Você está na página 1de 18

1.

LAGRANGIAN MECHANICS
Beauty, at least in theoretical physics, is perceived in the simplicity and
compactness of the equations that describe the phenomena we observe about us. Dirac
has emphasized this point and said “It is more important to have beauty in one’s
equations than to have them fit experiment…. It seems that if one is working from the
point of view of getting beauty in one’s equations, and if one has really a sound insight,
one is on a sure line of progress.” In this sense the beauty of classical physics lies in the
fact that it can all be derived from the postulates of relativity together with just one
hypothesis, which we call Hamilton’s principle. This includes all of classical mechanics
and all of electricity and magnetism. In fact, if we postulate other interactions, such as
the Yukawa potential, the mathematical form of these interactions is very restricted. The
flexibility in the choice of natural laws is very limited.
In the future, as so-called “grand unified theories” are developed, it is expected
that even this limited flexibility will be removed. One of the remarkable developments of
modern physics has been the growing perception that the laws of physics are inevitable.
Hawking may have gone beyond the realm of pure physics when he asked the question
“Did God have any choice?” in the way She wrote the laws of physics. However, it
seems that if the universe consists of three spatial dimensions and time, and we require
causality, then there is little choice in the laws of physics.

1.1. Hamilton’s principle and the postulates of relativity

Figure 1 Reference frames


Newton stated as his first law of motion that unless acted upon by an outside
force, a body at rest will remain at rest, and a body in motion will remain in uniform
motion. A frame of reference K in which this is true is called an inertial frame. The
postulates of special relativity state that the laws of physics observed in an inertial
reference frame K are identical to those observed in another inertial reference frame K '
that moves with respect to K as shown in Figure 1. Clearly, if the reference frame K' is
also an inertial frame, it moves relative to K with at most a constant velocity, and vice
versa. Newton also said that “…absolute, true, and mathematical time, of itself, and
from its own nature, flows equably and without relation to anything external.” We now
know that this is true only in the nonrelativistic limit, but we assume here that it is true.
Although the discussion really deserves to be fully relativistic, we restrict our attention to
the nonrelativistic case. If time is absolute, then the coordinates r and r ' and the times
t and t ' in the two inertial reference frames are related by
r ' = r − Vt , (1)
t'=t, (2)
where V is the velocity of K ' in K . These are known as a Galilean transformation.
Hamilton’s principle says that as a system moves from state a to state b , it does
so along the trajectory that makes the action integral
b
S = ∫ Ldt (3)
a

an extremum, generally a minimum, subject to the constraint that the endpoints a and b
(including both the coordinates and the times) are fixed. That is, in the notation of the
calculus of variations,
b
δ S = δ ∫ Ldt = 0 (4)
a

for variations δ r of the trajectory that vanish at the endpoints, as shown in Figure 2. The
quantity L is called the Lagrangian for the system, and its form depends on the nature of
the system under consideration. The task in classical mechanics and classical field theory
therefore consists of two parts. First we must determine the Lagrangian L for the
system, and second we must find the equations of motion that minimize the action S . As
we shall see, the form of the Lagrangian follows from the postulates of relativity. Only
the few parameters that appear in the equations must be determined from experiment.

Figure 2 Variation of a trajectory.

1.2. Lagrangian for a Free Particle


Up to this point we have not said anything about the physical system we are
trying to describe, which may consist of matter, or fields, or both. We begin with a
simple, structureless, point particle, described by the coordinates r and time t . We
hypothesize that the Lagrangian depends only on the coordinates, the time, and the
velocity, but no higher derivatives of the position, so that
L = L ( r, v,t ) (5)
where v = dr / dt is the velocity of the particle. In fact, since space and time are
homogeneous, the Lagrangian of a free particle cannot depend explicitly on the
coordinates or the time, but only on the velocity. Otherwise, the behavior of the particle
would be different at different places and different times. Thus, the Lagrangian must be
simply
L = L ( v) (6)

But the Lagrangian cannot depend on the direction of v , since space is isotropic, so it can
depend only on the magnitude v 2 = v ⋅ v and have the form
L = T ( v ⋅ v) (7)

for some function T that we must determine.


