You are on page 1of 12

10/22/2014

Hamiltonian systems - Scholarpedia

Hamiltonian systems
Ja m es Meiss (2 0 0 7 ), Sch ola r pedia ,
2 (8 ):1 9 4 3 .

Recommend this on Google

doi:1 0 .4 2 4 9 /sch ola r pedia .1 9 4 3

r ev ision #1 2 9 9 2 5 [lin k t o/cit e t h is


a r t icle]

Prof. James Meiss, Applied Mathematics University of Colorado, Boulder, CO, USA
A dynamical system of 2n , first order, ordinary differential equations
= J H(z, t),
z

J = (

) ,

is an n degree-of-freedom (d.o.f.) Hamiltonian system (when it is nonautonomous it has n

(1)

+ 1/2

d.o.f.).

Here H is the ''Hamiltonian'', a smooth scalar function of the extended phase space variables z and time t , the
2n 2n matrix J is the Poisson matrix and I is the n n identity matrix. The equations naturally split into
two sets of n equations for canonically conjugate variables, z = (q, p) , i.e.
= H/p,
q

= H/q .
p

(2)

Here the n coordinates q represent the configuration variables of the system (e.g. positions of the component
parts) and their canonically conjugate momenta p represent the impetus gained by movement.
Hamiltonian systems are universally used as models for virtually all of physics.

Contents
1 Formulation
2 Examples
2.1 Springs
2.2 Pendulum
2.3 N-body problem
2.4 Electromagnetic Forces
3 Geometric Structure
3.1 Conservation of Energy
3.2 Liouville's Theorem
3.3 Poincar's Invariant
3.4 Symplectic Maps
4 Integrable Systems
4.1 Action-Angle Variables
4.2 Liouville Integrability
5 KAM Theory
6 Hamiltonian Chaos
7 References
8 External Links
http://www.scholarpedia.org/article/Hamiltonian_systems

1/12

10/22/2014

Hamiltonian systems - Scholarpedia

9 See Also

Formulation
In 1834 William Rowan Hamilton showed that Newton's equations F = ma for a set of particles in a
conservative force field F = V with "potential energy" V could be derived from a single function that he
called the "characteristic function",
n

H(q, p) =
i=1

|pi |

+ V (q , q , , q
1

2mi

) .

(3)

Here q is the position of the i particle whose mass is m , and p is its canonical momentum p = m q . The
equations of motion are obtained by (2), which can in turn be converted to Newton's second order form by
differentiating the equation q = p /m .
th

At first it seems that Hamilton's formulation gives only a convenient restatement of Newton's system---the
convenience perhaps most evident in that the scalar function H(q, p) encodes all of the information of the 2n
first order dynamical equations. However, a Hamiltonian formulation gives much more than just this
simplification. Indeed, if we allow more general functions H(q, p, t) and a more general relationship between
the canonical momenta and the velocities q then virtually all of the models of classical physics have a
Hamiltonian formulation, including electromagnetic forces, which are not derivable from a (scalar) potential.
Moreover, waves in inviscid fluids such as surface water waves or magnetohydrodynamic waves also have a
Hamiltonian (PDE) formulation. Quantum mechanics is formally obtained from classical mechanics by
replacing the canonical momentum in the Hamiltonian by a differential operator.
Hamiltonian structure provides strong constraints on the flow. Most simply, when H does not depend upon
time (autonomous) then its value is constant along trajectories: the energy E = H(q, p) is constant, see Energy
Conservation. Similarly if the Hamiltonian is independent of one of the configuration variables (the variable is
ignorable), then (2) implies that the corresponding canonical momentum is an invariant. This gives a simple
explanation for the relation between symmetries (for example rotational symmetry) and invariants (for
example angular momentum)---see Noether's Theorem.
One of the stronger constraints imposed by Hamiltonian structure relates to stability: it is impossible for a
trajectory to be asymptotically stable in a Hamiltonian system. Even more structure applies: for each
eigenvalue of an equilibrium there is a corresponding opposite eigenvalue . For example an equilibrium of
a one degree-of-freedom system must either be a center (two imaginary eigenvalues, i or a saddle (two real
eigenvalues, ) or have a double zero eigenvalue.
Another geometric implication is that knowledge of n invariants is enough to fully characterize a solution of the
2n equations for an n degree-of-freedom system, i.e., the Hamiltonian is integrable. This follows from
Liouville's Integrability Theorem. Moreover, if the orbits of such a system are bounded, then almost all of them
must lie on n-dimensional tori. Kolmogorov, Arnold and Moser proved that a sufficiently smooth, nearlyintegrable Hamiltonian system still has many such invariant tori (see KAM theory). This strong structural
stability of Hamiltonian dynamics was unexpected even in the middle of the 20 century when physicists
began the first computer simulations of dynamical systems (see Fermi Pasta Ulam problem).
th

