Skip to main content

Calculus of Variations, Conjugate Points and Morse Index

  • Chapter
  • First Online:
Elementary Symplectic Topology and Mechanics

Part of the book series: Lecture Notes of the Unione Matematica Italiana ((UMILN,volume 16))

  • 1880 Accesses

Abstract

Let us re-examine the classical conditions of Calculus of Variations geared to obtain a strong minimum (in the topology of the uniform convergence) for an arbitrary Lagrangian function \(L(t,q,\dot{q})\), convex in the velocities \(\dot{q}\). The result can be obtained in the geometric setup of symplectic geometry using the Hamiltonian description of the problem: the joint use of the theory of Poincaré-Cartan Integral Invariant and Young Inequality rapidly leads to the thesis. The Morse index of stationary curves is discussed in the mechanical case.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 49.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 64.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    The above uniform convexity may be weakened by the so-called Tonelli condition: (i) L is (simply) \(\dot{q}\)-convex and (ii) L is \(\dot{q}\)-superlinear at infinity: \(\lim _{\vert \dot{q}\vert \rightarrow +\infty }\frac{L(q,\dot{q})} {\vert \dot{q}\vert } = \infty \). In such a case also \(\mathcal{H}\) is Tonelli, in the p variables.

  2. 2.

    E.g. homotopically related each other by a continuous deformation over Λ n+1 .

  3. 3.

    In a unique way, thanks to the transversality condition.

  4. 4.

    See below: \(\varGamma _{t_{0},t_{1}}^{q_{0},q_{1}}\).

  5. 5.

    To be precise, we should write:

    $$\displaystyle\begin{array}{rcl} & & [t_{0},t_{1}] \times B^{n} \ni (t,v)\longmapsto (t,q(t,v); -\mathcal{H}(t,q(t,v),p(t,v)),p(t,v)) {}\\ & & \quad =\overbrace{ (t,q(t); -\mathcal{H}(t,q(t),p(t)),p(t))}^{\gamma (t)} {}\\ & & \qquad + (0,Q(t,v); \mathcal{H}(t,q(t),p(t)) -\mathcal{H}(t,q(t) + Q(t,v),p(t) + P(t,v)),P(t,v)) \in \mathbb{R}^{2n+2}, {}\\ & & \text{where}\ \ \ x(t,v) = x(t) + f(t,v) = (q(t) + Q(t,v),p(t) + P(t,v)), {}\\ & & \qquad f(t, 0) = (Q(t, 0),P(t, 0)) = 0. {}\\ \end{array}$$
  6. 6.

    We can see that the essential part of the system (5.15) for γ(t) is given by \(\dot{x}(t) = \mathbb{E}\nabla _{x}\mathcal{H}\big(t,x(t)\big)\), since the evolution of q 0(t) = t and \(p_{0}(t) = -\mathcal{H}\) is a trivial consequence.

  7. 7.

    Parameters t and λ are independent between them.

  8. 8.

    The same regularity (5.40) of L, thanks to Proposition 5.3.

  9. 9.

    The n × n-matrices \(A,B,\dot{A},\dot{B}\) are evalued for initial and final times τ α and τ α+1.

  10. 10.

    Continuous, but not differentiable.

  11. 11.

    To be honest, the introduction proposed above of functions q(⋅ ) twice-differentiable is only needed to establish the equivalence between Lagrange equations and Gateaux-stationary points for Hamilton’s functional. We overcome this a priori strange requirement by introducing the DuBois-Reymond Lemma.

  12. 12.

    \(h \in H_{0}^{1}([a,b], \mathbb{R}^{n})\;\Longleftrightarrow\;h \in H^{1}([a,b], \mathbb{R}^{n})\ \text{and}\ h(a) = 0 = h(b)\).

  13. 13.

    Note that J′[q]h and J″[q](h,h) are the first and second Fréchet derivatives with respect the norm \(\|\cdot \|.\)

  14. 14.

    See proposition 2.3 in [1].

  15. 15.

    See e.g. [89].

  16. 16.

    \(h \in H_{0}^{1}([t_{0},t_{1}], \mathbb{R}^{n})\;\Longleftrightarrow\;h \in H^{1}([t_{0},t_{1}], \mathbb{R}^{n})\ \text{e}\ h(t_{0}) = 0 = h(t_{1})\).

  17. 17.

