Sie sind auf Seite 1von 12

Ph125c lecture notes, 4/5/01

Introducing the spectroscopic notation nL j , we have thus found the fine-structure energy
corrections
E 2s 1/2 
5 mc 2 4 , 
128
E 2p 1/2  
5 mc 2 4 , 
128
E 2p 3/2  
1 mc 2 4 . 
128
[Cohen-Tannoudji et al Ch. XII Fig. 2]

The hyperfine Hamiltonian

In real hydrogen the nucleus (proton) has an intrinsic spin of 1/2. The hydrogen nuclear
magnetic moment operator may be written
 
gp N
 
 I I,


where I is the usual spin-1/2 vector operator, g p 5.585 is the proton “g-factor” (having to

do with its internal quark structure) and N is the “nuclear magneton” 
N
qp
 
2m p
(where q p    
q e ). It should be noted that N is smaller than the electron’s “Bohr magneton”
B
qe
2m e 


by a factor of m e /m p 1/2000. The electron’s magnetic moment operator is 
S
ge B
S,  
with g e  2.

1
   
The hyperfine Hamiltonian is given by




   
W hf 0

q
L. I  
1 3 .n
I .n S. I
8
   
. r ,


4 mer3   r3 S 3 S I
where n is the unit vector r/ r . It can be shown that the first two terms here are of order
2000 times smaller than W SO , and the third term is likewise about 2000 times smaller than
W D (which also contains a delta-function). The first term of W hf reflects the interaction of the
nuclear magnetic moment with the magnetic field 0 /2
   " !
qL/m e r 3 created by the orbiting
electron. The second term represents the magnetic dipole-dipole interaction between the
nuclear and electronic spins. The third term, known as Fermi’s ‘contact term,’ has to do with
the internal magnetic structure of the proton.

The important thing to note about the hyperfine Hamiltonian is that it leads us to consider
coupling of angular momenta between the nuclear spin I and the total electron angular
momentum J, yielding 
 # &$ % $
F I J.
The rationale for this coupling scheme (as opposed, e.g., to coupling S and I first and then
adding L) is that the spin-orbit coupling is so much larger than the hyperfine interaction. As
a result, in the absence of an applied magnetic field, W hf leads to small additional splittings
within the nL j fine-structure levels [Cohen-Tannoudji et al, Ch. XII Figs. 3 and 4]:

' (*)
(A /2 1420405751.768 + 0.001 Hz, corresponding to the famous 21 cm line in hydrogen).

2
Zeeman effect of the 1s ground state hyperfine structure

[Cohen-Tannoudji, Diu, and Laloe Ch. XII section E]

In this lecture we’ll see our first good physical example of a situation where “good quantum
numbers go bad” as a function of some external parameter. The basic idea to keep in mind
is that we’ll have an overall Hamiltonian


H H0 W,
where H 0 and W have comparable eigenvalues, we suppose H 0 , W  
0, and is a

When
 
real-valued scalar parameter that can be varied by changing experimental conditions.
1, we can assume that the eigenstates of H are essentially those of H 0 , with W


merely contributing small energy corrections. Therefore the eigenvalues of H 0 are a good

quantum number. On then other hand when 1, we expect to be able to treat H 0 as a

perturbation on W. Then the eigenvalues of W should be a good quantum number in the
regime of large . Nice and simple!

Last time, we discussed the fine and hyperfine structure in atomic hydrogen. We started
from our basic results on the non-relativistic Coulomb Hamiltonian
p2 e2 ,
H0
2m r
whose energy spectrum
me 4
En
2 2n2
is highly degenerate, and considered how the successive perturbations

3
   
Wf  p4
8m 3e c 2
 1 1 dV r L.S
2m 2e c 2 r dr
  8m c  V r  , 2
2 2
2

.  r1  3  .n .n .  83  . 


e


q

0
W hf I S I L S I S I r ,
4 mer3 3

lifted some of the degeneracies and led us to the definition of new quantum numbers
J L S, ! "
F ! I " J.
(Note that I’ll dispense with the vector notation on these operators when there is no risk of
ambiguity.) As a result, we end up with the following modified picture of the hydrogen 1s
shell [C-T et al, Ch. XII Fig. 3]:

Let’s review what’s going on here.


