You are on page 1of 37

Canonical

transformation

In Hamiltonian mechanics, a canonical


transformation is a change of canonical
coordinates (q, p, t) → (Q, P, t) that
preserves the form of Hamilton's
equations. This is sometimes known as
form invariance. It need not preserve the
form of the Hamiltonian itself. Canonical
transformations are useful in their own
right, and also form the basis for the
Hamilton–Jacobi equations (a useful
method for calculating conserved
quantities) and Liouville's theorem (itself
the basis for classical statistical
mechanics).

Since Lagrangian mechanics is based on


generalized coordinates, transformations
of the coordinates q → Q do not affect the
form of Lagrange's equations and, hence,
do not affect the form of Hamilton's
equations if we simultaneously change the
momentum by a Legendre transformation
into
Therefore, coordinate transformations
(also called point transformations) are a
type of canonical transformation. However,
the class of canonical transformations is
much broader, since the old generalized
coordinates, momenta and even time may
be combined to form the new generalized
coordinates and momenta. Canonical
transformations that do not include the
time explicitly are called restricted
canonical transformations (many
textbooks consider only this type).
For clarity, we restrict the presentation
here to calculus and classical mechanics.
Readers familiar with more advanced
mathematics such as cotangent bundles,
exterior derivatives and symplectic
manifolds should read the related
symplectomorphism article. (Canonical
transformations are a special case of a
symplectomorphism.) However, a brief
introduction to the modern mathematical
description is included at the end of this
article.

Notation
Boldface variables such as q represent a
list of N generalized coordinates that need
not transform like a vector under rotation,
e.g.,

A dot over a variable or list signifies the


time derivative, e.g.,

The dot product notation between two


lists of the same number of coordinates is
a shorthand for the sum of the products of
corresponding components, e.g.,
The dot product (also known as an "inner
product") maps the two coordinate lists
into one variable representing a single
numerical value.

Direct approach
The functional form of Hamilton's
equations is
By definition, the transformed coordinates
have analogous dynamics

where K(Q, P) is a new Hamiltonian


(sometimes called the Kamiltonian[1]) that
must be determined.

In general, a transformation
(q, p, t) → (Q, P, t) does not preserve the
form of Hamilton's equations. For time
independent transformations between
(q, p) and (Q, P) we may check if the
transformation is restricted canonical, as
follows. Since restricted transformations
have no explicit time dependence (by
definition), the time derivative of a new
generalized coordinate Qm is

where {⋅, ⋅} is the Poisson bracket.

We also have the identity for the conjugate


momentum Pm
If the transformation is canonical, these
two must be equal, resulting in the
equations

The analogous argument for the


generalized momenta Pm leads to two
other sets of equations
These are the direct conditions to check
whether a given transformation is
canonical.

Liouville's theorem
The direct conditions allow us to prove
Liouville's theorem, which states that the
volume in phase space is conserved under
canonical transformations, i.e.,
By calculus, the latter integral must equal
the former times the Jacobian J

where the Jacobian is the determinant of


the matrix of partial derivatives, which we
write as

Exploiting the "division" property of


Jacobians yields
Eliminating the repeated variables gives

Application of the direct conditions above


yields J = 1.

Generating function
approach
To guarantee a valid transformation
between (q, p, H) and (Q, P, K), we may
resort to an indirect generating function
approach. Both sets of variables must
obey Hamilton's principle. That is the
Action Integral over the Lagrangian
and

respectively, obtained by the Hamiltonian


via ("inverse") Legendre transformation,
both must be stationary (so that one can
use the Euler–Lagrange equations to
arrive at equations of the above-
mentioned and designated form; as it is
shown for example here):
One way for both variational integral
equalities to be satisfied is to have

Lagrangians are not unique: one can


always multiply by a constant λ and add a
total time derivative dG
dt and yield the same
equations of motion (see for reference:
[1] ).

In general, the scaling factor λ is set equal


to one; canonical transformations for
which λ ≠ 1 are called extended canonical
transformations. dG
dt is kept, otherwise the
problem would be rendered trivial and
there would be not much freedom for the
new canonical variables to differ from the
old ones.

Here G is a generating function of one old


canonical coordinate (q or p), one new
canonical coordinate (Q or P) and
(possibly) the time t. Thus, there are four
basic types of generating functions
(although it should be noted that mixtures
of these four types can exist), depending
on the choice of variables. As will be
shown below, the generating function will
define a transformation from old to new
canonical coordinates, and any such
transformation (q, p) → (Q, P) is
guaranteed to be canonical.

Type 1 generating function

The type 1 generating function G1 depends


only on the old and new generalized
coordinates

To derive the implicit transformation, we


expand the defining equation above

Since the new and old coordinates are


each independent, the following 2N + 1
equations must hold

These equations define the transformation


(q, p) → (Q, P) as follows. The first set of
N equations

define relations between the new


generalized coordinates Q and the old
canonical coordinates (q, p). Ideally, one
can invert these relations to obtain
formulae for each Qk as a function of the
old canonical coordinates. Substitution of
these formulae for the Q coordinates into
the second set of N equations

yields analogous formulae for the new


generalized momenta P in terms of the old
canonical coordinates (q, p). We then
invert both sets of formulae to obtain the
old canonical coordinates (q, p) as
functions of the new canonical
coordinates (Q, P). Substitution of the
inverted formulae into the final equation

yields a formula for K as a function of the


new canonical coordinates (Q, P).

