(t)) \dot\gamma(t), X_t \rangle \; \dt \\ & = \norm{\frac\nabla\dt X}_0^2 - \langle R_\gamma X, X \rangle_0 \end{split} \] \end{proof} Next we discuss some further applications of the theory we've developed so far. In particular, we will work towards Morse's index theorem and and describe how one can apply it to establish the Jacobi-Darboux theorem. We begin with a technical lemma, whose proof amounts to an uninspiring exercise in analysis -- see lemma 2.4.6 of \cite{klingenberg}. \begin{lemma}\label{thm:inclusion-submnfds-is-compact} Let \(\Omega_{p q}^0 M \subset C^0(I, M)\) be the space of continuous curves joining \(p\) to \(q\). Then the inclusion \(\Omega_{p q} M \longhookrightarrow \Omega_{p q}^0 M\) is continuous and compact. Likewise, if \(M\) is compact and \(\Lambda^0 M \subset C^0(I, M)\) is the space of continuous free loops then the inclusion \(\Lambda M \longhookrightarrow \Lambda^0 M\) is continuous and compact. \end{lemma} As a first consequence, we prove\dots \begin{proposition}\label{thm:morse-index-e-is-finite} Given a critical point \(\gamma\) of \(E\) in \(\Omega_{p q} M\), the self-adjoint operator \(A_\gamma : T_\gamma \Omega_{p q} M \to T_\gamma \Omega_{p q} M\) given by \[ \langle A_\gamma X, Y \rangle_1 = \langle X, A_\gamma Y \rangle_1 = d^2 E_\gamma(X, Y) \] has the form \(A_\gamma = \operatorname{Id} + K_\gamma\) where \(K_\gamma : T_\gamma \Omega_{p q} M \to T_\gamma \Omega_{p q} M\) is a compact operator. The same holds for \(\Lambda M\) if \(M\) is compact. \end{proposition} \begin{proof} Consider \(K_\gamma = - \left( \operatorname{Id} - \frac{\nabla^2}{\dt^2} \right)^{-1} \circ (\operatorname{Id} + R_\gamma)\). We will show that \(K_\gamma\) is compact and that \(A_\gamma = \operatorname{Id} + K_\gamma\) for \(\gamma\) in both \(\Omega_{p q} M\) and \(\Lambda M\) -- in which case assume \(M\) is compact. Let \(\gamma \in \Omega_{p q} M\) be a critical point. By theorem~\ref{thm:critical-points-char-in-submanifolds} we know that \(\gamma\) is a geodesic. Let \(X, Y \in \Gamma(\gamma^* TM)\) with \(X_0 = X_1 = Y_0 = Y_1 = 0\). Then \begin{equation}\label{eq:compact-partial-result} \begin{split} \langle X, Y \rangle_1 & = \langle X, Y \rangle_0 + \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle \\ & = \langle X, Y \rangle_0 + \int_0^1 \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle \; \dt \\ & = \langle X, Y \rangle_0 + \int_0^1 \frac\dd\dt \langle X_t, Y_t \rangle - \left\langle \frac{\nabla^2}{\dt^2} X, Y \right\rangle \; \dt \\ & = \langle X, Y \rangle_0 - \left\langle \frac{\nabla^2}{\dt^2} X, Y \right\rangle_0 + \left.\langle X_t, Y_t \rangle\right|_{t = 0}^1 \\ & = \left\langle \left(\operatorname{Id} - \frac{\nabla^2}{\dt^2}\right) X, Y \right\rangle_0 \end{split} \end{equation} Since \(\Gamma(\gamma^* TM) \subset H^1(\gamma^* TM)\) is dense, (\ref{eq:compact-partial-result}) extends to all of \(T_\gamma \Omega_{p q} M\). Hence given \(X, Y \in T_\gamma \Omega_{p q} M\) we have \[ \begin{split} \langle A_\gamma X, Y \rangle_1 & = \left\langle \frac\nabla\dt X, \frac\nabla\dt Y \right\rangle_0 - \langle R_\gamma X, Y \rangle_0 \\ & = \langle X, Y \rangle_1 - \langle X, Y \rangle_0 - \langle R_\gamma X, Y \rangle_0 \\ & = \langle X, Y \rangle_1 - \langle (\operatorname{Id} + R_\gamma) X, Y \rangle_0 \\ & = \langle X, Y \rangle_1 - \left\langle \left( \operatorname{Id} - \frac{\nabla^2}{\dt^2} \right)^{-1} \circ (\operatorname{Id} + R_\gamma) X, Y \right\rangle_1 \\ & = \langle X, Y \rangle_1 + \langle K_\gamma X, Y \rangle_1 \\ \end{split} \] Now consider a critical point \(\gamma \in \Lambda M\) -- i.e. a closed geodesic. Equation (\ref{eq:compact-partial-result}) also holds for \(X, Y \in \Gamma(\gamma^* TM)\) with \(X_0 = X_1\) and \(Y_0 = Y_1\), so it holds for all \(X, Y \in T_\gamma \Lambda M\). Hence by applying the same argument we get \(\langle A_\gamma X, Y \rangle_1 = \langle (\operatorname{Id} + K_\gamma) X, Y \rangle_1\). As for the compactness of \(K_\gamma\) in the case of \(\Omega_{p