Hamilton’s principle for a free particle may now be stated in the form
b
δ S = δ ∫ T ( v 2 ) dt = 0 (8)
a

Using the methods of the calculus of variations, we compute


dδ r dT
b b b
dT
δ S = ∫ δ T ( v ⋅ v ) dt = ∫ 2 v ⋅ δ v 2
dt = 2∫ v ⋅ 2
dt = 0 (9)
a a
dv a
dt dv
and integrating once by parts, we get
b
d ⎛ dT ⎞
b
dT
δ S = 2 δ r ⋅ v 2 − 2∫ δ r ⋅ ⎜ v 2 ⎟ dt = 0 (10)
dv a a
dt ⎝ dv ⎠
The first term vanishes because δ r = 0 at the endpoints. Since the second term vanishes
for all variations δ r , the rest of the integrand must vanish identically:
d ⎛ dT ⎞
⎜v ⎟=0 (11)
dt ⎝ dv 2 ⎠
Thus, v is a constant. The trajectory of a free particle is a straight line.
The form of the function T ( v ⋅ v ) is determined by the requirements of Galilean
relativity, which state that Hamilton’s principle must be equally valid in both the
reference frames K ' and K . That is,
b b
δ S ' = δ ∫ T' ( v '⋅ v ') dt ' = 0 = δ ∫ T ⎡⎣( v '+ V ) ⋅ ( v '+ V ) ⎤⎦ dt ' (12)
a a

For this to be true, it is necessary that T ⎡⎣( v '+ V ) ⋅ ( v '+ V ) ⎤⎦ and T ( v '⋅ v ' ) differ by at
most the time derivative of a function of the coordinates and the time,
d Λ ( r ', t ')
T ⎡⎣( v '+ V ) ⋅ ( v '+ V ) ⎤⎦ - T ( v '⋅ v ' ) = (13)
dt '
for in this case
b b
d Λ ( r ', t ')
δ ∫ {T ⎡⎣( v '+ V ) ⋅ ( v '+ V ) ⎤⎦ - T ( v '⋅ v ')} dt ' = δ ∫ dt ' = δΛ ( r ', t ' ) a = 0
b
(14)
a a
dt '
since the variation of the coordinates vanishes at the endpoints. But
d Λ ( r ', t ') ∂Λ
= ∇ ' Λ ⋅ v '+ (15)
dt ' ∂t '
so
∂Λ
T ⎡⎣( v '+ V ) ⋅ ( v '+ V ) ⎤⎦ - T ( v '⋅ v ') = ∇ ' Λ ⋅ v '+ (16)
∂t '
But T is independent of the coordinates and time, so ∇ ' Λ and ∂Λ / ∂t ' , which depend
only on the coordinates and time, must be constants, and we get
T [ v '⋅ v '+ 2V ⋅ v '+ V ⋅ V ] - T ( v '⋅ v ') = K1 ⋅ v '+ K 2 (17)
It is easily shown (by expanding in a power series, for example), that this can be true only
if the Lagrangian is
1 2
L (v) = T ( v ⋅ v) = mv + K (18)
2
for some constants m and K . Since it disappears from the equations of motion when the
variation is taken, we set K = 0 . We must determine the constant m by comparison with
experiment.

1.3. Lagrangian for a Particle Interacting with a Field


To describe the interaction of a particle with a field, we postulate a Lagrangian of
the form
1 2
L= mv − U ( r, t ) . (19)
2

where the first term is just the Lagrangian of a free particle. The variation of the action is
therefore
dδ r
b b
δ S = m∫ v ⋅ dt − ∫ δ Udt , (20)
a
dt a

But δ U = ∇ ⋅ δ r , so upon integrating once by parts we get


⎛ ⎞
b
dv
δ S = mv ⋅ δ r a − ∫ ⎜ m + ∇U ⎟ ⋅ δ rdt = 0
b
(21)
a⎝ ⎠
dt
According to Hamilton’s principle the first term vanishes because δ r = 0 at the
endpoints. Then, since the integral must vanish for arbitrary variations δ r , the rest of the
integrand must vanish identically. We therefore obtain the equations of motion
dv
m = −∇U (22)
dt
Comparing this with experiment, we identify the constant m as the mass of the particle
and U as the potential energy.

1.4. Invariance and Momentum

Figure 3 Translation of a trajectory.