    Note that \(\alpha _{2} =\alpha _{1}\frac{(t_{1}-t_{0})^{2}} {2}\).

References

  1. A. Abbondandolo, M. Schwarz, A smooth pseudo-gradient for the Lagrangian action functional. Adv. Nonlinear Stud. 9, 597–623 (2009)

    MATH  MathSciNet  Google Scholar 

  2. V. Alexeev, V. Tikhomirov, S. Fomin, Commande Optimale (éditions MIR, Moscou, 1982)

    Google Scholar 

  3. M. Berger, M. Berger, Perspectives in Nonlinearity (W. A. Benjamin, New York, 1968)

    MATH  Google Scholar 

  4. F. Cardin, G. De Marco, A. Sfondrini, Finite reduction and Morse index estimates for mechanical systems. NoDEA Nonlinear Differ. Equ. Appl. 18(5), 557–569 (2011)

    Article  MATH  Google Scholar 

  5. G. Contreras, R. Iturriaga, Global Minimizers of Autonomous Lagrangians. 22o Colóquio Brasileiro de Matemática (Instituto de Matemática Pura e Aplicada (IMPA), Rio de Janeiro, 1999), 148pp

    Google Scholar 

  6. M.P. do Carmo, Riemannian Geometry (Birkhäuser, Boston, 1992), xiv+300pp

    Google Scholar 

  7. B.A. Dubrovin, A.T. Fomenko, S.P. Novikov, Modern Geometry—Methods and Applications. Part III. Introduction to Homology Theory. Graduate Texts in Mathematics, vol. 124 (Springer, New York, 1990), x+416pp

    Google Scholar 

  8. L.C. Evans, Partial Differential Equations. Graduate Studies in Mathematics, vol. 19, 2nd edn. (American Mathematical Society, Providence, 2010), xxii+749pp

    Google Scholar 

  9. A. Fathi, Weak KAM theory in Lagrangian dynamics. Preliminary version

    Google Scholar 

  10. M. Mazzucchelli, The Lagrangian Conley conjecture. Comment. Math. Helv. 86(1), 189–246 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  11. M. Mazzucchelli, Critical Point Theory for Lagrangian Systems. Progress in Mathematics, vol. 293 (Birkhäuser/Springer, Basel, 2012), xii+187pp

    Google Scholar 

  12. J. Milnor, Morse Theory, vol. 51 (Princeton University Press, Princeton, 1963), vi+153pp

    Google Scholar 

  13. M. Schwarz, Morse Homology (Birkhäuser, Basel, 1993), ix+235pp

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendix

Appendix

5.1.1 Strong and Weak Minima

Consider the set (typicalFootnote 11) of the curves in the calculus of variations for the functional of Hamilton’s problem \(J[q(\cdot )] =\int _{ 0}^{T}L(q(t),\dot{q}(t),t)\mathit{dt}\):

$$\displaystyle{\varGamma = \left \{q(\cdot ) \in W^{1,2}([0,T]; \mathbb{R}^{n}) = H^{1}([0,T]; \mathbb{R}^{n}): \quad q(0) = q_{ 0},\ q(T) = q_{T}\right \}}$$
$$\displaystyle{\left \{\begin{array}{lcl} \vert \vert q(\cdot )\vert \vert _{H^{1}} &:=&\left ( \frac{1} {T}\int _{0}^{T}\vert q(t)\vert ^{2}\mathit{dt}\right )^{1/2} + \left ( \frac{1} {T}\int _{0}^{T}\vert \dot{q}(t)\vert ^{2}\mathit{dt}\right )^{1/2} \\ & \ & \\ \vert \vert q(\cdot )\vert \vert _{\mathcal{C}^{1}} &:=&\sup _{t\in [0,T]}\vert q(t)\vert +\sup _{t\in [0,T]}\vert \dot{q}(t)\vert,\\ &\ & \\ \vert \vert q(\cdot )\vert \vert _{\mathcal{C}^{0}} &:=&\sup _{t\in [0,T]}\vert q(t)\vert.\\ \end{array} \right.}$$

