415 lines
No EOL
21 KiB
Text
415 lines
No EOL
21 KiB
Text
|
|
§44
|
|
Maupertuis' principle
|
|
141
|
|
Writing the action in the form (43.8) and again replacing H by E, we have
|
|
(44.3)
|
|
The first term in this expression,
|
|
(44.4)
|
|
is sometimes called the abbreviated action.
|
|
Substituting (44.3) in (44.2), we find that
|
|
8S0=0.
|
|
(44.5)
|
|
Thus the abbreviated action has a minimum with respect to all paths which
|
|
satisfy the law of conservation of energy and pass through the final point
|
|
at any instant. In order to use such a variational principle, the momenta
|
|
(and so the whole integrand in (44.4)) must be expressed in terms of the
|
|
co-ordinates q and their differentials dq. To do this, we use the definition of
|
|
momentum:
|
|
(44.6)
|
|
and the law of conservation of energy:
|
|
E(g)
|
|
(44.7)
|
|
Expressing the differential dt in terms of the co-ordinates q and their differen-
|
|
tials dq by means of (44.7) and substituting in (44.6), we have the momenta
|
|
in terms of q and dq, with the energy E as a parameter. The variational prin-
|
|
ciple so obtained determines the path of the system, and is usually called
|
|
Maupertuis' principle, although its precise formulation is due to EULER and
|
|
LAGRANGE.
|
|
The above calculations may be carried out explicitly when the Lagrangian
|
|
takes its usual form (5.5) as the difference of the kinetic and potential energies:
|
|
The momenta are
|
|
and the energy is
|
|
The last equation gives
|
|
dt
|
|
(44.8)
|
|
142
|
|
The Canonical Equations
|
|
§44
|
|
substituting this in
|
|
Epides
|
|
we find the abbreviated action:
|
|
(44.9)
|
|
In particular, for a single particle the kinetic energy is T = 1/2 m(dl/dt)2,
|
|
where m is the mass of the particle and dl an element of its path; the variational
|
|
principle which determines the path is
|
|
${/[2m(B-U)]dl=0
|
|
(44.10)
|
|
where the integral is taken between two given points in space. This form is
|
|
due to JACOBI.
|
|
In free motion of the particle, U = 0, and (44.10) gives the trivial result
|
|
8 I dl = 0, i.e. the particle moves along the shortest path between the two
|
|
given points, i.e. in a straight line.
|
|
Let us return now to the expression (44.3) for the action and vary it with
|
|
respect to the parameter E. We have
|
|
substituting in (44.2), we obtain
|
|
(44.11)
|
|
When the abbreviated action has the form (44.9), this gives
|
|
=
|
|
(44.12)
|
|
which is just the integral of equation (44.8). Together with the equation of
|
|
the path, it entirely determines the motion.
|
|
PROBLEM
|
|
Derive the differential equation of the path from the variational principle (44.10).
|
|
SOLUTION. Effecting the variation, we have
|
|
f
|
|
In the second term we have used the fact that dl2 = dr2 and therefore dl d8l = dr. d&r.
|
|
Integrating this term by parts and then equating to zero the coefficient of Sr in the integrand,
|
|
we obtain the differential equation of the path:
|
|
§45
|
|
Canonical transformations
|
|
143
|
|
Expanding the derivative on the left-hand side and putting the force F = - auld gives
|
|
d2r/dl2=[F-(F.t)t]/2(E-U),
|
|
where t = dr/dl is a unit vector tangential to the path. The difference F-(F. t)t is the com-
|
|
ponent Fn of the force normal to the path. The derivative d2r/dl2 = dt/dl is known from
|
|
differential geometry to be n/R, where R is the radius of curvature of the path and n the unit
|
|
vector along the principal normal. Replacing E-U by 1mv2, we have (mv2/R)n = Fn, in
|
|
agreement with the familar expression for the normal acceleration in motion in a curved
|
|
path.