We return to (21) for a moment, which we may now write in the form
⎛ ⎞
b
dv
δ S = mv ⋅ δ r a − ∫ ⎜ m + ∇U ⎟ ⋅ δ rdt = 0
b
(23)
a⎝ ⎠
dt
and consider the case of a translation of the entire trajectory by the constant amount
δ r = ε = constant (24)
as illustrated in Figure 3. Provided that the potential U is invariant under the translation
δ r = ε , the Lagrangian is unchanged. Therefore, the action is unchanged by the
translation and δ S = 0 for this variation of the trajectory. But the result of the translation
remains a valid trajectory, so the integral in (23) still vanishes identically. However, the
variation δ r is no longer zero at the endpoints. Therefore, we see from the first term in
(23) that
mv a = mv ( b ) − mv ( a ) = 0
b
(25)

That is, the quantity mv is conserved along the trajectory. We call those quantities that
are conserved in a translationally invariant system the momenta, so the momentum must
be
p = mv (26)
For a system of particles that attract and repel one another through central
potentials, the Lagrangian has the form
1 1
L = ∑ mi vi2 − ∑ Uij ri − r j
i 2 2 i, j
( ) (27)

where the factor of ½ appears in the second term because we have counted the interaction
between each pair of particles twice. If the positions of all the particles are translated or
rotated together, the Lagrangian is unchanged. Therefore, the total momentum and
angular momentum of a system of particles interacting according to (27) are conserved.
We can also see this by looking at the equation of motion that we derive from (27), which
is
dp k ∂L dp k ∂U
− = + ∑ kj = 0 (28)
dt ∂rk dt j ∂ri
The factor of one half has disappeared because in the sum over i two terms survive for
k = i , these being U kj and U jk = U kj . Summing over all particles we obtain

dp k ∂U
∑k dt
+ ∑ kj = 0
kj ∂rk
(29)

But
∂U kj ∂U kj
=− (30)
∂rk ∂r j
That is, the forces of interaction on the two particles are equal and opposite. Therefore,
the sum vanishes by cancellation and the total canonical momentum of the interacting
particles is conserved:
dp k
∑k dt
=0 (31)

Note that this result depends on the fact that the interaction can be written in the form
(27), in which only the instantaneous positions appear in Uij . That is, the interaction is
felt instantaneously by both particles, so every action has an equal and opposite reaction,
as represented by (30) and stated by Newton. This is valid only in nonrelativistic theory.
In relativistic theory, the interaction propagates at the speed of light, and is not felt
instantaneously by another particle. Therefore, the total momentum of the particles is not
conserved. Instead, the total momentum of the particles and fields is conserved.
Just as translational invariance is associated with linear momentum, rotational
invariance is associated with angular momentum. For an infinitesimal rotation of the
position of a particle about the origin, the increment in the coordinates is linear in the
angle of rotation and in the coordinates of the particle. The variation of the trajectory is
therefore
δ r = δω × r (32)
where δω is the rotation vector.
Figure 4 Rotation of a trajectory.
When the trajectory is rotated as shown in Figure 4, the variation of the action is
given by (23), as before. Substituting (32) for δ r , we now get
⎛ ⎞
b
dv
δ S = mv ⋅ (δω × r ) a − ∫ ⎜ m
b
+ ∇U ⎟ ⋅ δ rdt (33)
a⎝ ⎠
dt
For a system which is rotationally invariant, the action is unaffected by this
transformation, so δ S = 0 . But the trajectory remains valid, so the integral on the right
still vanishes. Therefore, upon rearranging the triple product, we find that

m v ⋅ (δω × r ) a = 0 = δω ⋅ ( r × p ) a
b b
(34)

That is, l = r × p is a constant of the motion. When the Lagrangian is invariant under a
rotation in space, the angular momentum l is conserved. For a set of particles that
interact through central potentials, as described by (27), the Lagrangian is invariant under
a rotation of the positions of all the particles. Therefore, we can repeat the arguments
used for linear momentum to show that the total angular momentum of all the particles is
conserved.
2. HAMILTONIAN MECHANICS