By definition, the Sobolev space H 1 is precisely composed by the completion of the \(\mathcal{C}^{1}\) curves with respect to the above norm \(\vert \vert \cdot \vert \vert _{H^{1}}\); it comes out that they are not necessarily differentiable: for the Sobolev immersion theorem, \(H^{1}\hookrightarrow \mathcal{C}^{0}\), one has that the curves are continuous. For every curve \(q(\cdot ) \in \varGamma \cap \mathcal{C}^{1}\) one has that

$$\displaystyle{\vert \vert q(\cdot )\vert \vert _{\mathcal{C}^{0}} \leq \vert \vert q(\cdot )\vert \vert _{\mathcal{C}^{1}},}$$

and hence, the \(q(\cdot ) \in \varGamma \cap \mathcal{C}^{1}\) such that \(\vert \vert q(\cdot )\vert \vert _{\mathcal{C}^{1}} \leq r\) and such that q(⋅ ) belongs to B 1(r), are all in B 0(r): this means that the \(\mathcal{C}^{1}\) topology is finer (or stronger) than the \(\mathcal{C}^{0}\) topology.

Despite this,

  1. (i)

    If q (⋅ ) is such that for some \(\varepsilon > 0\):

    $$\displaystyle{J[q^{{\ast}}(\cdot )] \leq J[q(\cdot )],\qquad \forall q(\cdot ) - q^{{\ast}}(\cdot ) \in B_{ 0}(\varepsilon ) (\star )}$$

    we will say that q (⋅ ) is a strong minimum,

  2. (ii)

    If q (⋅ ) is such that for some \(\varepsilon > 0\):

    $$\displaystyle{J[q^{{\ast}}(\cdot )] \leq J[q(\cdot )],\qquad \forall q(\cdot ) - q^{{\ast}}(\cdot ) \in B_{ 1}(\varepsilon ) (\star \star )}$$

    we will say that q (⋅ ) is a weak minimum.

Observe that, if q (⋅ ) is a strong minimum, then it also is a weak minimum, but the vice-versa is not true, in fact:

$$\displaystyle{B_{1}(\varepsilon ) \subset B_{0}(\varepsilon ).}$$

5.1.2 Variational Theory in H 1 = W 1, 2

Lemma 5.3 (DuBois-Reymond Lemma)

Let \(\varphi \in L^{2}([a,b], \mathbb{R}^{n})\) and suppose that Footnote 12

$$\displaystyle{\int _{a}^{b}\varphi (t)\dot{h}(t)\mathit{dt} = 0\quad \forall h \in H_{ 0}^{1}([a,b], \mathbb{R}^{n})}$$

Then

$$\displaystyle{\varphi (t) \equiv \text{const.}\ \ a.e.}$$

Proof

Define \(\mu:= \frac{1} {b-a}\int _{a}^{b}\varphi (t)\mathit{dt}\), it is meaningful since \(\varphi \in L^{2} \subset L^{1}\). Let \(h(t):=\int _{ a}^{t}(\varphi (s)-\mu )ds\), we see that this choice of h is in H 0 1: h(a) = 0 = h(b), \(\dot{h} =\varphi -\mu \in L^{2}\) and \(\vert \vert h\vert \vert _{L^{2}}^{2} =\int _{ a}^{b}(\int _{a}^{t}(\varphi (s)-\mu )ds)\cdot (\int _{a}^{t}(\varphi (s)-\mu )\mathit{ds})\mathit{dt} \leq (b-a)\sup _{t\in [a,b]}\vert \int _{a}^{t}(\varphi (s)-\mu )ds\vert ^{2} < +\infty \) because \(\varphi \in L^{1}\).

$$\displaystyle{0 =\int _{ a}^{b}\varphi (t)\dot{h}(t)\mathit{dt} =\int _{ a}^{b}\varphi (t)(\varphi (t)-\mu )\mathit{dt} =\int _{ a}^{b}(\varphi (t)-\mu )^{2}\mathit{dt},}$$

since \(\varphi -\mu \in L^{2}\), then: \(\varphi (t) \equiv \mu \ \ a.e.\) □ 

Lemma 5.4 (First Variation Lemma)

Suppose

$$\displaystyle{\mathit{dJ}(q)h =\int _{ a}^{b}[L_{ q}(q(t),\dot{q}(t))h(t) + L_{\dot{q}}(q(t),\dot{q}(t))\dot{h}(t)]\mathit{dt} = 0}$$

\(\forall h \in H_{0}^{1}([a,b], \mathbb{R}^{n})\) and

$$\displaystyle{q \in \varGamma _{a,b}^{q_{0},q_{1} }:= \left \{q(\cdot ) \in H^{1}([a,b], \mathbb{R}^{n}):\ q(a) = q_{ 0},\ q(b) = q_{1}\right \}}$$