q} M\), from (\ref{eq:compact-partial-result}) we get \(\norm{K_\gamma X}_1^2 = - \langle (\operatorname{Id} + R_\gamma) X, K_\gamma X \rangle_0\), so that proposition~\ref{thm:continuous-inclusions-sections} implies \begin{equation}\label{eq:compact-operator-quota} \norm{K_\gamma X}_1^2 \le \norm{\operatorname{Id} + R_\gamma} \cdot \norm{K_\gamma X}_\infty \cdot \norm{X}_0 \le \sqrt{2} \norm{\operatorname{Id} + R_\gamma} \cdot \norm{K_\gamma X}_1 \cdot \norm{X}_0 \end{equation} Given a bounded sequence \((X_n)_n \subset T_\gamma \Omega_{p q} M\), it follow from lemma~\ref{thm:inclusion-submnfds-is-compact} that \((X_n)_n\) is relatively compact as a \(C^0\)-sequence. From (\ref{eq:compact-operator-quota}) we then get that \((K_\gamma X_n)_n\) is relatively compact as an \(H^1\)-sequence, as desired. The same argument holds for \(\Lambda M\) if \(M\) is compact -- so that we can once more apply lemma~\ref{thm:inclusion-submnfds-is-compact}. \end{proof} Once again, the first part of this proposition is a particular case of a broader result regarding the space of curves joining submanifolds of \(M\): if \(N \subset M\) is a totally geodesic manifold of codimension \(1\) and \(\gamma \in H_{N, \{q\}}^1(I, M)\) is a critical point of the restriction of \(E\) then \(A_\gamma = \operatorname{Id} + K_\gamma\). These results aren't that appealing on their own, but they allow us to establish the following result, which is essential for stating Morse's index theorem. \begin{corollary} Given a critical point \(\gamma\) of \(E\!\restriction_{\Omega_{p q} M}\), there is an orthogonal decomposition \[ T_\gamma \Omega_{p q} M = T_\gamma^- \Omega_{p q} M \oplus T_\gamma^0 \Omega_{p q} M \oplus T_\gamma^+ \Omega_{p q} M, \] where \(T_\gamma^- \Omega_{p q} M\) is the finite-dimensional subspace spanned by eigenvectors with negative eigenvalues, \(T_\gamma^0 \Omega_{p q} M = \ker A_\gamma\) and \(T_\gamma^+ \Omega_{p q} M\) is the proper Hilbert subspace given by the closure of the subspace spanned by eigenvectors with positive eigenvalues. The same holds for critical points \(\gamma\) of \(E\!\restriction_{\Lambda M}\) and \(T_\gamma \Lambda M\) if \(M\) is compact. \end{corollary} \begin{definition}\label{def:morse-index} Given a critical point \(\gamma\) of \(E\!\restriction_{\Omega_{p q} M}\) we call the number \(\dim T_\gamma^- \Omega_{p q} M\) \emph{the \(\Omega\)-index of \(\gamma\)}. Likewise, we call \(\dim T_\gamma^- \Lambda M\) for a critical point \(\gamma\) of \(E\!\restriction_{\Lambda M}\) \emph{the \(\Lambda\)-index of \(\gamma\)}. Whenever the submanifold \(\gamma\) lies in is clear from context we refer to the \(\Omega\)-index or the \(\Lambda\)-index of \(\gamma\) simply by \emph{the index of \(\gamma\)}. \end{definition} This definition highlights one of the greatest strengths of our approach: while the index of a geodesic \(\gamma\) can be defined without the aid of the tools developed in here, by using of the Hessian form \(d^2 E_\gamma\) we can place definition~\ref{def:morse-index} in the broader context of Morse theory. In fact, the geodesics problems and the energy functional where among Morse's original proposed applications. Proposition~\ref{thm:morse-index-e-is-finite} and definition~\ref{def:morse-index} amount to a proof that the Morse index of \(E\) at a critical point \(\gamma\) is finite. We are now ready to state Morse's index theorem. \begin{theorem}[Morse] Let \(\gamma \in \Omega_{p q} M\) be a critical point of \(E\). Then the index of \(\gamma\) is given of the sum of the multiplicities of the proper conjugate points\footnote{By ``conjugate points of a geodesic $\gamma$'' we of course mean points conjugate to $\gamma(0) = p$ along $\gamma$.