2.1. Canonical Momentum


In the Lagrangian formulation of mechanics, the variables in the Lagrangian are
the coordinates r and t , and the velocity v = dr / dt ,
1 2
L= mv − U ( r, t ) = L ( r, v, t ) (35)
2
The definition of the velocity v = dr / dt is used explicitly in the variation of the
trajectory to obtain the Euler-Lagrange equations of motion. But another approach is
possible, called Hamiltonian mechanics. In the Hamiltonian formulation of mechanics
we introduce the canonical momentum P and change from the variables r , v , and t to
the variables r , P , and t , and we give the coordinates and momenta equal standing. The
space spanned by the coordinates r and the momenta P is called the phase space of a
system, and the motion of a system is described by a trajectory in phase space with t
acting as a parameter along the trajectory. For example, the nonrelativistic motion of a
one-dimensional harmonic oscillator, for which P = p , is an ellipse in the x - px phase
plane, as shown in Figure 5, with time acting as a parameter. To change variables from
( r, v,t ) to ( r, P, t ) we use what is called a Legendre transformation, and to make the new
variables independent of one another we carry out the variation of the trajectory
separately for r and P , ignoring the connection between position and momentum. The
resulting equations of motion, called the canonical equations of motion, are simpler and
more symmetric than the Lagrangian equations. However, there are now twice as many
equations of motion since each is a first-order differential equation instead of second-
order.

Figure 5 Phase space of a harmonic oscillator


To begin, we must first define the canonical momentum P . For a Lagrangian
L ( r, v,t ) the variation of the action is

⎛ ∂L ∂L ⎞
b
δ S = ∫⎜ ⋅δ r + ⋅ δ v ⎟ dt (36)
a⎝
∂r ∂v ⎠
But the variation of the velocity is
dδ r
δv = (37)
dt
Substituting this into (36) and integrating once, by parts, in the usual fashion, we find that
∂L ⎛ d ∂L ∂L ⎞
b
δS = ⋅δ r a − ∫ ⎜ − ⎟ ⋅ δ rdt = 0
b
(38)
∂v a⎝
dt ∂v ∂r ⎠
by Hamilton’s principle. But the variation δ r vanishes at the endpoints, and is arbitrary
in between. Therefore, the first term vanishes, and from the second term we obtain the
equations of motion
d ∂L ∂L
− =0 (39)
dt ∂v ∂r
These are known as the Euler-Lagrange equations, and they may be used to find the
equations of motion for any Lagrangian L ( r, v,t ) .

Returning to (38), we consider a translation δ r = ε of the trajectory. For a system


that is translationally invariant we may use the arguments given previously to show that
the quantity
∂L
P= (40)
∂v
is conserved. We call this the canonical momentum. It is a simple matter to confirm that
for a particle in a potential Φ ( r,t ) , this definition agrees with the previous definition of
the momentum of a particle. However, it is more general than that definition, which
makes Hamiltonian (and, for that matter Lagrangian) mechanics so useful. In terms of
the canonical momentum, the Euler-Lagrange equations of motion are expressed in the
suggestive form
dP ∂L
− =0 (41)
dt ∂r
When ∂L / ∂r = −∇U = F for some force F , (41) looks just like Newton’s equation of
motion.

2.2. Legendre Transformation


The change from the variables r , v and t to the variables r , P and t is
accomplished by using a Legendre transformation. Legendre transformations are
frequently used in thermodynamics to change from one set of variables to another. For
example, a liquid or gas may be characterized by its pressure p , volume V , and
temperature T . From the definition of the entropy S we see that the heat added to a
system in a reversible process is dQ = TdS and the work done by the system is
dW = pdV . According to the First Law of Thermodynamics the change of the internal
energy in a reversible process is
dU = dQ − dW = TdS − pdV (42)
This shows that the internal energy does not change in a process that occurs at constant
entropy and volume. Therefore, the internal energy U is a function of the entropy S and
volume V . For processes such as a change of state that take place at constant
temperature and pressure, however, the Gibbs function is more useful. The Gibbs
function is related to the internal energy by the Legendre transformation
G = U + pV − TS (43)
Since
dG = dU + pdV + Vdp − TdS − SdT = Vdp − SdT (44)
we see that the Gibbs function is a function of the variables p and T . For example, if
we boil water at constant pressure, the temperature remains constant as we add heat, but
the steam expands at constant pressure. In this case the internal energy of the system
increases, but the Gibbs function, which is a function of the variables p and T , remains
constant as the water changes to steam.
Returning to the equations of motion of a particle in a field, we see that to change
the variables of the system from r , v , and t to r , P , and t we use a Legendre
transformation and introduce the Hamiltonian function
H = P⋅v − L (45)
Using the definition (40) of the canonical momentum, we find that
∂L ∂L
d H = P ⋅ dv + v ⋅ dP − d L = v ⋅ dP − ⋅ dr − dt (46)
∂r ∂t
Evidently, the Hamiltonian is the desired function of r , P , and t .