At the level of H 0 and before considering spins, the 1s shell has no degeneracy since
! ! !
l 0. With s 1/2 and I 1/2 however, we should expect there to be a total of four states in
the shell, which could be taken for example to be
! !
|n 1 , l 0 , s 1/2 , I 1/2 ; m s !
1/2 , m I !
1/2 , !" !" #
|n ! 1, l ! 0, s ! 1/2 , I ! 1/2 ; m s ! " 1/2 , m
 I ! $ 1/2 # ,

|n ! 1, l ! 0, s ! 1/2 , I ! 1/2 ; m s !$ 1/2 , m I !" 1/2 # ,
! !
|n 1 , l 0 , s 1/2 , I 1/2 ; m s ! I ! !$ 1/2 , m !$ 1/2 # .
%
With this choice of basis in the orbital electron-spin nuclear-spin Hilbert space we have %
picked the eigenvalues of S z , I z as quantum numbers.
The fine structure Hamiltonian W f does not lift this degeneracy at all. Thinking about
each term individually, the mass-velocity term
W mv
p4
8m 3e c 2
!$
does not act on the spin degrees of freedom so it can only contribute an overall shift of the

4
1s shell, which evaluates (using the non-relativistic Coulomb radial function) to
W mv 1s 5 mc 2 4 .   
8
Likewise for the Darwin term

  V r ,
2
2
WD
8m 2e c 2
which evaluates to
WD   1s
1 mc 2  4
.
2
Hence we are left with the spin-orbit term
W SO  1 1 dV r L.S.
2 2 r


2m e c dr
As we discussed last time the presence of the L.S operator should generally lead us to

work in a coupled basis for the orbital and electron-spin angular momenta, the general
reasoning being that
L, W SO 0, 
S, W SO  0,
J, W SO   0,
since



1 J2 L2 S2 .
L.S  
2
Hence in order to apply degenerate perturbation theory for W SO we should switch to the
basis
 
|n 1 , j 1/2 , I 1/2 ; m j 
1/2 , m I 
1/2 ,  
|n  1 , j  1/2 , I  1/2 ; m    1/2 , m   1/2  ,
|n  1 , j  1/2 , I  1/2 ; m   1/2 , m  1/2  ,
j I

|n  1 , j  1/2 , I  1/2 ; m   1/2 , m   1/2  ,


j I

j I

(where j  1/2 is the only possible value with l  0 and s  1/2). Hence we have now
switched to using the eigenvalues of J , I as quantum numbers. Of course, since we are
talking about an s shell  
z z

W   L  S  L  S  L  S 
SO 1s x x y y z z 1s

 0
as
 
L x |l 0 L y |l 0 L z |l 0 0.    
Still, we proceed in this fashion in order to maintain consistency with the overall procedure
that must be used in subshells of higher angular momentum.
In conclusion, we find that the fine-structure term merely shifts the 1s 1/2 level as a whole
by 
W f 1s 1 mc 2 4 .
8
   

As for the hyperfine Hamiltonian

5
    



  
W hf
4
q
mer3
0
L. I 1 3 .n
r3 S I .n  
S. I
8
3 S I
. r ,     
we immediately see that the first term will be zero in an s shell for the same reason as W SO .
It turns out that the second term has zero mean value as well, due to spherical symmetry of
the s-state angular wave function. What remains then is to evaluate the contact term.
The matrix elements of the Fermi contact term have the form

n 1 , l 0 ; ms , mI |  2 0
.       !
r |n 1 , l 0 ; m s , m I " "
# $ $ &% &%
3 S I
" n 1, l 0| " 2 0
"   ! " "
r |n 1 , l 0 m s , m I | S . I |m s , m I '
( # $ $
3
A m s , m I |S.I |m s , m I , '
where the prefactor A now contains both the radial matrix element and the gyromagnetic
ratios. It evaluates to
* ,
A 4 gp me mec2 4 1 me 3 1 ,
3 mp mp ) 2
+ -
.-,
which is easily seen to be a factor m e /m p smaller than fine structure (recall that the spin
operators will cancel out the trailing 2 ). Now in the same way that we treated the spin-orbit
coupling, we see that we should go to a coupled basis for the electron-spin and
nuclear-spin angular momenta
F S I, ) +
namely
|n 1 , j 1/2 , I 1/2 ; F 1 , m F) 1, ) ) ) )+ /
|n 0 1, j 0 1/2 , I 0 1/2 ; F 0 1 , mF 0 0/ ,
|n 0 1, j 0 1/2 , I 0 1/2 ; F 0 1 , mF 01 1 / ,
0
|n 1 , j 1/2 , I 1/2 ; F 0 , m F 0 0 0 0 0/ ,
0
with F 1, 0 being the only possibilities for s 1/2 and I 1/2. Hence we are now in a 0 0
2
situation where in the 4D electron-spin nuclear-spin Hilbert space we are taking the
eigenvalues of F 2 , F z as quantum numbers. In this basis the contact term perturbation is
diagonal, since
AS.I A F2 S2 I2 .
2
0 1 1
Looking at this expression, we also note that S.I depends only on F and not m F , 3 /
0
AS.I |n 1 , j 1/2 , I 1/2 ; F, m F 0 0 /
0 1
A F 2 S 2 I 2 |n 1 , j 1/2 , I 1/2 ; F, m
2
1 F 0 0 0 /
0 4
A 2 FF 1
2
3
4
5 6 71
3 |n 1 , j 1/2 , I 1/2 ; F, m ,
4 F 1 0 0 0 /
so we should expect the hyperfine coupling to split the 1s 1/2 shell into an F 1 triplet (shifted 0
3 / 06 4
by W hf A 2 /4) and an F 0 singlet (shifted by W hf 0 3A 2 /4). Have another look at 3 /0 1 4
the figure at the beginning of this lecture, as well as the corresponding diagram for the n 2 0
shell [C-T et al, Ch. XII Fig. 4]:

6
The important point is that W f makes j a good quantum number for lifting the degeneracies
associated with l and s, but then W hf ultimately makes F the good quantum number when
we consider nuclear spin I as well.

Believe it or not, this is now the starting point for our discussion of the Zeeman effect and of
the general picture
H H0 W.

Specifically, we will consider H 0 to be the hydrogen Hamiltonian with fine and hyperfine
couplings, for which n, j, F are the good quantum numbers (recall that the 2s 1/2 2p 1/2 
splitting is strictly due to the Lamb shift, which does not come from any Hamiltonian we
have considered!). Now we are going to add a further perturbation associated with an
externally-applied magnetic field.


The overall form of the Zeeman Hamiltonian should be pretty familiar by now,
Wz  B0    l  s  I .
[Before moving on let us note that in principle there ought to be an additional term,
quadratic in B 0 , in the Zeeman Hamiltonian – see Cohen-Tannoudji et al p. 1233 for more

details.] Here B 0 is the applied magnetic field (created by some sort of laboratory magnet),


and the ’s are magnetic-moment operators. We already know that

 
qg p
 
qg e
s S, I I,
2m e 2m p
 
where g e 2 and g p 5.585. Just as with any current ‘loop,’ there is also a magnetic
moment associated with the orbiting electron charge:
 
q
L. 

l 2m e
(Comparison of this expression with the ones for s,I may help motivate the definition of g
factors). If we simply take the z coordinate axis to coincide with the orientation of the applied

magnetic field B 0 , we may simplify
Wz 0 Lz
   
2S z NIz   
 B0,

7
 
 
where B 0 B 0 z and thus
q
B , 0
2m e 0
N
q
g B ,  
2m p p 0
are the Larmor frequencies. Let’s begin by considering the weak-field case where B 0 is
assumed to be small.
   
Note that N 0 since m p m e . In order to simplify drastically the following

 
discussion, we will henceforth neglect the N I z term. In the 1s shell, we can likewise drop the
0 L z term since l 0 for these states. Hence we are left considering the effect of the

 
perturbation
WZ 0 2S z

    

on the set of states

 0
,
|n 1 , j 1/2 , I 1/2 ; F 1 , m F 1,
   
 1
,
|n 1, j 1/2 , I 1/2 ; F 1 , mF
   
 0
.
|n 1, j 1/2 , I 1/2 ; F 1 , mF
|n 
0 , mF 1, j  1/2 , I  1/2 ; F 
As only the last (F 

0) state differs in energy from the rest, we should begin by trying to
diagonalize S z in the F 1 manifold.

Using our favorite method (e.g. a table of Clebsch-Gordon coefficients) we can write the
 


F 1, m F basis states in terms of the eigenstates of S z , I z for s 1/2 and I 1/2 :
|F 1 , m F 
1 1 1 , 

 
2 s 2 I
|F 1 , m F 0  1 1 1  1 1 , 
2 2 2 2 I

  
2 s I s

|F 1 , m F 
1 1 1 .  