If \(L: \mathbb{R}^{n} \times \mathbb{R}^{n} \rightarrow \mathbb{R}\) is such that the above Gateaux differential there exists and

  1. (i)

    \(P(t):=\int _{ a}^{t}L_{q}(q(s),\dot{q}(s))ds \in L^{2}\) ,

  2. (ii)

    \(L_{\dot{q}}(q(t),\dot{q}(t)) \in L^{2}\) ,

then:

$$\displaystyle{ P - L_{\dot{q}} = \text{const.}\quad a.e. }$$
(5.55)

Remark 5.2

Note that the above relation is precisely the integral version of the Euler-Lagrange equations.

Proof

\(0 =\int _{ a}^{b}[L_{q}(q(t),\dot{q}(t))h(t)+L_{\dot{q}}(q(t),\dot{q}(t))\dot{h}(t)]\mathit{dt} = P(t)h(t)\vert _{a}^{b}-\int _{a}^{b}(P-L_{\dot{q}})\dot{h}\mathit{dt}\), the above Lemma leads us to the thesis. □ 

The mechanical case Let \(L = \frac{1} {2}m\vert \dot{q}\vert ^{2} - V (q)\), \(L_{\dot{q}} = m\dot{q}\) and if q ∈ H 1 then \(m\dot{q} \in L^{2}\). If \(V: \mathbb{R}^{n} \rightarrow \mathbb{R}\) is \(\mathcal{C}^{1}\), then

$$\displaystyle{P(t) = -\int _{a}^{t}\nabla U(q(s))\mathit{ds},}$$

is a continuous function on the compact set [a, b], then P ∈ L 2.

5.1.3 Local (for Short Time Interval) Minimum in the Calculus of Variations

First, we recall a general fact on the role of the coercivity in calculus of variations

Proposition 5.1 (Local coercivity induces local minimum)

  1. (i)

    Let \((X,\|\cdot \|)\) be a normed affine space of curves between two fixed configurations, \(X = X_{t_{0},t_{1}}^{q_{0},q_{1}} = q + X_{t_{ 0},t_{1}}^{0,0}\) for a fixed q ∈ X,

  2. (ii)

    Let \(J: (X,\|\cdot \|) \rightarrow \mathbb{R}\) be a real-valued functional of class \(\mathcal{C}^{2}(X; \mathbb{R})\) ,

  3. (iii)

    Let q ∈ X be a stationary curve Footnote 13 for J:

    $$\displaystyle{J'[q]h = 0,\ \forall h \in X_{t_{0},t_{1}}^{0,0},}$$
  4. (iv)

    Let suppose there exists a positive constant α > 0 such that (local coercivity):

    $$\displaystyle{J''[q](h,h) \geq \alpha \| h\|^{2}\quad \text{for any}\ h \in X_{ t_{0},t_{1}}^{0,0}}$$

Then q is a strict local minimum for J in the induced topology from \(\|\cdot \|\) .

Proof

From Taylor expansion,

$$\displaystyle{J[q + h] - J[q] = \frac{1} {2!}J''[q](h,h) +\| h\|^{2}g(q,h),}$$

where g(q, h) is an infinitesimal function for \(\|h\| \rightarrow 0\). Thus,

$$\displaystyle{J[q + h] - J[q] \geq \big (\frac{1} {2}\alpha + g(q,h)\big)\|h\|^{2}}$$

and for small h (in the choosen norm \(\|\cdot \|\)) and different from zero (i.e., the curve h(t) ≡ 0), surely \((\frac{1} {2}\alpha + g(q,h)) > 0\) and hence

$$\displaystyle{J[q + h] > J[q].}$$

    □ 

Recently, it has been proved that the functional J is \(\mathcal{C}^{2}\) in H 1 if and only if the restriction \(\dot{q}\mapsto L(t,q,\dot{q})\) is polynomial at most of degree 2,Footnote 14 just it precisely happens in the mechanical and geodesic cases!

