









Study with the several resources on Docsity
Earn points by helping other students or get them with a premium plan
Prepare for your exams
Study with the several resources on Docsity
Earn points to download
Earn points by helping other students or get them with a premium plan
Community
Ask the community for help and clear up your study doubts
Discover the best universities in your country according to Docsity users
Free resources
Download our free guides on studying techniques, anxiety management strategies, and thesis advice from Docsity tutors
The properties of symplectic vector fields in the context of conservative mechanical systems. It explains how these systems have equations of motion that are symplectic and can be expressed in Hamiltonian form. The document also highlights the differences between symplectic and all smooth vector fields, including the presence of a first integral (energy) and a preserved volume, and the fact that equilibrium points cannot be asymptotically stable in their energy level.
Typology: Exams
1 / 17
This page cannot be seen from the preview
Don't miss anything!
Conservative mechanical systems have equations of mo- tion that are symplectic and can be expressed in Hamilto- nian form. The generic properties within the class of sym- plectic vector fields are quite different from those within the class of all smooth vector fields: the system always has a first integral (“energy”) and a preserved volume, and equilibrium points can never be asymptotically stable in their energy level. — John Guckenheimer
fast track: chapter 9, p. 155
8.1 Hamiltonian flows
“· · · to do this business right is a thing of far greater diffi- culty than I was aware of.” — Sir Isaac Newton, in a letter to Edmund Halley
appendix A
remark 2.
example 8. p. 149
remark 8.
example 8. p. 149
example ?? p. ??
8.3 Stability of Hamiltonian flows
8.3.1 Canonical transformations
Figure 8.3: Stability of a symplectic map in R^4.
complex saddle saddle−center
degenerate saddle
(2) (2)
real saddle
generic center degenerate center
(2)
(2)
example 8. p. 150
example 8. p. 150
example 8. p. 151
8.5 Poincaré invariants
f t(C)
C
′
C
C
i= 1
V
ΩV
in depth: appendix A8.1, p. 845
Résumé
maximal possible rank. The odd symplectic groups Sp(2D + 1) are not semisimple. If you care about group theory for its own sake (the dynamical systems symmetry reduction techniques of chapter 12 are still too primitive to be applicable to Quantum Field Theory), chapter 14 of ref. [4] is fun, too.
Referring to the Sp(d) Lie algebra elements as ‘Hamiltonian matrices’ as one some- times does [5, 28] conflicts with what is meant by a ‘Hamiltonian matrix’ in quantum mechanics: the quantum Hamiltonian sandwiched between vectors taken from any com- plete set of quantum states. We are not sure where this name comes from; Dragt cites refs. [8, 10], and chapter 17 of his own book in progress [6]. Fulton and Harris [8] use it. Certainly Van Loan [22] uses in 1981, and Taussky in 1972. Might go all the way back to Sylvester?
Dream student Henriette Roux wants to know: “Dynamics equals a Hamiltonian plus a bracket. Why don’t you just say it?” A: “It is true that in the tunnel vision of atomic mechanicians the world is Hamiltonian. But it is much more wondrous than that. This chapter starts with Newton 1687: force equals acceleration, and we always replace a higher order time derivative with a set of first order equations. If there are constraints, or fully relativistic Quantum Field Theory is your thing, the tool of choice is to recast New- ton equations as a Lagrangian 1788 variational principle. If you still live in material but non-relativistic world and have not gotten beyond Heisenberg 1925, you will find Hamil- ton’s 1827 principal function handy. The question is not whether the world is Hamiltonian
together– was introduced into mathematics by Hermann Weyl. ‘Canonical’ lineage is
church-doctrinal: Greek ‘kanon’, referring to a reed used for measurement, came to mean
in Latin a rule or a standard.
in (8.3), is set by the convention that the Hamilton’s principal function (for energy con-
serving flows) is given by R(q, q′, t) =
∫ (^) q′ q pidqi^ −^ Et. With this sign convention the action along a classical path is minimal, and the kinetic energy of a free particle is positive. Any finite-dimensional symplectic vector space has a Darboux basis such that ω takes form (8.6). Dragt [5] convention for phase-space variables is as in (8.2). He calls the dynam- ical trajectory x 0 → x(x 0 , t) the ‘transfer map’, something that we will avoid here, as it conflicts with the well established use of ‘transfer matrices’ in statistical mechanics.