http://www.scholarpedia.org/article/Hamiltonian_systems

2/12

10/22/2014

Hamiltonian systems - Scholarpedia

Examples
For many mechanical systems, the Hamiltonian takes the form H(q, p) = T (q, p) + V (q) , where T (q, p) is the
kinetic energy, and V (q) is the potential energy of the system. Such systems are called natural Hamiltonian
systems. The simplest case is when the kinetic energy is of the form in (3) for a set of particles with kinetic
momenta p R and masses m . More generally, when the extent of the bodies is taken into account the
kinetic energy can depend upon the configuration of the system, but it is typically a quadratic function of the
3

momenta, so that T (q, p) =

1
2

M (q)

p ,

where the n

as the inertia of the system, and the vector p

mass matrix M (q) represents the shape as well

includes both linear momenta and angular momenta.

Springs
A harmonic spring has potential energy of the form

where k is

the spring's force coefficient (the force per unit length of extension) or
the spring constant, and x is the length of the spring relative to its
unstressed, natural length. Thus a point particle of mass m connected
to a harmonic spring with natural length L that is attached to a fixed
support at the origin and allowed to move in one dimension has a
Hamiltonian of the form H(q, p) =
p +
(q L) and thus its
1

2m

Figure 1 : Coupled Springs

equations of motion are


= p/m ,
q

= k(q L) .
p

If the spring is hanging vertically in a constant gravitational field, then the new equations are obtained by
simply adding the gravitational potential energy mgq to H .
A set of point masses that are coupled by springs has potential energy given by the sum of the potential energies
of each spring in the system. For example suppose that there are two masses connected to three springs as
shown in (Figure 1). The Hamiltonian is
1
H(q, p) =

p
m1

1
+

p
m2

k1

2
1

k2
2

(q 2 q 1 )

k3
2

(L q 2 )

One advantage of the Hamiltonian formulation of mechanics is that the equations for arbitrarily complicated
arrays of springs and masses can be obtained by simply finding the expression for the total energy of the system
(However, it is often easier to do this using the Lagrangian formulation of mechanics which does not require
knowing the form of the canonical momenta in advance).

Pendulum
The ideal, planar pendulum is a particle of mass m in a constant gravitational field, that is attached to a rigid,
massless rod of length L , as shown in (Figure 2). The canonical momentum of this system is the angular
momentum p = mL and the potential energy is the gravitational energy
from the vertical. The Hamiltonian is
2

mgL cos ,

where is the angle

H(, p) =

2mL

http://www.scholarpedia.org/article/Hamiltonian_systems

mgL cos .

(4)

3/12

10/22/2014

Hamiltonian systems - Scholarpedia

This gives the equations of motion


p

mL

= mgL sin .
p

While these equations are simple, their explicit solution requires elliptic functions. However, the trajectories of
the pendulum are easy to visualize since the energy is conserved, see (Figure 3). When
the energy is below mgL the angle cannot exceed and the pendulum oscillates. Since
the energy is conserved, the orbit must be periodic. For energies larger than mgL , the
pendulum rotates, and the angle either monotonically grows with time (if the angular
momentum is positive) or decreases (negative p). The critical level set is the separatrix;
the two orbits on this level set asymptotically approach the equilibrium (, 0) as
t .

These are called homoclinic orbits.