In practice, this procedure is easier than it


sounds, because the generating function
is usually simple. For example, let

This results in swapping the generalized


coordinates for the momenta and vice
versa
and K = H. This example illustrates how
independent the coordinates and
momenta are in the Hamiltonian
formulation; they are equivalent variables.

Type 2 generating function

The type 2 generating function G2 depends


only on the old generalized coordinates
and the new generalized momenta
where the terms represent a
Legendre transformation to change the
right-hand side of the equation below. To
derive the implicit transformation, we
expand the defining equation above

Since the old coordinates and new


momenta are each independent, the
following 2N + 1 equations must hold
These equations define the transformation
(q, p) → (Q, P) as follows. The first set of
N equations

define relations between the new


generalized momenta P and the old
canonical coordinates (q, p). Ideally, one
can invert these relations to obtain
formulae for each Pk as a function of the
old canonical coordinates. Substitution of
these formulae for the P coordinates into
the second set of N equations

yields analogous formulae for the new


generalized coordinates Q in terms of the
old canonical coordinates (q, p). We then
invert both sets of formulae to obtain the
old canonical coordinates (q, p) as
functions of the new canonical
coordinates (Q, P). Substitution of the
inverted formulae into the final equation
yields a formula for K as a function of the
new canonical coordinates (Q, P).

In practice, this procedure is easier than it


sounds, because the generating function
is usually simple. For example, let

where g is a set of N functions. This


results in a point transformation of the
generalized coordinates
Type 3 generating function

The type 3 generating function G3 depends


only on the old generalized momenta and
the new generalized coordinates

where the terms represent a


Legendre transformation to change the
left-hand side of the equation below. To
derive the implicit transformation, we
expand the defining equation above
Since the new and old coordinates are
each independent, the following 2N + 1
equations must hold

These equations define the transformation


(q, p) → (Q, P) as follows. The first set of
N equations
define relations between the new
generalized coordinates Q and the old
canonical coordinates (q, p). Ideally, one
can invert these relations to obtain
formulae for each Qk as a function of the
old canonical coordinates. Substitution of
these formulae for the Q coordinates into
the second set of N equations

yields analogous formulae for the new


generalized momenta P in terms of the old
canonical coordinates (q, p). We then
invert both sets of formulae to obtain the
old canonical coordinates (q, p) as
functions of the new canonical
coordinates (Q, P). Substitution of the
inverted formulae into the final equation

yields a formula for K as a function of the


new canonical coordinates (Q, P).

In practice, this procedure is easier than it


sounds, because the generating function
is usually simple.

Type 4 generating function

The type 4 generating function


depends only on the old and
new generalized momenta

where the terms represent


a Legendre transformation to change both
sides of the equation below. To derive the
implicit transformation, we expand the
defining equation above

Since the new and old coordinates are


each independent, the following 2N + 1
equations must hold
These equations define the transformation
(q, p) → (Q, P) as follows. The first set of
N equations

define relations between the new


generalized momenta P and the old
canonical coordinates (q, p). Ideally, one
can invert these relations to obtain
formulae for each Pk as a function of the
old canonical coordinates. Substitution of
these formulae for the P coordinates into
the second set of N equations

yields analogous formulae for the new


generalized coordinates Q in terms of the
old canonical coordinates (q, p). We then
invert both sets of formulae to obtain the
old canonical coordinates (q, p) as
functions of the new canonical
coordinates (Q, P). Substitution of the
inverted formulae into the final equation
yields a formula for K as a function of the
new canonical coordinates (Q, P).

Motion as a canonical
transformation
Motion itself (or, equivalently, a shift in the
time origin) is a canonical transformation.
If and
, then Hamilton's
principle is automatically satisfied
since a valid trajectory
should always satisfy Hamilton's principle,
regardless of the endpoints.

Modern mathematical
description
In mathematical terms, canonical
coordinates are any coordinates on the
phase space (cotangent bundle) of the
system that allow the canonical one-form
to be written as

up to a total differential (exact form). The


change of variable between one set of
canonical coordinates and another is a
canonical transformation. The index of the
generalized coordinates q is written here
as a superscript ( ), not as a subscript as
done above ( ). The superscript conveys
the contravariant transformation
properties of the generalized coordinates,
and does not mean that the coordinate is
being raised to a power. Further details
may be found at the symplectomorphism
article.

History
The first major application of the
canonical transformation was in 1846, by
Charles Delaunay, in the study of the Earth-
Moon-Sun system. This work resulted in
the publication of a pair of large volumes
as Mémoires by the French Academy of
Sciences, in 1860 and 1867.

See also
Symplectomorphism
Hamilton–Jacobi equation
Liouville's theorem (Hamiltonian)
Mathieu transformation
Linear canonical transformation

References
1. Goldstein 1980, p. 380
Goldstein, Herbert (1980). Classical
mechanics (2d ed.). Reading, Mass.:
Addison-Wesley Pub. Co. p. 380. ISBN 0-
201-02918-9.
Landau, L. D.; Lifshitz, E. M. (1975)
[1939]. Mechanics. Translated by Bell, S.
J.; Sykes, J. B. (3rd ed.). Amsterdam:
Elsevier. ISBN 978-0-7506-28969.

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Canonical_transformation&oldid=880258833"

Last edited 2 months ago by Lolinat…


Content is available under CC BY-SA 3.0 unless
otherwise noted.

You might also like