For general Lagrangians with \(\dot{q}\mapsto L(t,q,\dot{q})\) convex and quadratically growing at infinity we can simply say that J is \(\mathcal{C}^{1,1}\) in H 1, that is \(\mathcal{C}^{1}\) with first derivative Lipschitz.Footnote 15

We are ready to the following

Proposition 5.2 (Local minimum in H 1, the mechanical case)

Let \(J(q) =\int _{ t_{0}}^{t_{1}}[\frac{1} {2}m\vert \dot{q}(t)\vert ^{2} - V (q(t))]\mathit{dt}\) and Footnote 16

$$\displaystyle{J'[q]h = 0\quad \forall h \in H_{0}^{1}([t_{ 0},t_{1}], \mathbb{R}^{n})}$$

Under the hypothesis that the interval [t 0 ,t 1 ] is small enough,

$$\displaystyle{t_{1} - t_{0} < \sqrt{2m \left (\max _{t\in [t_{0 },t_{1 } ] } \left (\mathrm{max }\vert \mathrm{spec }\nabla ^{2 } V \left (q(t) \right ) \vert \right ) \right ) ^{-1}},}$$

there exists a suitable constant α > 0 such that

$$\displaystyle{J''[q](h,h) = \frac{d^{2}} {d\lambda ^{2}}J[q +\lambda h]\vert _{\lambda =0} \geq \alpha \| h\|_{H^{1}}^{2},\qquad \forall h \in H_{ 0}^{1},}$$

so that q is a local minimum for J in H 1 .

Proof

$$\displaystyle{\begin{array}{rcl} \frac{d^{2}} {d\lambda ^{2}} J[q +\lambda h]\vert _{\lambda =0} & =&\frac{d} {d\lambda }\{\frac{d} {d\lambda }J[q +\lambda h]\}\vert _{\lambda =0}\\ \\ & =&\frac{d} {d\lambda }\{\int _{t_{0}}^{t_{1}}[m(\dot{q} +\lambda \dot{ h}) \cdot \dot{ h} -\nabla V (q +\lambda h)] \cdot h\,\mathit{dt}\}\vert _{\lambda =0} \\ \\ & =&\int _{t_{0}}^{t_{1}}m\vert \dot{h}\vert ^{2}\mathit{dt} -\int _{t_{ 0}}^{t_{1}}\nabla _{\mathit{ ij}}^{2}V (q)h^{i}h^{j}\mathit{dt}\\ \\ & \geq &\int _{t_{0}}^{t_{1}}m\vert \dot{h}(t)\vert ^{2}\mathit{dt} -\int _{t_{ 0}}^{t_{1}}\mathrm{max}\vert \mathrm{spec}\nabla ^{2}V (q(t))\vert \cdot \vert h(t)\vert ^{2}\mathit{dt}, \end{array} }$$
$$\displaystyle{ \begin{array}{rcl} \frac{d^{2}} {d\lambda ^{2}} J[q +\lambda h]\vert _{\lambda =0} & \geq &\int _{t_{0}}^{t_{1}}m\vert \dot{h}\vert ^{2}\mathit{dt} - c\int _{t_{ 0}}^{t_{1}}\vert h\vert ^{2}\mathit{dt},\end{array} }$$
(5.56)

where \(c:=\max _{t\in [t_{0},t_{1}]}(\mathrm{max}\vert \mathrm{spec}\nabla ^{2}V (q(t))\vert )\). In order to estimate \(\int _{t_{0}}^{t_{1}}\vert \dot{h}\vert ^{2}\mathit{dt}\) with \(\int _{t_{0}}^{t_{1}}\vert h\vert ^{2}\mathit{dt}\) we recall the Cauchy-Schwarz inequality:

$$\displaystyle{\left (\int _{a}^{b}\mathit{fgdt}\right )^{2} \leq \int _{ a}^{b}f^{2}\mathit{dt}\int _{ a}^{b}g^{2}\mathit{dt}.}$$

Using the fact that h(t 0) = 0, we have that

$$\displaystyle{\int _{t_{0}}^{t_{1} }\vert h(t)\vert ^{2}\mathit{dt} =\int _{ t_{0}}^{t_{1} }\sum _{i=1}^{3}(h^{i}(t))^{2}\mathit{dt} =\int _{ t_{0}}^{t_{1} }\sum _{i=1}^{3}\left (\int _{ t_{0}}^{t}\dot{h}^{i}(\tau )d\tau \right )^{2}\mathit{dt}.}$$