come paired as Λ, 1/Λ, and complex eigenvalues come either in Λ, Λ∗^ pairs, |Λ| = 1, or
Λ, 1/Λ, Λ∗, 1/Λ∗^ loxodromic quartets. As most maps studied in introductory nonlinear
dynamics are 2d, you have perhaps never seen a loxodromic quartet. How likely are we to run into such things in higher dimensions? According to a very extensive study of
periodic orbits of a driven billiard with a four dimensional phase space, carried in ref. [16], the three kinds of eigenvalues occur with about the same likelihood.
of short periodic pulses, as can be physically implemented, for instance, by pulsed optical
lattices in cold atoms physics. On the theoretical side, standard maps illustrate a number of important features: small k values provide an example of KAM perturbative regime
(see ref. [14]), while larger k’s illustrate deterministic chaotic transport [3, 18], and the transition to global chaos presents remarkable universality features [12, 13, 25]. The
quantum counterpart of this model has been widely investigated, as the first example
where phenomena like quantum dynamical localization have been observed [1]. Stability residue was introduced by Greene [12]. For some hands-on experience of the standard
map, download Meiss simulation code [19].
system exhibits ‘chaos’ if its orbits are locally unstable (positive Lyapunov exponent) and globally mixing (positive entropy). In sect. 6.2 we shall define Lyapunov exponents and
discuss their evaluation, but already at this point it would be handy to have a few quick nu- merical methods to diagnose chaotic dynamics. Laskar’s frequency analysis method [15]
is useful for extracting quasi-periodic and weakly chaotic regions of state space in Hamil- tonian dynamics with many degrees of freedom. For pointers to other numerical methods,
see ref. [26].
References
Example 8.1. Unforced undamped Duffing oscillator. When the damping term is removed from the Duffing oscillator (2.22), the system can be written in Hamiltonian form,
H(q, p) =
p^2 2
q^2 2
q^4 4
This is a 1-dof Hamiltonian system, with a 2-dimensional state space, the plane (q, p). The Hamilton’s equations (8.1) are
q˙ = p , p˙ = q − q^3. (8.26)
For 1-dof systems, the ‘surfaces’ of constant energy (8.5) are curves that stratify the phase
plane (q, p), and the dynamics is very simple: the curves of constant energy are the tra- jectories, as shown in figure 8.1. click to return: p. 139
Example 8.2. Collinear helium. In the quantum chaos part of ChaosBook.org we shall apply the periodic orbit theory to the quantization of helium. In particular, we will study collinear helium, a doubly charged nucleus with two electrons arranged on a line, an electron on each side of the nucleus. The Hamiltonian for this system is
p^21 +
p^22 −
r 1
r 2
r 1 + r 2
Collinear helium has 2 dof, and thus a 4-dimensional phase space M, which energy con-
servation stratified by 3-dimensional constant energy hypersurfaces. In order to visualize it, we often project the dynamics onto the 2-dimensional configuration plane, the (r 1 , r 2 ),
ri ≥ 0 quadrant, figure 8.2. It looks messy, and, indeed, it will turn out to be no less chaotic than a pinball bouncing between three disks. As always, a Poincaré section will
be more informative than this rather arbitrary projection of the flow. The difference is that
in such projection we see the flow from an arbitrary perspective, with trajectories criss- crossing. In a Poincaré section the flow is decomposed into intrinsic coordinates, a pair
along the marginal stability time and energy directions, and the rest transverse, revealing the phase-space structure of the flow. click to return: p. 140
Example 8.3. Symplectic form for D = 2. For two degrees of freedom the phase space is 4-dimensional, x = (q 1 , q 2 , p 1 , p 2 ) , and the symplectic 2-form is
ω =
The symplectic bilinear form 〈x(1)|x(2)〉 is the sum over the areas of the parallelepipeds spanned pairwise by components of the two vectors,
〈x(1)|x(2)〉 = (x(1))>ω x(2)^ = (q(1) 1 p(2) 1 − q(2) 1 p(1) 1 ) + (q(1) 2 p(2) 2 − q(2) 2 p(1) 2 ). (8.29)
It is this sum over oriented areas (not the Euclidean distance between the two vectors, |x(2)^ − x(1)|) that is preserved by the symplectic transformations. click to return: p. 141
Example 8.4. Hamiltonian flows are canonical. For Hamiltonian flows it follows from (8.14) that (^) dtd
J>ωJ
= 0, and since at the initial time J^0 (x 0 ) = 1 , Jacobian matrix is a symplectic transformation (8.6). This equality is valid for all times, so a Hamiltonian flow f t(x) is a canonical transformation, with the linearization ∂x f t(x) a symplectic trans- formation (8.6): For notational brevity here we have suppressed the dependence on time
Figure 8.4: Stability exponents of a Hamiltonian equi- librium point, 2-dof.