|
|
$45. Canonical transformations
|
|
The choice of the generalised co-ordinates q is subject to no restriction;
|
|
they may be any S quantities which uniquely define the position of the system
|
|
in space. The formal appearance of Lagrange's equations (2.6) does not
|
|
depend on this choice, and in that sense the equations may be said to be
|
|
invariant with respect to a transformation from the co-ordinates q1, q2,
|
|
to any other independent quantities Q1, Q2, The new co-ordinates Q are
|
|
functions of q, and we shall assume that they may explicitly depend on the
|
|
time, i.e. that the transformation is of the form
|
|
Qi=Qi(q,t)
|
|
(45.1)
|
|
(sometimes called a point transformation).
|
|
Since Lagrange's equations are unchanged by the transformation (45.1),
|
|
Hamilton's equations (40.4) are also unchanged. The latter equations, how-
|
|
ever, in fact allow a much wider range of transformations. This is, of course,
|
|
because in the Hamiltonian treatment the momenta P are variables inde-
|
|
pendent of and on an equal footing with the co-ordinates q. Hence the trans-
|
|
formation may be extended to include all the 2s independent variables P
|
|
and q:
|
|
Qt=Qi(p,q,t),
|
|
Pi = Pi(p, q,t).
|
|
(45.2)
|
|
This enlargement of the class of possible transformations is one of the im-
|
|
portant advantages of the Hamiltonian treatment.
|
|
The equations of motion do not, however, retain their canonical form
|
|
under all transformations of the form (45.2). Let us derive the conditions
|
|
which must be satisfied if the equations of motion in the new variables P, Q
|
|
are to be of the form
|
|
(45.3)
|
|
with some Hamiltonian H'(P,Q). When this happens the transformation is
|
|
said to be canonical.
|
|
The formulae for canonical transformations can be obtained as follows. It
|
|
has been shown at the end of §43 that Hamilton's equations can be derived
|
|
from the principle of least action in the form
|
|
(45.4)
|
|
144
|
|
The Canonical Equations
|
|
§45
|
|
in which the variation is applied to all the co-ordinates and momenta inde-
|
|
pendently. If the new variables P and Q also satisfy Hamilton's equations,
|
|
the principle of least action
|
|
0
|
|
(45.5)
|
|
must hold. The two forms (45.4) and (45.5) are equivalent only if their inte-
|
|
grands are the same apart from the total differential of some function F of
|
|
co-ordinates, momenta and time.t The difference between the two integrals
|
|
is then a constant, namely the difference of the values of F at the limits of
|
|
integration, which does not affect the variation. Thus we must have
|
|
=
|
|
Each canonical transformation is characterised by a particular function F,
|
|
called the generating function of the transformation.
|
|
Writing this relation as
|
|
(45.6)
|
|
we see that
|
|
Pi = 0F/dqi, =-0F/JQi H' = H+0F/dt;
|
|
(45.7)
|
|
here it is assumed that the generating function is given as a function of the
|
|
old and new co-ordinates and the time: F = F(q, Q, t). When F is known,
|
|
formulae (45.7) give the relation between p, q and P, Q as well as the new
|
|
Hamiltonian.
|
|
It may be convenient to express the generating function not in terms of the
|
|
variables q and Q but in terms of the old co-ordinates q and the new momenta
|
|
P. To derive the formulae for canonical transformations in this case, we must
|
|
effect the appropriate Legendre's transformation in (45.6), rewriting it as
|
|
=
|
|
The argument of the differential on the left-hand side, expressed in terms of
|
|
the variables q and P, is a new generating function (q, P, t), say. Thent
|
|
= Qi = ID/OPi, H' = H+d
|
|
(45.8)
|
|
We can similarly obtain the formulae for canonical transformations in-
|
|
volving generating functions which depend on the variables P and Q, or
|
|
p and P.