N-body problem
A set of point masses interacting by Newton's gravitational force is also a Hamiltonian
system of the natural form (3) with potential energy

Figure 2:
Planar
Pendulum

Gmi mj
V (q 1 , q n ) =
i<j

where q

||q i q j ||

is the position of the i

th

body. In

addition to the conserved energy H = E this system


has additional conserved quantities. Since H is a
function only of the difference between particle
positions, the total momentum
n

P = pi

(5)

i=1

Figure 3: Phase Space of the Pendulum

is conserved. Since H is a function only of the distance


between the bodies, the total angular momentum is
also conserved
n

L = q

i=1

For the case of two bodies, the Hamiltonian has six degrees of freedom (the three components of the position
and momentum for each body), however, the conservation of total momentum means that if we choose
coordinates moving with the center of mass

http://www.scholarpedia.org/article/Hamiltonian_systems
Q =

4/12
i

10/22/2014

Hamiltonian systems - Scholarpedia


n

mi q
i

Q =
M

where M

= mi

i=1

is the total mass, then the Hamiltonian is independent of Q , so that its conjugate

momentum (5) is constant. Thus the system is reduced to three degrees of freedom, depending only upon the
inter-particle vector q = q q , and its conjugate momentum, p = q , where =
is the reduced
m1 m2

mass. In these coordinates the Hamiltonian becomes


P

H(q, Q, p, P ) =

2M

Gm1 m2

(6)

||q||

The total angular momentum splits as well L = Q P + q p . Since P is constant, and Q


= M P , the first
term is itself individually conserved, so l = q p is also constant, a fact that can also be seen from (6) directly.

The dynamics of three or more bodies can be extremely complex.

Electromagnetic Forces
A nonrelativistic charged particle in an electromagnetic field has the equations of motion
= eE(q, t) +
mq

B(q, t)
q

where E is the electric field, and B is the magnetic field and we use Gaussian (cgs) units. This system is
Hamiltonian, with
1

e
(p

H(q, p, t) =
2m

A)

+ e,

(7)

where the scalar and vector potentials and A are defined through
A
E = +

B = A.

The momentum occurring in (7) is not the kinetic momentum mq , but rather a canonical momentum defined
by p = mq + A . For systems that also have a Lagrangian formulation, the canonical momentum is defined
e
c

by
)
L(q, q
p =

Note that the first term in the Hamiltonian (7) is simply the kinetic energy as usual, and the last term is the
electrical potential energy.

Geometric Structure
http://www.scholarpedia.org/article/Hamiltonian_systems

5/12

10/22/2014

Hamiltonian systems - Scholarpedia

Much of the elegance of the Hamiltonian formulation stems from its geometric structure. Hamiltonian phase
space is an even dimensional space with a natural splitting into two sets of coordinates, the configuration
variables q and the momenta p . For most physical systems the momenta are similar to velocities, which are
tangent vectors to trajectories, but the difference--emphasized in the electromagnetic example--is that they are
cotangent vectors, as we will explain further below. In this case the Hamiltonian phase space is the cotangent
bundle of the configuration space.
More abstractly, the phase space of a Hamiltonian system is an even dimensional manifold M that is endowed
with a nondegenerate two-form, . This two-form allows us to define a pairing between vectors and covectors.
Given a Hamiltonian function H

: M R ,

the Hamiltonian vector field z = X(z) is defined by


i X (X, ) = dH.

(8)

This is just a coordinate-free version of (1). Indeed, a famous theorem of Darboux implies that near each point
in M there exists a set of canonical variables z = (q, p) , such that
= dq dp ,

where is the "wedge product". In terms of these coordinates, (v, w) = v

Jw ,

where J is the Poisson

matrix (1), and the equations (8) become


J

X = H,

which is a restatement of (1).

Conservation of Energy
If a Hamiltonian does not depend explicitly on time, then its value, the energy, is constant. Indeed
differentiating along a trajectory gives
dH

dq

=
dt

dp

dt

+
q

by (2). Thus H(q(t), p(t)) = H(q(0), p(0)) = E

dt

= 0,

While Hamiltonian systems are often referred to as conservative systems, these two types of dynamical
systems should not be confounded. In the autonomous case, a Hamiltonian system conserves energy, however,
it is easy to construct nonHamiltonian systems that also conserve an energy-like quantity. Moreover, in the
nonautonomous case, the Hamiltonian depends explicitly on time H(q, p, t) and there is no conserved energy.