complex saddle saddle-center
degenerate saddle
(2) (2)
real saddle
generic center degenerate center
(2) (2)
and the initial point, J = Jt(x 0 ). By elementary properties of determinants it follows from (8.6) that Hamiltonian flows are phase-space volume preserving, |det J| = 1. The initial condition (4.10) for J is J^0 = 1 , so one always has
det J = + 1. (8.30)
click to return: p. 143
Example 8.5. Hamiltonian Hénon map, reversibility. By (4.45) the Hénon map (3.18) for b = −1 value is the simplest 2-dimensional orientation preserving area- preserving map, often studied to better understand topology and symmetries of Poincaré sections of 2 dof Hamiltonian flows. We find it convenient to multiply (3.19) by a and absorb the a factor into x in order to bring the Hénon map for the b = −1 parameter value into the form
xi+ 1 + xi− 1 = a − x^2 i , i = 1 , ..., np , (8.31)
The 2-dimensional Hénon map for b = −1 parameter value
xn+ 1 = a − x^2 n − yn yn+ 1 = xn. (8.32)
is Hamiltonian (symplectic) in the sense that it preserves area in the [x, y] plane.
For definitiveness, in numerical calculations in examples to follow we shall fix (arbi- trarily) the stretching parameter value to a = 6, a value large enough to guarantee that all
roots of 0 = f n(x) − x (periodic points) are real. exercise 9. click to return: p. 144 Example 8.6. 2-dimensional symplectic maps. In the 2-dimensional case the eigen- values (5.5) depend only on tr Mt
Λ 1 , 2 =
tr Mt^ ±
(tr Mt^ − 2)(tr Mt^ + 2)
Greene’s residue criterion states that the orbit is (i) elliptic if the stability residue |tr Mt| − 2 ≤ 0, with complex eigenvalues Λ 1 = eiθt, Λ 2 = Λ∗ 1 = e−iθt. If |tr Mt| − 2 > 0, λ is real, and the trajectory is either
(ii) hyperbolic Λ 1 = eλt^ , Λ 2 = e−λt^ , or (8.34) (iii) inverse hyperbolic Λ 1 = −eλt^ , Λ 2 = −e−λt^. (8.35)
click to return: p. 144
and
G′(x) = −g(x). (8.40)
Important features of this map, including transition to global chaos (destruction of the last invariant torus), may be tackled by detailed investigation of the stability of periodic orbits. A family of periodic orbits of period Q already present in the k = 0 rotation maps can be labeled by its winding number P/Q The Greene residue describes the stability of a P/Q-cycle:
2 − tr MP/Q
If RP/Q ∈ (0, 1) the orbit is elliptic, for RP/Q > 1 the orbit is hyperbolic orbits, and for RP/Q < 0 inverse hyperbolic.