|
|
t We do not consider such trivial transformations as Pi = api, Qi = qt,H' = aH, with a an
|
|
arbitrary constant, whereby the integrands in (45.4) and (45.5) differ only by a constant
|
|
factor.
|
|
+ If the generating function is = fi(q, t)Pi, where the ft are arbitrary functions, we
|
|
obtain a transformation in which the new co-ordinates are Q = fi(q, t), i.e. are expressed
|
|
in terms of the old co-ordinates only (and not the momenta). This is a point transformation,
|
|
and is of course a particular canonical transformation.
|
|
§45
|
|
Canonical transformations
|
|
145
|
|
The relation between the two Hamiltonians is always of the same form:
|
|
the difference H' - H is the partial derivative of the generating function with
|
|
respect to time. In particular, if the generating function is independent of
|
|
time, then H' = H, i.e. the new Hamiltonian is obtained by simply substitut-
|
|
ing for P, q in H their values in terms of the new variables P, Q.
|
|
The wide range of the canonical transformations in the Hamiltonian treat-
|
|
ment deprives the generalised co-ordinates and momenta of a considerable
|
|
part of their original meaning. Since the transformations (45.2) relate each
|
|
of the quantities P, Q to both the co-ordinates q and the momenta P, the
|
|
variables Q are no longer purely spatial co-ordinates, and the distinction
|
|
between Q and P becomes essentially one of nomenclature. This is very
|
|
clearly seen, for example, from the transformation Q = Pi, Pi = -qi,
|
|
which obviously does not affect the canonical form of the equations and
|
|
amounts simply to calling the co-ordinates momenta and vice versa.
|
|
On account of this arbitrariness of nomenclature, the variables P and q in
|
|
the Hamiltonian treatment are often called simply canonically conjugate
|
|
quantities. The conditions relating such quantities can be expressed in terms
|
|
of Poisson brackets. To do this, we shall first prove a general theorem on the
|
|
invariance of Poisson brackets with respect to canonical transformations.
|
|
Let [f,g]p,a be the Poisson bracket, for two quantities f and g, in which
|
|
the differentiation is with respect to the variables P and q, and [f,g]p,Q that
|
|
in which the differentiation is with respect to P and Q. Then
|
|
(45.9)
|
|
The truth of this statement can be seen by direct calculation, using the for-
|
|
mulae of the canonical transformation. It can also be demonstrated by the
|
|
following argument.
|
|
First of all, it may be noticed that the time appears as a parameter in the
|
|
canonical transformations (45.7) and (45.8). It is therefore sufficient to prove
|
|
(45.9) for quantities which do not depend explicitly on time. Let us now
|
|
formally regard g as the Hamiltonian of some fictitious system. Then, by
|
|
formula (42.1), [f,g]p,a = df/dt. The derivative df/dt can depend only on
|
|
the properties of the motion of the fictitious system, and not on the particular
|
|
choice of variables. Hence the Poisson bracket [f,g] is unaltered by the
|
|
passage from one set of canonical variables to another.
|
|
Formulae (42.13) and (45.9) give
|
|
[Qi, Qk]p,a = 0, [Pi,Pk]p,a = 0,
|
|
(45.10)
|
|
These are the conditions, written in terms of Poisson brackets, which must
|
|
be satisfied by the new variables if the transformation P, q P, Q is canonical.
|
|
It is of interest to observe that the change in the quantities P, q during the
|
|
motion may itself be regarded as a series of canonical transformations. The
|
|
meaning of this statement is as follows. Let qt, Pt be the values of the canonical
|
|
t Whose generating function is
|
|
6*
|
|
146
|
|
The Canonical Equations
|
|
§46
|
|
variables at time t, and qt+r, Pt+r their values at another time t +T. The latter
|
|
are some functions of the former (and involve T as a parameter):
|
|
If these formulae are regarded as a transformation from the variables Qt, Pt
|
|
to qt+r, Pttr, then this transformation is canonical. This is evident from the
|
|
expression ds = for the differential of the action S(qt++,
|
|
qt) taken along the true path, passing through the points qt and qt++ at given
|
|
times t and t + T (cf. (43.7)). A comparison of this formula with (45.6) shows
|
|
that - S is the generating function of the transformation.
