You are on page 1of 8

Physics 342 Lecture 35

Fine Structure
Lecture 35

Physics 342
Quantum Mechanics I

Monday, April 26th, 2010

There are relativistic and electromagnetic effects we have missed in our treat-
ment of the pure Coulombic, classical approach. These are relatively easy to
put back in perturbatively. Fine structure consists of two separate physical
effects: one relativistic correction, one associated with spin-orbit coupling.
Here we will focus on the relativistic corrections to Hydrogen, in preparation
for a fully relativistic treatment. The perturbative calculations are relevant,
but not necessary to understanding the issue (relativistic modification).

35.1 Relativistic Hamiltonian

In classical mechanics, a free particle is described by the action:


Z
1
m x 2 + y 2 + z 2 dt

Sc = (35.1)
2

leading to the usual Lagrangian L = 21 m v 2 . But this free particle action


can be written suggestively by noting that the distance travelled along the
dynamical trajectory (a curve x(t) parametrized by t) is

d`2 = dx2 + dy 2 + dz 2 = x 2 + y 2 + z 2 dt2



(35.2)

so that the classical action is basically the length (squared) along the curve.
When we extremize Sc in the force free case, then, we expect to get length-
exteremized trajectories, or straight lines in this setting.
The same is true for relativistic mechanics we start with a length given
by the relativistic line element:

ds2 = c2 dt2 + dx2 + dy 2 + dz 2 , (35.3)

1 of 8
35.1. RELATIVISTIC HAMILTONIAN Lecture 35

and then the length along a curve parametrized by (some parameter, not
necessarily time) is
q
ds = c2 t2 + x 2 + y 2 + z 2 d. (35.4)
If we want the same basic free-particle action, one proportional to length-
along-the-curve, then we would naturally take:
Z q
Sr = c2 t2 + x 2 + y 2 + z 2 d. (35.5)

This action is manifestly reparametrization invariant meaning that we


can change without changing the fundamental interpretation of extremal
length for the solutions to the equations of motion. To see this, note that if
we had another parameter (), then the change-of-variables in the action
would be governed by:
dx dx d
x = = , (35.6)
d d d
for example, so that
s  2  2  2  2
Z
2
dt dx dy dz d d
Sr = c + + + d
d d d d d d (35.7)
Z q
= c2 t2 + x 2 + y 2 + z 2 d,

where now dots refer to derivatives w.r.t. .


We can, in particular, and in order to compare with classical mechanics, take
= t, the coordinate time. This means that we will be using the relativistic
action, but in the context of the laboratory frame (with the clock on its
wall as our parameter for motion). With this parametrization, the action
becomes: Z r
v2
Z p
Sr = 2 2
c + v dt = i c 1 2 dt. (35.8)
c
For this action, our relativistic Lagrangian becomes (just the integrand of
the above) r
v2
Lr = i c 1 2 . (35.9)
c
We can set the overall constant by taking the slow-motion limit and de-
manding that the relativistic Lagrangian reduce to the classical one:
1 v2 v2
 
1
Lr i c 1 = i c i . (35.10)
2 c2 2 c

2 of 8
35.1. RELATIVISTIC HAMILTONIAN Lecture 35

Additive constants dont change the Euler-Lagrange equations, so the con-


stant factor i c is irrelevant to the predictions of this low-velocity limit.
From the above, we see that we must have
i
= m = i m c. (35.11)
c
Our final form for the relativistic Lagrangian is
r
2 v2
Lr = m c 1 . (35.12)
c2

The point of all of this is to find out what the free-particle Hamiltonian is,
that way we would know how to correct the Schrodinger equation. From
the above Lagrangian, we find the Hamiltonian in the usual way, first by
identifying the relativistic momenta:

L mv
p= =q , (35.13)
v 2
1 vc2

and then, forming the Hamiltonian via Legendre transform:


 
2 m c2 1 v2
mv c2
H =vpL= q + q
2 2
1 vc2 1 vc2
. (35.14)
m c2
=q
2
1 vc2

If we finish the job and write the Hamiltonian entirely in terms of the
(relativistic) momentum by inverting (35.13)
r
p2 c2 p2
v2 = 2 H = m c2 +1 . (35.15)
p + m2 c2 m2 c2

We see that this total energy is made up of a contribution from the motion
of the particles and the rest energy of the particles (note that H = m c2 at
p = 0). To make the kinetic energy portion by itself we subtract off the rest
energy: r
2 2 p2
T = H mc = mc + 1 m c2 . (35.16)
m2 c2

3 of 8
35.2. HYDROGEN CORRECTION Lecture 35

What we have been doing so far, with the Schrodinger equation, is taking
p2
2 m as the kinetic energy (with the classical p = m v), and using the replace-
ment: p ~i to generate the quantum system. Now we see that there
are . . . relativistic difficulties. Our first move is to replace (in our minds)
the classical p with the relativistic form. That doesnt change anything in
theory. The more important shift is to expand this relativistic kinetic en-
ergy (it is difficult to modify it directly with the square root in place) and
generate corrective terms based on the low-p expansion:
 2 2 !
2 1 p2 1 p
T mc 1 + + . . . m c2
2 m2 c2 8 m2 c2
(35.17)
p2 p4
= .
2 m 8 m3 c2

35.2 Hydrogen Correction

We see that our perturbation is, effectively,  p4 with  = 8 m13 c4 . If we


want to calculate the perturbed energies of the Hydrogen atom, then, we
must be able to evaluate:
E E10 E1 =  hn | p4 |n i . (35.18)
For most states of Hydrogen, p4 is a Hermitian operator, and we can factor
the operator into a p2 portion acting on |n i and another acting on hn |:
E =  p2 n p2 |n i .