Liouville's Theorem
One direct consequence of the form (2) is that the divergence of a Hamiltonian vector field is zero
H
X = J H = J ij
i,j

since J is antisymmetric and the Hessian matrix

D H

= 0.
z i z j

is symmetric. This immediately implies that the volume

of any bundle of trajectories is preserved. That is, suppose A is a set of initial conditions with volume
http://www.scholarpedia.org/article/Hamiltonian_systems

V (A) =

dz.

6/12

10/22/2014

Hamiltonian systems - Scholarpedia


V (A) =

dz.
A

If z evolves to

(z) ,

the flow of the vector field, then the new volume

dz =
(A)

(t (z) y)dy,
A

is the same as the original volume V (A) .


This is known as Liouville's theorem. It is valid for any divergence free vector field, X = 0 .
Note that Hamiltonian flow is volume preserving even when it is nonautonomous.

Poincar's Invariant
In addition to preserving volume, Hamiltonian systems also preserve a loop action, or Poincar invariant.
Given any loop L in the extended phase space (q, p, t) , let
A(L) =

pdq H(q, p, t)dt.

(9)

Then under a Hamiltonian flow the loop action is preserved


A( (L)) = A(L).
t

Even more generally, suppose T is the two dimensional tube obtained from the flow of L
T = {t (L) : t R}

and L is any loop on T that is homotopic to L . Then A(L ) = A(L) .

This fact is used, for example, in the construction of a Poincar section for Hamiltonian systems.

Symplectic Maps
A map f

: M M

is symplectic if it preserves the symplectic form . Geometrically, we say that f

= ,

which becomes in components


Df

where Df is the 2n

2n

J Df = J

(10)

Jacobian matrix

q
Df (q, p) =

f p

p
f p
p

The preservation of the loop action (9) implies that the time-T map of any Hamiltonian flow is symplectic. This
follows from Stokes's theorem and the fact that for a loop at a fixed value of time, the loop action reduces to

pdq .

Note that this holds even if the Hamiltonian depends explicitly on time H(q, p, t) .

http://www.scholarpedia.org/article/Hamiltonian_systems

7/12

10/22/2014

Hamiltonian systems - Scholarpedia

Another way in which symplectic maps arise is for Poincar sections of autonomous Hamiltonian flows on an
energy surface. For example if the surface Q

= {(q, p) : q n = 0, q
> 0, H(q, p) = E}
n

resulting return map to Q is symplectic with the form |

n1
i=1

d q i d pi .

is selected, then the

This is especially useful for the

visualization of the motion of a two-degree-of-freedom system, since the resulting map is two-dimensional.
The set of linear mappings that obey (10) is called the symplectic group; it is a Lie group. Any quadratic
Hamiltonian
1
H(z) =

Kz ,

where K is a (constant) symmetric matrix, has a linear flow that is generated by the exponential (t) = e

tJ K

Each of the matrices in the curve (t) is symplectic. Indeed, the collection {J K

: K

= K}

forms the Lie

Algebra of the symplectic group.

Integrable Systems
A dynamical system is integrable when it can be solved in some way. One (rather restrictive) way in which this
can happen is if the flow of the vector field can be constructed analytically. However, since this can almost
never be done (in terms of elementary functions), this is not an especially useful class of systems.
However, there is a class of Hamiltonian systems, action-angle systems, whose solutions can be obtained
analytically, and there is a well-accepted definition of integrability for Hamiltonian dynamics due to Liouville in
which each integrable Hamiltonian is (locally) equivalent to these action-angle systems.

Action-Angle Variables
A Hamiltonian system is written in action-angle form if there is a set of canonical variables (, I ) where
and I R and such that H depends only upon the actions

H(I ) .

In this case the equations of motion (1) become simple indeed:

= H(I ) = (I ) ,

I = 0

(11)

These equations can be easily solved, giving


((t), I (t)) = ( o + (I o )t, I o )

Thus the angles move along the invariant torus I

= Io

with a fixed frequency vector

For example, the simple harmonic oscillator Hamiltonian


1
H(q, p) =

(p

+ q

http://www.scholarpedia.org/article/Hamiltonian_systems

(q, p) = (

sin ,

cos ) .

8/12

10/22/2014

Hamiltonian systems - Scholarpedia

can be written in action angle form by setting (q, p) = (2I


canonical since dq

dp = d dI

sin , 2I cos ) .

The new variables are

(i.e., the transformation is canonical). In the new coordinates the

Hamiltonian becomes H(, I ) = I . Thus it is in action-angle form with

= 1 .