We estimate this last term by means of Cauchy-Schwarz,

$$\displaystyle{\sum _{i=1}^{n}\left (\int _{ t_{0}}^{t}1 \cdot \dot{ h}^{i}(\tau )d\tau \right )^{2} \leq \sum _{ i=1}^{n}\int _{ t_{0}}^{t}d\tau \int _{ t_{0}}^{t}\left (\dot{h}^{i}(\tau )\right )^{2}d\tau = (t - t_{ 0})\int _{t_{0}}^{t}\vert \dot{h}(\tau )\vert ^{2}d\tau.}$$

Hence

$$\displaystyle\begin{array}{rcl} & \int _{t_{0}}^{t_{1}}\vert h(t)\vert ^{2}\mathit{dt} \leq \int _{t_{ 0}}^{t_{1}}(t - t_{ 0})\left (\int _{t_{0}}^{t}\vert \dot{h}(\tau )\vert ^{2}d\tau \right )\mathit{dt} & {}\\ & \leq \int _{t_{0}}^{t_{1}}(t - t_{0})\mathit{dt}\int _{t_{ 0}}^{t_{1}}\vert \dot{h}(\tau )\vert ^{2}d\tau = \frac{1} {2}(t_{1} - t_{0})^{2}\int _{ t_{0}}^{t_{1}}\vert \dot{h}\vert ^{2}\mathit{dt}.& {}\\ \end{array}$$

Then we have that

$$\displaystyle{ \|h\|_{L^{2}}^{2} \leq \frac{1} {2}(t_{1} - t_{0})^{2}\|\dot{h}\|_{ L^{2}}^{2}. }$$
(5.57)

We obtain for (5.56):

$$\displaystyle\begin{array}{rcl} & J''[q](h,h) = \frac{d^{2}} {d\lambda ^{2}} J[q +\lambda h]\vert _{\lambda =0} \geq m\|\dot{h}\|_{L^{2}}^{2} - c\|h\|_{L^{2}}^{2}& {}\\ & \geq \left ( \frac{2m} {(t_{1}-t_{0})^{2}} - c\right )\|h\|_{L^{2}}^{2}. & {}\\ \end{array}$$

We see that \(\alpha _{1}:= \frac{2m} {(t_{1}-t_{0})^{2}} - c > 0\) if and only if \(t_{1} - t_{0} < \sqrt{\frac{2m} {c}}\). On the other hand, still with (5.57), the following estimate is holding

$$\displaystyle{J''[q](h,h) \geq m\|\dot{h}\|_{L^{2}}^{2} - c\|h\|_{ L^{2}}^{2} \geq \left (m - c\frac{1} {2}(t_{1} - t_{0})^{2}\right )\|\dot{h}\|_{ L^{2}}^{2}}$$

where \(\alpha _{2}:= m -\frac{1} {2}c(t_{1} - t_{0})^{2} > 0\) if and only if (as above) \(t_{1} - t_{0} < \sqrt{\frac{2m} {c}}\). At the end, the obtain an estimateFootnote 17 in the Sobolev H 1 norm:

$$\displaystyle{J''[q](h,h) \geq \frac{1} {2}\mathrm{min}\{\alpha _{1},\alpha _{2}\}(\|h\|_{L^{2}}^{2} +\|\dot{ h}\|_{ L^{2}}^{2}) = \frac{1} {2}\mathrm{min}\{\alpha _{1},\alpha _{2}\}\|h\|_{H^{1}}^{2} =\alpha \| h\|_{ H^{1}}^{2}}$$

 □ 

Remark 5.3 (On the existence)

Our effort here has been to highlight the quality of the critical curves using terms involving the second derivatives of the functional; if we are interested simply in search of the minimum of the functional, the only condition of uniform convexity of L with respect to \(\dot{q}\), without restriction on the size of the time interval, would be sufficient to ensure the existence of the minimum in H 1: this is a Tonelli like problem, see e.g. [60], Th. 5, p. 453. An alternative way to capture the existence could be to observe that J, if sup | V ″(q) |  < +, can be finitely reduced to a Generating Function Quadratic at Infinity (GFQI, see Chap. 7) which is Palais-Smale, see [32].

5.1.4 A Regularity Result

By means of the Legendre diffeomorphism \(\mathcal{T}\) – see (5.4) – we are able to establish the following main connection between the regularity of the solutions q(⋅ ) and the regularity of L: The H 1 critical curves q(⋅ ) have the same regularity of the Lagrangian L.

Proposition 5.3

Let q(⋅) ∈ H 1 be a curve satisfying the integral version  (5.55) of the Euler-Lagrange equations. Suppose that L is a C k (k ≥ 2) Tonelli Lagrangian, then q(⋅) is C k .