2.3. Canonical Equations of Motion


To derive the canonical equations of motion we begin, as before, with Hamilton’s
principle,
b b
δ S = δ ∫ Ldt = δ ∫ ( P ⋅ v − H ) dt = 0 (47)
a a

Since the endpoints are fixed, we can take the variation inside the integral and get
dδ r
b b b
dr
δ S = ∫δ P ⋅ dt + ∫ P ⋅ dt − ∫ δ Hdt = 0 (48)
a
dt a
dt a

But the variation of the Hamiltonian is


∂H ∂H
δH = ⋅δ r + ⋅δ P (49)
∂r ∂P
If we substitute this into (48) and integrate the second term by parts, remembering that
the variation δ r vanishes at the endpoints, we find that the variation of the action is
given by two terms,
⎛ dr ∂H ⎞ ⎛ dP ∂H ⎞
b b

∫a ⎜⎝ dt − ∂P ⎟⎠ ⋅ δ Pdt − ∫a ⎜⎝ dt + ∂r ⎟⎠ ⋅ δ rdt = 0 (50)

Now, in the Hamiltonian formulation the coordinates r and P of phase space are given
equal standing, so the variations δ P and δ r are individually arbitrary. Therefore, the
quantities in parentheses must individually vanish, and we arrive at the canonical
equations of motion
dr ∂H
= (51)
dt ∂P
dP ∂H
=− (52)
dt ∂r
We also see that the total time derivative of the Hamiltonian is
d H ∂H dr ∂H dP ∂H ∂H
= ⋅ + ⋅ + = (53)
dt ∂r dt ∂P dt ∂t ∂t
Therefore, unless the Hamiltonian is explicitly time dependent, it is a constant of the
motion. Since it is conserved in a time-invariant system, the Hamiltonian must be the
total energy, or something proportional to it. For a particle in a field we compute
⎛1 ⎞ 1
H = P ⋅ v − L = mv ⋅ v − ⎜ mv 2 − U ⎟ = mv 2 + U (54)
⎝2 ⎠ 2
Note carefully, however, that when using the canonical equations of motion it is
important that the Hamiltonian be expressed in terms of the coordinates and the canonical
momenta, not in terms of the velocities or, when they are different, the ordinary
momenta. For example,
p2
H= +U (55)
2m
is the proper form of the Hamiltonian for a particle in a field.

2.4. Liouville’s theorem


Hamiltonian mechanics is the natural framework for describing the motion of
physical systems in phase space, and phase space is where we live in statistical
mechanics, chaos theory, and the physics of particle beams. Fundamental to the
discussion of systems in phase space is Liouville’s theorem.
We consider a system described by the phase-space coordinates ( Pi , qi ) . For a
single particle in one dimension, the coordinates are ( x, p x ) , and the phase space is
shown in Figure 6. Within this phase space, we consider a volume V enclosed by the
surface S . Inside this surface are a number N of systems. The surface S moves with
⎛ dq dP ⎞
the local phase velocity v phase = ⎜ i , i ⎟ . Since the trajectories of two particles cannot
⎝ dt dt ⎠
cross in phase space, no particles can cross the surface S , and the number of particles
inside remains constant.

Figure 6 Liouville’s theorem


Interestingly, and importantly, as the surface S moves around in phase space, its
volume V remains constant. To see this, we observe that the rate of change of the
volume in phase space is
dV
dt v∫S
= v phase ⋅ nˆ dS (56)

where n̂ is a unit vector normal to the surface S . But by the divergence theorem,
dV
dt v∫S
= v phase ⋅ nˆ dS = ∫ ∇ phase ⋅ v phase dV (57)
V

where the divergence of the velocity in phase space is


⎛ ∂q ∂P ⎞ ⎛ ∂ ∂H ∂ ∂H ⎞
∇ phase ⋅ v phase = ∑ ⎜ i + i ⎟ = ∑ ⎜ − ⎟=0 (58)
i ⎝ ∂qi ∂Pi ⎠ i ⎝ ∂qi ∂pi ∂pi ∂qi ⎠
when we use the canonical equations of motion. Therefore, we see that
dV
=0 (59)
dt
which is called Poincare’s theorem. But since this is true for an arbitrarily small volume,
and the number of points in the volume remains constant, we see that the density ρ of
points in phase space is a constant of the motion. That is,

=0 (60)
dt
This is called Liouville’s theorem. Note that it is derived from the canonical equations of
motion for a conservative system. It doesn’t apply when there are dissipative forces, such
as friction or drag. For example, for a harmonic oscillator with dissipation, the volume of
phase space that encloses any set of particles will eventually collapse to a point at the
origin. All the particles will eventually come to rest.