|
|
46. Liouville's theorem
|
|
For the geometrical interpretation of mechanical phenomena, use is often
|
|
made of phase space. This is a space of 2s dimensions, whose co-ordinate axes
|
|
correspond to the S generalised co-ordinates and S momenta of the system
|
|
concerned. Each point in phase space corresponds to a definite state of the
|
|
system. When the system moves, the point representing it describes a curve
|
|
called the phase path.
|
|
The product of differentials dT = dq1 ... dqsdp1 dps may be regarded
|
|
as an element of volume in phase space. Let us now consider the integral
|
|
I dT taken over some region of phase space, and representing the volume of
|
|
that region. We shall show that this integral is invariant with respect to
|
|
canonical transformations; that is, if the variables P, q are replaced by
|
|
P, Q by a canonical transformation, then the volumes of the corresponding
|
|
regions of the spaces of P, and P, Q are equal:
|
|
...dqsdp1...dps =
|
|
(46.1)
|
|
The transformation of variables in a multiple integral is effected by the
|
|
formula I .jdQ1...dQsdP1...dPz = S... I Ddq1 dp1...dps,
|
|
where
|
|
(46.2)
|
|
is the Jacobian of the transformation. The proof of (46.1) therefore amounts
|
|
to proving that the Jacobian of every canonical transformation is unity:
|
|
D=1.
|
|
(46.3)
|
|
We shall use a well-known property of Jacobians whereby they can be
|
|
treated somewhat like fractions. "Dividing numerator and denominator" by
|
|
0(91, ..., qs, P1, Ps), we obtain
|
|
Another property of Jacobians is that, when the same quantities appear in
|
|
both the partial differentials, the Jacobian reduces to one in fewer variables,
|
|
§47
|
|
The Hamilton-Jacobi equation
|
|
147
|
|
in which these repeated quantities are regarded as constant in carrying out
|
|
the differentiations. Hence
|
|
(46.4)
|
|
P=constant
|
|
q=constant
|
|
The Jacobian in the numerator is, by definition, a determinant of order s
|
|
whose element in the ith row and kth column is Representing the
|
|
canonical transformation in terms of the generating function (q, P) as in
|
|
(45.8), we have = In the same way we find that the
|
|
ik-element of the determinant in the denominator of (46.4) is
|
|
This means that the two determinants differ only by the interchange of rows
|
|
and columns; they are therefore equal, so that the ratio (46.4) is equal to
|
|
unity. This completes the proof.
|
|
Let us now suppose that each point in the region of phase space considered
|
|
moves in the course of time in accordance with the equations of motion of the
|
|
mechanical system. The region as a whole therefore moves also, but its volume
|
|
remains unchanged:
|
|
f dr = constant.
|
|
(46.5)
|
|
This result, known as Liouville's theorem, follows at once from the invariance
|
|
of the volume in phase space under canonical transformations and from the
|
|
fact that the change in p and q during the motion may, as we showed at the end
|
|
of §45, be regarded as a canonical transformation.
|
|
In an entirely similar manner the integrals
|
|
11 2 dae dph
|
|
,
|
|
in which the integration is over manifolds of two, four, etc. dimensions in
|
|
phase space, may be shown to be invariant.
|
|
47. The Hamilton-Jacobi equation
|
|
In §43 the action has been considered as a function of co-ordinates and
|
|
time, and it has been shown that the partial derivative with respect to time
|
|
of this function S(q, t) is related to the Hamiltonian by
|
|
and its partial derivatives with respect to the co-ordinates are the momenta.