For k = 0 all points on the y 0 = P/Q line are periodic with period Q, winding number P/Q and marginal stability RP/Q = 0. As soon as k > 0, only a 2Q of such orbits survive, according to Poincaré -Birkhoff theorem: half of them elliptic, and half hyperbolic. If
we further vary k in such a way that the residue of the elliptic Q-cycle goes through 1, a bifurcation takes place, and two or more periodic orbits of higher period are generated. click to return: p. 144
Exercises
8.1. Complex nonlinear Schrödinger equation. Con- sider the complex nonlinear Schrödinger equation in one spatial dimension [17]:
i
∂φ ∂t
∂^2 φ ∂x^2
(a) Show that the function ψ : R → C defining the traveling wave solution φ(x, t) = ψ(x−ct) for c > 0 satisfies a second-order complex differential equa- tion equivalent to a Hamiltonian system in R^4 rel- ative to the noncanonical symplectic form whose matrix is given by
wc =
− 1 0 0 −c 0 − 1 c 0
(b) Analyze the equilibria of the resulting Ha- miltonian system in R^4 and determine their linear stability properties. (c) Let ψ(s) = eics/^2 a(s) for a real function a(s) and determine a second order equation for a(s). Show
that the resulting equation is Hamiltonian and has heteroclinic orbits for β < 0. Find them. (d) Find ‘soliton’ solutions for the complex nonlinear Schrödinger equation.
(Luz V. Vela-Arevalo)
8.2. Symplectic vs. Hamiltonian matrices. In the language of group theory, symplectic matrices form the symplectic Lie group Sp(d), while the Hamiltonian ma- trices form the symplectic Lie algebra sp(d), or the al- gebra of generators of infinitesimal symplectic transfor- mations. This exercise illustrates the relation between the two:
(a) Show that if a constant matrix A satisfy the Hamil- tonian matrix condition (8.9), then J(t) = exp(tA) , t ∈ R, satisfies the symplectic condition (8.6), i.e., J(t) is a symplectic matrix. (b) Show that if matrices Ta satisfy the Hamiltonian matrix condition (8.9), then g(φ) = exp(φ · T) , φ ∈ RN^ , satisfies the symplectic condition (8.6), i.e., g(φ) is a symplectic matrix.
exerNewton - 13jun2008 ChaosBook.org edition16.0, Jan 13 2018
(A few hints: (i) expand exp(A) , A = φ · T , as a power series in A. Or, (ii) use the linearized evolution equation (8.13). ) 8.3. When is a linear transformation canonical?
(a) Let A be a [n × n] invertible matrix. Show that the map φ : R^2 n^ → R^2 n^ given by (q, p) 7 → (Aq, (A−^1 )>p) is a canonical transformation. (b) If R is a rotation in R^3 , show that the map (q, p) 7 → (R q, R p) is a canonical transformation.
(Luz V. Vela-Arevalo) 8.4. Determinants of symplectic matrices. Show that the determinant of a symplectic matrix is +1, by going through the following steps:
(a) use (8.21) to prove that for eigenvalue pairs each member has the same multiplicity (the same holds for quartet members), (b) prove that the joint multiplicity of λ = ±1 is even, (c) show that the multiplicities of λ = 1 and λ = − 1 cannot be both odd. Hint: write
P(λ) = (λ − 1)^2 m+^1 (λ + 1)^2 l+^1 Q(λ)
and show that Q(1) = 0.
8.5. Cherry’s example. What follows refs. [2, 20] is mostly a reading exercise, about a Hamiltonian system that is linearly stable but nonlinearly unstable. Consider the Hamiltonian system on R^4 given by
(q^21 + p^21 ) − (q^22 + p^22 ) +
p 2 (p^21 − q^21 ) − q 1 q 2 p 1.
(a) Show that this system has an equilibrium at the origin, which is linearly stable. (The linearized system consists of two uncoupled oscillators with frequencies in ratios 2:1). (b) Convince yourself that the following is a family of solutions parameterize by a constant τ:
q 1 = −
cos(t − τ) t − τ
, q 2 = cos 2(t − τ) t − τ
p 1 =
sin(t − τ) t − τ
, p 2 =
sin 2(t − τ) t − τ
These solutions clearly blow up in a finite time; however they start at t = 0 at a distance
3 /τ from the origin, so by choosing τ large, we can find solutions starting arbitrarily close to the ori- gin, yet going to infinity in a finite time, so the origin is nonlinearly unstable.
(Luz V. Vela-Arevalo)
exerNewton - 13jun2008 ChaosBook.org edition16.0, Jan 13 2018