A more general, anharmonic

oscillator, with a natural Hamiltonian of the form (3) with a potential energy V (q) with a unique minimum at
q = 0 has a Hamiltonian that depends in a nonlinear way upon the action, but which nevertheless can be
reduced to action-angle form.
Hamiltonian systems with two or more degrees of freedom cannot always be reduced to action-angle form,
giving rise to chaotic motion.

Liouville Integrability
Liouville and Arnold showed that the motion in a larger class of Hamiltonian systems is as simple as that of (11).
Suppose that an n degree-of-freedom Hamiltonian system (2) has a set of n invariants F that are almost
i

everywhere independent (their gradients span an n-dimensional space except on sets of zero measure) and that
are in involution, that is, their Poisson brackets vanish:
{F i , F j } (F i , F j ) = 0 .

Then if a regular level set of the invariants L

= {F i (q, p) = ci : i = 1, n}

is compact it must be a torus .

Moreover, there is a neighborhood of L in which there exist action-angle coordinates such that the equations
c

of motion reduce to (11). See (Arnold, 1978).


For example, every one degree-of-freedom, autonomous Hamiltonian system is Liouville integrable. However,
the action-angle coordinates may not be globally defined. In the case of the pendulum (4), there are actionvariables away from the separatrix.
Generically the dynamics on an invariant torus are quasiperiodic.

KAM Theory
Andrey Kolmogorov discovered a general method for the study of perturbed, integrable Hamiltonian systems.
The method lead to theorems by Vladimir Arnold for analytic Hamiltonian systems (Arnold, 1963) and by
Jurgen Moser for smooth enough area-preserving mappings (Moser 1962), and the ideas have become known as
KAM theory.
Roughly speaking, KAM theory implies that a Hamiltonian system of the form
H(, I ) = H 0 (I ) + H 1 (, I ) ,

which is integrable at = 0 , still has a large set of invariant tori if is small enough (a set whose measure
approaches the total measure as 0 ). In order that KAM theory apply, the Hamiltonian must be sufficiently
smooth, and (for the simplest version of the theorem) the unperturbed Hamiltonian must satisfy a
nondegeneracy or twist condition, that D

H 0 (I )

is nonsingular.

For more details see Kolmogorov-Arnold-Moser Theory.

http://www.scholarpedia.org/article/Hamiltonian_systems

9/12

10/22/2014

Hamiltonian systems - Scholarpedia

Hamiltonian Chaos
Though many invariant tori of an
integrable system typically persist
upon a perturbation, tori that
commensurate or nearly
commensurate are typically
destroyed. Chaotic dynamics often
occurs in the neighborhood of these
destroyed tori.
An invariant torus is characterized by
its frequency vector . It is
commensurate if there exists an
integer vector m

such that

m = 0

Figure 4: Poincar section at t = 2k of the two-wav e


Hamiltonian (1 2) for a = 4 , b = 6 and = 0.1 . Resonant
island chains with rotation numbers 0/1, 1/2, 2/3, 4/5 and 1/1
are shown.

Commensurate tori of an integrable


system are generically destroyed by
any perturbation.
For example, consider the 1.5 degreeof-freedom system
1
H(q, p) =

+ (acos(2q) + bcos(2(q t))

(12)

that represents the motion of (for example) a charge particle in the field of two electrostatic waves. Here the
phase space can be taken to be T

since H is a periodic function of q and t For = 0 , the momentum is

constant and the orbits lie on two-dimensional tori with the frequency vector

= (p, 1)

Consequently every

torus with a rational value of p is commensurate--indeed such orbits are periodic in this case.
KAM theory implies that if p is "sufficiently" irrational, then the torus is preserved for || 1 . However
commensurate tori and nearby irrational tori are destroyed. For small the destroyed tori are replaced by
chains of islands formed from a pair of periodic orbits, one a saddle and the other elliptic (see Stability of
Hamiltonian Flows). Surrounding the elliptic orbit are a family of two-dimensional tori with a new topology (not
homotopic to p = constant , see (Figure 4). Moreover, the stable and unstable manifolds of the saddle
typically intersect transversely, giving rise to a Smale horseshoe and chaotic motion (albeit chaos that is limited
to a narrow layer about the separatrix. As grows these chaotic layers also grow, and they can envelope larger
regions of phase space, see (Figure 5).