|
|
Accordingly replacing the momenta P in the Hamiltonian by the derivatives
|
|
as/aq, we have the equation
|
|
(47.1)
|
|
which must be satisfied by the function S(q, t). This first-order partial
|
|
differential equation is called the Hamilton-Jacobi equation.
|
|
148
|
|
The Canonical Equations
|
|
§47
|
|
Like Lagrange's equations and the canonical equations, the Hamilton-
|
|
Jacobi equation is the basis of a general method of integrating the equations
|
|
of motion.
|
|
Before describing this method, we should recall the fact that every first-
|
|
order partial differential equation has a solution depending on an arbitrary
|
|
function; such a solution is called the general integral of the equation. In
|
|
mechanical applications, the general integral of the Hamilton-Jacobi equation
|
|
is less important than a complete integral, which contains as many independent
|
|
arbitrary constants as there are independent variables.
|
|
The independent variables in the Hamilton-Jacobi equation are the time
|
|
and the co-ordinates. For a system with s degrees of freedom, therefore, a
|
|
complete integral of this equation must contain s+1 arbitrary constants.
|
|
Since the function S enters the equation only through its derivatives, one
|
|
of these constants is additive, so that a complete integral of the Hamilton-
|
|
Jacobi equation is
|
|
Sft,q,saas)+
|
|
(47.2)
|
|
where X1, ..., as and A are arbitrary constants.
|
|
Let us now ascertain the relation between a complete integral of the
|
|
Hamilton-Jacobi equation and the solution of the equations of motion which
|
|
is of interest. To do this, we effect a canonical transformation from the
|
|
variables q, P to new variables, taking the function f (t, q; a) as the
|
|
generating function, and the quantities a1, A2, ..., as as the new momenta.
|
|
Let the new co-ordinates be B1, B2, ..., Bs. Since the generating function
|
|
depends on the old co-ordinates and the new momenta, we use formulae
|
|
(45.8): Pi = af/dqi, Bi = af/dar, H' = H+dfdd. But since the function f
|
|
satisfies the Hamilton-Jacobi equation, we see that the new Hamiltonian is
|
|
zero: H' = H+af/dt = H+as/t = 0. Hence the canonical equations in
|
|
the new variables are di = 0, Bi = 0, whence
|
|
ay=constant,
|
|
Bi = constant.
|
|
(47.3)
|
|
By means of the S equations af/dai = Bi, the S co-ordinates q can be expressed
|
|
in terms of the time and the 2s constants a and B. This gives the general
|
|
integral of the equations of motion.
|
|
t Although the general integral of the Hamilton-Jacobi equation is not needed here, we
|
|
may show how it can be found from a complete integral. To do this, we regard A as an arbi-
|
|
trary function of the remaining constants: S = f(t, q1, ..., q8; a1, as) +A(a1, as). Re-
|
|
placing the Ai by functions of co-ordinates and time given by the S conditions asidar = 0,
|
|
we obtain the general integral in terms of the arbitrary function A(a1,..., as). For, when the
|
|
function S is obtained in this manner, we have
|
|
as
|
|
The quantities (as/dqs)a satisfy the Hamilton-Jacobi equation, since the function S(t, q; a)
|
|
is assumed to be a complete integral of that equation. The quantities asida therefore satisfy
|
|
the same equation.
|
|
§48
|
|
Separation of the variables
|
|
149
|
|
Thus the solution of the problem of the motion of a mechanical system by
|
|
the Hamilton-Jacobi method proceeds as follows. From the Hamiltonian,
|
|
we form the Hamilton-Jacobi equation, and find its complete integral (47.2).
|
|
Differentiating this with respect to the arbitrary constants a and equating
|
|
the derivatives to new constants B, we obtain S algebraic equations
|
|
asidar=Bt,
|
|
(47.4)
|
|
whose solution gives the co-ordinates q as functions of time and of the 2s
|
|
arbitrary constants. The momenta as functions of time may then be found
|
|
from the equations Pi = aslaqi.