Figure 7 Particle beams in real space, viewed at the indicated sections.


We have already seen the usefulness of the phase plane for the theory of chaotic
motion. To illustrate the use of Liouville’s theorem to describe particle beams, we
consider the beam illustrated in Figure 7 and Figure 8. In Figure 7, both beams are
presumed to have uniform, round distributions in transverse position and transverse
momentum. Clearly, however, the beams are different; one is at a rough focus, while the
other is emerging from a sharp focus. The differences are much more apparent in Figure
8. In the first beam, the momentum and position in the x direction (likewise the y
direction) is uncorrelated, but in the second case the x -momentum is proportional to the
x -position and the distribution in phase space is very different. In the case of a perfectly
focused beam the area occupied by the beam vanishes.

Figure 8 Particle beams in phase space, viewed at the indicated section.

The usefulness of the phase-space picture is shown in Figure 9. As the beam passes
through the focus, the area in phase space occupied by the beam shears in the horizontal
direction as the trajectories take the particles to new values of y , keeping p y constant.
From Liouville’s theorem we know that the area of the distribution remains constant
during this evolution. At the focus, the ellipse circumscribing the particles is erect, and
the width in the y -direction is a minimum. In fact, if we pass the beam through a
focusing magnetic field, we shear the beam vertically (the momenta p y change but, at
least for a thin lens, the vertical positions y are almost unchanged), but the area remains
constant. As the beam drifts toward the focus, the distribution shears in the horizontal
direction (the positions change, but the momenta are constant). When the beam reaches a
focus, the ellipse is again erect, and the width in the y -direction is a minimum. Since the
area is a constant, we can find the width in the y -direction by knowing the width in the
p y -direction. This is established by the lens, which determines the shear in the vertical
( p y ) direction in the phase plane. We can obtain a narrower focus by extending the
distribution in the vertical direction, or we can collimate the beam (narrow the
distribution in the p y -direction), but only by spreading out the beam in the y -direction
to keep the area constant.

Figure 9 Particle-beam evolution through a focus, viewed at the indicated sections.


3. VELOCITY-DEPENDENT POTENTIALS

3.1. Lagrangian Mechanics


Thus far, we have restricted our attention to forces that depend only on the
position of the particle and not on its motion. This excludes magnetic forces on charged
particles and friction and other dissipative forces in more complex systems. We shall not
discuss friction, here, but magnetic forces are easy to include in the theory. To do this,
we postulate a Lagrangian of the form
1 2
L= mv − U ( r, v, t ) (61)
2
where U need not be an energy, just an interaction term with the units of energy. We
assume further that the potential U is linear in the velocity. If this doesn’t agree with
experiment, we can correct it later. Since the Lagrangian is a scalar, the vector velocity
can appear only in a scalar product, so the Lagrangian must have the form
1 2
L ( r , v, t ) = mv − qΦ ( r, t ) + qv ⋅ A ( r, t ) (62)
2
where q is a parameter that we will later identify with the charge on the particle, Φ is
called the electric potential, and A is called the magnetic vector potential.
For motion in the x -direction, the Euler-Lagrange equation of motion (39) is
d ∂L ∂L
− =0 (63)
dt ∂vx ∂x
which becomes
d ∂Φ ∂A
( mvx + qAx ) + q − qv ⋅ = 0 (64)
dt ∂x ∂x
But the derivatives appearing here are
dAx ∂A ∂A ∂A ∂A
= vx x + v y x + v z x + x (65)
dt ∂x ∂y ∂z ∂t

∂A ∂A ∂A ∂A
v⋅ = v x x + v y y + vz z (66)
∂x ∂x ∂x ∂x
If we substitute these into (64) and rearrange the terms, we get
dvx ⎡ ∂Φ ∂Ax ⎛ ∂A ∂A ⎞ ⎛ ∂Ax ∂Az ⎞ ⎤
m = q ⎢− − + vy ⎜ y − x ⎟ − vz ⎜ − ⎟⎥ (67)
dt ⎣ ∂x ∂t ⎝ ∂x ∂y ⎠ ⎝ ∂z ∂x ⎠ ⎦

If we define the electric and magnetic fields by


∂A
E = −∇Φ − (68)
∂t
B = ∇× A (69)

then the equation of motion becomes


dv
m = q (E + v × B) (70)
dt
which is just the Lorentz force law for a particle in an electromagnetic field. We see now
that for a force linear in the velocity, this is the only possible equation of motion.