References
Abraham, R. and J. E. Marsden (1978). Foundations of Mechanics. Reading, Benjamin.
http://www.scholarpedia.org/article/Hamiltonian_systems

10/12

10/22/2014

Hamiltonian systems - Scholarpedia

Arnold, V. I. (1963). Proof of a Theorem of A.N. Kolmogorov on the Invariance of Quasiperiodic Motions
Under Small Perturbations of the Hamiltonian. Russ. Math. Surveys 18:5: 9-36.
Arnold, V. I. (1978). Mathematical Methods of Classical Mechanics. New York, Springer.
MacKay, R. S. and J. D. Meiss, Eds.
(1987). Hamiltonian Dynamical
Systems: a reprint selection. London,
Adam-Hilgar Press.
McDuff, D. and D. Salamon (1995).
Introduction to Symplectic Topology.
Oxford, Clarendon Press.
Meyer, K. R. and G. R. Hall (1992).
Introduction to the Theory of
Hamiltonian Systems. New York,
Springer-Verlag.
Moser, J. K. (1962). On Invariant
Curves of Area-Preserving Mappings
of an Annulus. Nachr. Akad. Wiss.
Gttingen, II Math. Phys. 1: 1-20.

Figure 5: Poincar section of the two-wav e Hamiltonian (1 2) for


= 0.2 .

Siegel, C. L. and J. K. Moser (1971).


Lectures on Celestial Mechanics. New
York, Springer-Verlag.
Internal references
Paul M.B. Vitanyi (2007) Andrey Nikolaevich Kolmogorov. Scholarpedia, 2(2):2798.
Yuri A. Kuznetsov (2007) Conjugate maps. Scholarpedia, 2(12):5420.
James Meiss (2007) Dynamical systems. Scholarpedia, 2(2):1629.
Eugene M. Izhikevich (2007) Equilibrium. Scholarpedia, 2(10):2014.
Ferdinand Verhulst (2007) Hamiltonian normal forms. Scholarpedia, 2(8):2101.
Jeff Moehlis, Kresimir Josic, Eric T. Shea-Brown (2006) Periodic orbit. Scholarpedia, 1(7):1358.
Anatoly M. Samoilenko (2007) Quasiperiodic oscillations. Scholarpedia, 2(5):1783.
Steve Smale and Michael Shub (2007) Smale horseshoe. Scholarpedia, 2(11):3012.
Philip Holmes and Eric T. Shea-Brown (2006) Stability. Scholarpedia, 1(10):1838.
Alain Chenciner (2007) Three body problem. Scholarpedia, 2(10):2111.

External Links
James Meiss's website (http://amath.colorado.edu/faculty/jdm/)

See Also

http://www.scholarpedia.org/article/Hamiltonian_systems

11/12

10/22/2014

Hamiltonian systems - Scholarpedia

Dynamical systems, Ordinary differential equations, Stability of Hamiltonian flows, Hamiltonian normal forms,
Kolmogorov-Arnold-Moser theory, Resonance, Melnikov function, Symplectic maps, Nontwist maps, Three
body problem, Quasiperiodic oscillations, Andrey Kolmogorov, Henri Poincar
Sponsored by: Eugene M. Izhikevich, Editor-in-Chief of Scholarpedia, the peer-reviewed open-access
encyclopedia
Reviewed by (http://www.scholarpedia.org/w/index.php?title=Hamiltonian_systems&oldid=18653) :
Anonymous
Accepted on: 2007-08-19 16:06:03 GMT (http://www.scholarpedia.org/w/index.php?
title=Hamiltonian_systems&oldid=18653)
Categories:

Dynamical Systems Celestial mechanics


This page w as las t m odified
on 15 January 2013, at 04:10.
This page has been acces s ed
80,209 tim es .
"Ham iltonian s ys tem s " by
Jam es Meis s is licens ed
under a Creative Com m ons
Attribution-NonCom m ercialShareAlik e 3.0 Unported
Licens e. Perm is s ions beyond
the s cope of this licens e are
des cribed in the Term s of
Us e

http://www.scholarpedia.org/article/Hamiltonian_systems

12/12

You might also like