} of \(\gamma\) in the interior of \(I\). \end{theorem} Unfortunately we do not have the space to include the proof of Morse's theorem in here, but see theorem 2.5.9 of \cite{klingenberg}. The index theorem can be generalized for \(H_{N, \{q\}}^1(I, M)\) by replacing the notion of conjugate point with the notion of focal points of \(N\) -- see theorem 7.5.4 of \cite{gorodski} for the classical approach. What we are really interested in, however, is the following consequence of Morse's theorem. \begin{theorem}[Jacobi-Darboux]\label{thm:jacobi-darboux} Let \(\gamma \in \Omega_{p q} M\) be a critical point of \(E\). \begin{enumerate} \item If there are no conjugate points of \(\gamma\) then there exists a neighborhood \(U \subset \Omega_{p q} M\) of \(\gamma\) such that \(E(\eta) > E(\gamma)\) for all \(\eta \in U\) with \(\eta \ne \gamma\). \item Let \(k > 0\) be the sum of the multiplicities of the conjugate points of \(\gamma\) in the interior of \(I\). Then there exists an immersion \[ i : B^k \to \Omega_{p q} M \] of the unit ball \(B^k = \{v \in \mathbb{R}^k : \norm{v} < 1\}\) with \(i(0) = \gamma\), \(E(i(v)) < E(\gamma)\) and \(L(i(v)) < L(\gamma)\) for all nonzero \(v \in B^k\). \end{enumerate} \end{theorem} \begin{proof} First of all notice that given \(\eta \in U_\gamma\) with \(\eta = \exp_\gamma(X)\), \(X \in H^1(W_\gamma)\), the Taylor series for \(E(\eta)\) is given by \(E(\eta) = E(\gamma) + \frac{1}{2} d^2 E_\gamma(X, X) + \cdots\). More precisely, \begin{equation}\label{eq:energy-taylor-series} \frac {\abs{E(\exp_\gamma(X)) - E(\gamma) - \frac{1}{2} d^2 E_\gamma(X, X)}} {\norm{X}_1^2} \to 0 \end{equation} as \(X \to 0\). Let \(\gamma\) be as in \textbf{(i)}. Since \(\gamma\) has no conjugate points, it follows from Morse's index theorem that \(T_\gamma^- \Omega_{p q} M = 0\). Furthermore, by noticing that any piece-wise smooth \(X \in \ker A_\gamma\) is Jacobi field vanishing at \(p\) and \(q\) one can also show \(T_\gamma^0 \Omega_{p q} M = 0\). Hence \(T_\gamma \Omega_{p q} M = T_\gamma^+ \Omega_{p q} M\) and therefore \(d^2 E\!\restriction_{\Omega_{p q} M}\) is positive-definite. Now given \(\eta = \exp_\gamma(X)\) as before, (\ref{eq:energy-taylor-series}) implies that \(E(\eta) > E(\gamma)\), provided \(\norm{X}_1\) is sufficiently small. As for part \textbf{(ii)}, fix an orthonormal basis \(\{X_j : 1 \le j \le k\}\) of \(T_\gamma^- \Omega_{p q} M\) consisting of eigenvectors of \(A_\gamma\) with negative eigenvalues \(- \lambda_i\). Let \(\delta > 0\) and define \begin{align*} i : B^k & \to \Omega_{p q} M \\ v & \mapsto \exp_\gamma(\delta (v_1 \cdot X_1 + \cdots + v_k \cdot X_k)) \end{align*} Clearly \(i\) is an immersion for small enough \(\delta\). Moreover, from (\ref{eq:energy-taylor-series}) and \[ E(i(v)) = E(\gamma) - \frac{1}{2} \delta^2 \sum_j \lambda_j \cdot v_j + \cdots \] we find that \(E(i(v)) < E(\gamma)\) for sufficiently small \(\delta\). In particular, \(L(i(v))^2 \le E(i(v)) < E(\gamma) = L(\gamma)^2\). \end{proof} We should point out that part \textbf{(i)} of theorem~\ref{thm:jacobi-darboux} is weaker than the classical formulation of the Jacobi-Darboux theorem -- such as in theorem 5.5.3 of \cite{gorodski} for example -- in two aspects. First, we do not compare the length of curves \(\gamma\) and \(\eta \in U\). This could be amended by showing that the length functional \(L : H^1(I, M) \to \mathbb{R}\) is smooth and that its Hessian \(d^2 L_\gamma\) is given by \(C \cdot d^2 E_\gamma\) for some \(C > 0\). Secondly, unlike the classical formulation we only consider curves in an \(H^1\)-neighborhood of \(\gamma\) -- instead of a neighborhood of \(\gamma\) in \(\Omega_{p q} M\) in the uniform topology. On the other hand, part \textbf{(ii)} is definitively an improvement of the classical formulation: we can find curves \(\eta = i(v)\) shorter than \(\gamma\) already in an \(H^1\)-neighborhood of \(\gamma\). This concludes our discussion of the applications of our theory to the geodesics problem. We hope that these short notes could provide the reader with a glimpse of the rich theory of the calculus of variations and global analysis at large. We once again refer the reader to \cite[ch.~2]{klingenberg}, \cite[ch.~11]{palais} and \cite[sec.~6]{eells} for further insight on modern variational methods.