3.2. Hamiltonian Mechanics


For a Lagrangian of the form
1 2
L ( r , v, t ) = mv − qΦ ( r, t ) + qv ⋅ A ( r, t ) (71)
2
the canonical momentum defined earlier by (40) is
∂L
P= = mv + qA (72)
∂v
The Hamiltonian defined by (45) is then
1 2
H = P⋅v − L = mv + qΦ (73)
2
which is the total energy of the particle. However, the Hamiltonian is correctly expressed
in terms of the canonical momentum, not the velocity, so we write
P − qA
v= (74)
m
and the Hamiltonian becomes

( P − qA )
2

H= + qΦ (75)
2m
The canonical equations of motion are now
dr ∂H P − qA p
= = = =v (76)
dt ∂P 2m m

∂ ( P − qA )
2
dP ∂H
=− =− − q∇Φ (77)
dt ∂r ∂r 2m
Using the vector identity
∇ ( a ⋅ b ) = a × ( ∇ × b ) + b × ( ∇ × a ) + ( a ⋅∇ ) b + ( b ⋅∇ ) a (78)
and the definitions (68) and (69) of the electric and magnetic fields, the second equation
becomes
dp
= q (E + v × B) (79)
dt
as before. Ho hum.
Hamiltonian mechanics is useful for formal developments, such as Liouville’s
theorem, but the canonical momentum itself is frequently useful for solving problems.
For example, for a plane electromagnetic wave the field is necessarily perpendicular to
the direction of propagation (to satisfy Gauss’s law), and has the form
A = A ⊥ ( x, t ) (80)
where A ⊥ is independent of the coordinates in the y and z directions. We can ignore
the potential Φ . Since the Lagrangian (or the Hamiltonian, for that matter) is invariant in
the y and z directions, the transverse components of the canonical momentum are
conserved. Thus, the first integral of the motion is immediately
P⊥ = p ⊥ + qA ⊥ = constant (81)
and the transverse momentum is
p ⊥ = −qA ⊥ + constant (82)
where the constant is the momentum of the particle before the wave arrives.

Figure 10 Vector potential in a uniform magnetic field


As another example, we consider Larmor’s theorem for the behavior of atoms in a
magnetic field. For a uniform magnetic field B , the magnetic vector potential may be
expressed in the form
1
A = B×r (83)
2
This consists of circles about the z axis, as shown in Figure 10. The Lagrangian
1
L = ∑ mvi2 − q ∑ Φ ( ri , t ) + ∑ qv ⋅ A ( ri , t ) (84)
i 2 i i

consists of a central potential about the nucleus, plus the interaction of the electrons with
the nucleus and each other, represented by Φ , and the magnetic vector potential (83).
Since all the terms in (84) are invariant under rotations about the magnetic field, the total
canonical angular momentum about the direction of the magnetic field is conserved:

∑ r × P = ∑ r × (p
i
i i
i
i i − qA i ) = constant (85)

where the constant is the angular momentum of the atom in the absence of the magnetic
field. Thus, the field causes the atom to rotate about the direction of the magnetic field.
To see this, we change to a new coordinate system rotating at the frequency Ω about an
axis parallel to the field B . The velocity of the particles in the new frame of reference is
v 'i = v i − Ω × r 'i . (86)
If we substitute this into (84) we get
N
L = ∑ ⎡⎣ 12 mv 'i2 + m ( Ω × r 'i ) ⋅ v 'i + 12 q ( B × r 'i ) ⋅ v 'i ⎤⎦ − q ∑ Φ ( ri , t ) , (87)
i =1 i

to lowest order in Ω and B . If we choose the frequency of rotation to be


qB
ΩL = − (88)
2m
the effect of the magnetic field vanishes in the rotating coordinate system. We therefore
see that the motion of charges in a magnetic field is just the motion in the absence of the
magnetic field plus a rotation of the entire system at the Larmor frequency Ω L . This is
Larmor's theorem. Note that the Larmor frequency Ω L is half the cyclotron frequency of
the particles in the same magnetic field, and the direction of the rotation vector ΩL is
opposite the direction of the magnetic field B for positive charges.

Você também pode gostar