|
|
If we have an incomplete integral of the Hamilton-Jacobi equation, depend-
|
|
ing on fewer than S arbitrary constants, it cannot give the general integral
|
|
of the equations of motion, but it can be used to simplify the finding of the
|
|
general integral. For example, if a function S involving one arbitrary con-
|
|
stant a is known, the relation asida = constant gives one equation between
|
|
q1, ..., qs and t.
|
|
The Hamilton-Jacobi equation takes a somewhat simpler form if the func-
|
|
tion H does not involve the time explicitly, i.e. if the system is conservative.
|
|
The time-dependence of the action is given by a term -Et:
|
|
S = So(g)-Et
|
|
(47.5)
|
|
(see 44), and substitution in (47.1) gives for the abbreviated action So(q)
|
|
the Hamilton-Jacobi equation in the form
|
|
(47.6)
|
|
$48. Separation of the variables
|
|
In a number of important cases, a complete integral of the Hamilton-
|
|
Jacobi equation can be found by "separating the variables", a name given to
|
|
the following method.
|
|
Let us assume that some co-ordinate, q1 say, and the corresponding
|
|
derivative asia appear in the Hamilton-Jacobi equation only in some
|
|
combination (q1, which does not involve the other co-ordinates, time,
|
|
or derivatives, i.e. the equation is of the form
|
|
(48.1)
|
|
where qi denotes all the co-ordinates except q1.
|
|
We seek a solution in the form of a sum:
|
|
(48.2)
|
|
150
|
|
The Canonical Equations
|
|
§48
|
|
substituting this in equation (48.1), we obtain
|
|
(48.3)
|
|
Let us suppose that the solution (48.2) has been found. Then, when it is
|
|
substituted in equation (48.3), the latter must become an identity, valid (in
|
|
particular) for any value of the co-ordinate q1. When q1 changes, only the
|
|
function is affected, and so, if equation (48.3) is an identity, must be a
|
|
constant. Thus equation (48.3) gives the two equations
|
|
(48.4)
|
|
= 0,
|
|
(48.5)
|
|
where a1 is an arbitrary constant. The first of these is an ordinary differential
|
|
equation, and the function S1(q1) is obtained from it by simple integration.
|
|
The remaining partial differential equation (48.5) involves fewer independent
|
|
variables.
|
|
If we can successively separate in this way all the S co-ordinates and the
|
|
time, the finding of a complete integral of the Hamilton-Jacobi equation is
|
|
reduced to quadratures. For a conservative system we have in practice to
|
|
separate only S variables (the co-ordinates) in equation (47.6), and when this
|
|
separation is complete the required integral is
|
|
(48.6)
|
|
where each of the functions Sk depends on only one co-ordinate; the energy
|
|
E, as a function of the arbitrary constants A1, As, is obtained by substituting
|
|
So = in equation (47.6).
|
|
A particular case is the separation of a cyclic variable. A cyclic co-ordinate
|
|
q1 does not appear explicitly in the Hamiltonian, nor therefore in the Hamilton-
|
|
Jacobi equation. The function (91, reduces to as/da simply, and
|
|
from equation (48.4) we have simply S1 = x1q1, so that
|
|
(48.7)
|
|
The constant a1 is just the constant value of the momentum P1 = asida
|
|
corresponding to the cyclic co-ordinate.
|
|
The appearance of the time in the term - Et for a conservative system
|
|
corresponds to the separation of the "cyclic variable" t.
|
|
Thus all the cases previously considered of the simplification of the integra-
|
|
tion of the equations of motion by the use of cyclic variables are embraced
|
|
by the method of separating the variables in the Hamilton-Jacobi equation.
|
|
To those cases are added others in which the variables can be separated even
|
|
though they are not cyclic. The Hamilton-Jacobi treatment is consequently
|
|
the most powerful method of finding the general integral of the equations of
|
|
motion. |