(35.19)

Now in general, finding the expectation value of p4 would require taking a


bunch of derivatives and integrating. We are going to try to avoid that, by
noting that the unperturbed Hamiltonian for Hydrogen is
p2
H= + V (r) (35.20)
2m
so that the operator p2 acts according to
p2 |n i = 2 m (En V (r)) |n i , (35.21)
and we can write
D E 1
E =  4 m2 (En V (r))2 hEn i2 2 hEn V (r) i + hV (r)2 i ,

= 4
2mc
(35.22)

4 of 8
35.2. HYDROGEN CORRECTION Lecture 35

where all the expectation values are w.r.t. the state |n i i.e. for the above,
hEn i = hn | En |n i. The energy En is just a number, so comes out of all
expectation values. For Hydrogen, the potential is

e2
V (r) = (35.23)
4 0 r
and we see that in order to evaluate the expectation values hV (r)i and its
square, we will need to know the expectation values: h 1r i and h r12 i.
The full target expression is
   2  2  !
1 e2 1 e 1
E = En2 + 2 En + . (35.24)
2 m c4 4 0 r 4 0 r2

35.2.1 Feynman-Hellmann Formula

The Feynman-Hellmann theorem concerns the change in energy of a


quantum mechanical system given a change in some parameter in the
Hamiltonian. For a Hamiltonian dependent on a parameter : H(),
we have
H() |n i = En |n i , (35.25)
and both the energy and potentially the state |n i inherit dependence
on through the eigenvalue equation.
Suppose we perturb a bit: + d , how does the energy of a
particular eigenstate respond? This is basically the same question weve
been asking all along in this perturbation section, so we know the answer
already:
H
H( + d) = H() + d + O(d 2 ), (35.26)

n d , |n i |n i + d n , we

and assuming that En En + E
have
n d = En + hn | H |n i d.
En + E (35.27)

This tells us that
En = hn | H |n i (35.28)

5 of 8
35.2. HYDROGEN CORRECTION Lecture 35

to first order in d but viewing En () as itself a function of the


parameter , we can also Taylor expand: En (+d ) = En ()+ E
d +
n

. . ., leading to the final identification:


En H
= hn | |n i , (35.29)

which is the Feynman-Hellmann formula. Note that the other terms
in the expansion would be associated naturally with higher derivatives
of En , evaluated about the point .

The utility of (35.29) should be clear think of the radial Hamiltonian for
Hydrogen:
~2 d2 ~2 ` (` + 1) e2
Hr = + , (35.30)
2 m dr2 2 m r2 4 0 r
the parameter e leads to
Hr e 1
= , (35.31)
e 2 0 r
while the derivative w.r.t. the parameter ` gives

Hr ~2 1 + 2 `
= , (35.32)
` 2 m r2
and we need expectation values w.r.t. both of these r-dependencies. We also
happen to know the associated derivatives w.r.t. energy, since En is just

m e4
En = (35.33)
32 2 20 ~2 (jm + ` + 1)2

(where jm = n ` 1).
From (35.31), we have

4 m e3
 
e 1
2 2 = 2  , (35.34)
2 2
32 0 ~ (jm + ` + 1) 0 r

or, reverting to n notation,

4 m e2
 
1
= , (35.35)
r 16 0 ~2 n2

6 of 8
35.2. HYDROGEN CORRECTION Lecture 35

4  0 ~2
and finally, in terms of the Bohr radius, a = m e2
,
 
1 1
= . (35.36)
r a n2

For the r2 expectation value, we will use (35.32)

2 m e4 ~2 (1 + 2 `)
 
1
3 = , (35.37)
32 2 20 ~2 (jm + ` + 1) 2m r2
or
m2 e4
 
1 2
= 2 = 2 . (35.38)
r2 2 4
8 0 ~ (1 + 2 `) n 3 a (1 + 2 `) n3

Putting it all together back in (35.24), we have:


 2 2 !
1 e 2 1 e 2
E = En2 + 2 En + .
2 m c4 4 0 a n2 4 0 a2 (1 + 2 `) n3
(35.39)
We can clean this up a little,

En2
 
8n
E = 3 . (35.40)
2 m c2 1 + 2`

7 of 8
35.2. HYDROGEN CORRECTION Lecture 35

Homework

This homework is not due, but allows you to check that the (relativistic)
Lagrangian treatment of particles leads to predictions of motion (dynamics)
that are as we expect.

Problem 35.1

Here we will look at the change in motion implied by the relativistic La-
grangian.

a. For a particle of mass m that starts from rest at t = 0, moving


under the influence of gravity near the earth, the classical Lagrangian has
the form:
1
L = m v 2 m g x. (35.41)
2
Find v(t) from the Euler-Lagrange equation of motion (this is a one-
dimensional problem). Plot v(t) as a function of time there is a violation
of special relativity here, make sure this is clear on your plot.

b. The relativistic Lagrangian for a free particle of mass m is:


r
2 v2
L = m c 1 2. (35.42)
c
Introduce the potential from above to write the relativistic Lagrangian for
a particle of mass m moving under the influence of gravity near the earth.
Using the Euler-Lagrange equations of motion, again solve for v(t) with
v(0) = 0. Plot your result, the velocity should now be in accord with
special relativity, make sure this is demonstrated on your plot.

8 of 8

You might also like