Proof

By Sobolev embedding theorem, we know that q ∈ C 0. We combine the integral form of the Euler-Lagrange equations, which are satisfied by q(⋅ ), with the conjugate momenta. We obtain:

$$\displaystyle{ p(t) = \frac{\partial L} {\partial \dot{q}} (t,q(t),\dot{q}(t)) =\int _{ 0}^{t} \frac{\partial } {\partial q}L(\tau,q(\tau ),\dot{q}(\tau ))d\tau + K }$$
(5.58)

So p(⋅ ), being equal to an integral function, is continuous. By the \(\mathcal{C}^{k-1}\) diffeomorphism \(\mathcal{T}\), which is in particular a homeomorphism, we obtain that

$$\displaystyle{ (t,q(t),\dot{q}(t)) = \mathcal{T}^{-1}(t,q(t),p(t)) }$$
(5.59)

is continuous, and in particular \(\dot{q}(\cdot )\) is continuous, so \(q(\cdot ) \in \mathcal{C}^{1}\). The same technique can be repeated as many times as the order of regularity of \(\mathcal{T}\), until we obtain the thesis: inserting now \(q(\cdot ) \in \mathcal{C}^{1}\) into (5.58) we see that \(p(\cdot ) \in \mathcal{C}^{1}\), and, by (5.59), \(\dot{q}(\cdot ) \in \mathcal{C}^{\mathrm{min}(1,k-1)}\) or \(q(\cdot ) \in \mathcal{C}^{\mathrm{min}(2,k)}\). If k > 2, at the next step: \(q(\cdot ) \in \mathcal{C}^{\mathrm{min}(3,k)}\). All this, until \(q(\cdot ) \in \mathcal{C}^{k}\). □ 

5.1.5 On the Linearized Flow

Working in \(\mathbb{R}^{n}\) for the sake of simplicity, we denote by

$$\displaystyle{ \tilde{x}(t,x) =\varPhi _{ X}^{t}(x) }$$
(5.60)

the flow relative to the vector field X:

$$\displaystyle{ \mathbb{R}^{n} \ni x\mapsto (x,X(x)) \in T\mathbb{R}^{n}, }$$
(5.61)
$$\displaystyle{ \left \{\begin{array}{rcl} \frac{d} {\mathit{dt}}\tilde{x}(t,x)& =&X(\tilde{x}(t,x)),\\ \\ \tilde{x}(0,x)& =&x.\end{array} \right. }$$
(5.62)

Let consider the linearized differential equation of (5.62) around the curve solution (5.60):

$$\displaystyle{ \left \{\begin{array}{rcl} \frac{d} {\mathit{dt}}\tilde{h}(t,h_{0})& =&DX(\tilde{x}(t,x))\tilde{h}(t,h_{0}),\\ \\ \tilde{h}(0,h_{0})& =&h_{0}.\end{array} \right. }$$
(5.63)

Proposition 5.4

The flow of the linearized problem is the linearized of the flow of the non linear problem:

$$\displaystyle{ \tilde{h}(t,h_{0}) = D\varPhi _{X}^{t}(x)h_{ 0}\ \ \ \Big(= \frac{\partial \tilde{x}} {\partial x}(t,x)h_{0}\Big) }$$
(5.64)

Proof

$$\displaystyle\begin{array}{rcl} & \frac{d} {\mathit{dt}}\tilde{h}(t,h_{0}) = \frac{\partial ^{2}\tilde{x}} {\partial t\,\partial x}(t,x)h_{0} = \frac{\partial ^{2}\tilde{x}} {\partial x\,\partial t}(t,x)h_{0} = \frac{\partial } {\partial x}X(\tilde{x}(t,x))h_{0},& {}\\ & = DX(\tilde{x}(t,x))\frac{\partial \tilde{x}} {\partial x}(t,x)h_{0} = DX(\tilde{x}(t,x))\tilde{h}(t,h_{0}). & {}\\ \end{array}$$

 □ 

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Cardin, F. (2015). Calculus of Variations, Conjugate Points and Morse Index. In: Elementary Symplectic Topology and Mechanics. Lecture Notes of the Unione Matematica Italiana, vol 16. Springer, Cham. https://doi.org/10.1007/978-3-319-11026-4_5

Download citation

Publish with us

Policies and ethics