2 s 2 I

 1
  |F  1 , m  1
,
It is then easy to verify that
S z |F  1 , mF F


1  1  
2
S |F  1 , m  0 
z
1
F
1 1
2 2 2 2 2 2
 |F  0 , m  0
,
2 s I s I


|F  1 , m   1
.
F

    
2
S |F 1 , m
z 1 F F

Within the F  1 manifold then, we see that


2

1 0 0
Sz  0 0 0
2
0 0 1

and hence is already diagonal.
In the F 0 manifold,

8
S z |F 0 , mF 0
 Sz 1
 1   1   1   1
2 2 2 2
2
 
s I s I

1  1   1   1   1
2 2 2 2 2 2 2
 s

I s I

|F 1 , m F 0 .
2
Hence for the purposes of first-order perturbation theory we have
S 
z 1s 1/2 ,F 0 0,

 1   A 2  
0 ,
and the overall effect of the Zeeman Hamiltonian in the 1s shell is:
|F 1 , m F
4
|F 1 , m F 0  A ,
2
4
|F 1 , m F  1  A  2 
0 ,
4
|F 0 , m F 0  3A .
2
4
In the ‘linear’ regime where first-order perturbation theory is valid, we thus have an
energy-level diagram that looks like [C-T et al, Ch. XII Fig. 5]:

This lifting of the F 1 degeneracy is known as ‘Zeeman splitting.’

For very strong applied magnetic fields it is possible to make the Zeeman term much larger
than the hyperfine term in the overall Hamiltonian. In such a scenario we should treat W hf as
a perturbation on the (non-relativistic Coulomb  fine-structure  Zeeman) Hamiltonian!
Still limiting our attention to the 1s shell, let’s go back to the fine-structure level, including
nuclear spin in our description but before considering the hyperfine coupling of J and I into
F. Our fine-structure basis states are thus

9
 1/2 , m  1/2  ,
 1/2  ,
|n 1, l 0, j 1/2 , I 1/2 ; m j
 1/2 , m
I

 1/2  ,
|n 1, l 0, j 1/2 , I 1/2 ; m j
 1/2 , m
I

 1/2  .
|n 1, l 0, j 1/2 , I 1/2 ; m j
 1/2 , m
I

|n 1 , l 0 , j 1/2 , I 1/2 ; m j I

The effect of the Zeeman term


WZ 0 2S z


is very simple to treat exactly, since with l 0 the simultaneous eigenstates of J 2 , J z are the
same as those of S 2 , S z (since J L S S). Hence the energies of the basis states
(which start out being degenerate since we have not yet applied hyperfine couplings) shift

    

according to
|n 1 , l 0 , j 1/2 , I 1/2 ; m j 1/2 , m I
|n 1 , l 0 , j 1/2 , I 1/2 ; m  1/2 , m  1/2
  
 ,
1/2 0,

|n 1 , l 0 , j 1/2 , I 1/2 ; m  1/2 , m  1/2


  
 ,
j I 0

|n 1 , l 0 , j 1/2 , I 1/2 ; m  1/2 , m  1/2


  
 .
j

j
I

I
0

two 2-fold degenerate subspaces (corresponding to m  1/2), we should again start by


Next we must add the hyperfine term as a perturbation to this basis. As we currently have
j
diagonalizing...
As discussed above, in the 1s shell the hyperfine Hamiltonian may be written
W hf AS.I

A SxIx 
SyIy SzIz 

A SzIz 1 S               
S I I 1 S S I I
4 4

A SzIz 1 S I        
S I .
2

Considering only the restriction W hf of W hf to the degenerate subspaces with well-defined m j
 
(for the purposes of degenerate perturbation theory), the S and S terms may be ignored
since they only connect states of different m s m j . Hence, 
W hf 
AS z I z ,
and we again find that there is no need to change bases. As a result, we have the energy
spectrum
  
|n 1 , l 0 , j 1/2 , I 1/2 ; m j  1/2 , m I !
1/2 0
A 2,
4 ! "#%$'& ( )
* * *
|n 1 , l 0 , j 1/2 , I 1/2 ; m j * 1/2 , m I * (
1/2 0
A 2,
4
* +
 ,-%)'& + )
* * *
|n 1 , l 0 , j 1/2 , I 1/2 ; m j * 1/2 , m I 1/2* +
 0
A 2,
4
* ( ,-%+ )
& + )
* * *
|n 1 , l 0 , j 1/2 , I 1/2 ; m j * 1/2 , m I 1/2 * +
 0
A 2. * +
 ,-%+ )
& ( )
) . )
&
4
2
Under the assumption A 0 (valid in the regime of large B 0 ), we thus have the
following diagram [C-T et al, Ch. XII Fig. 7]:

10
Here the expressions are really only valid in the region with solid lines, but the dashed lines
provide extrapolations of the first-order perturbation results back to B 0 0. We know from
results above, of course, that the real energy diagram for small B 0 should have a singlet
and a triplet!
As it turns out, it is not so bad to consider the overall term
W hfZ W hf W Z


as a whole, and to use it directly as a single perturbation term to the fine-structure (do you
see why this is different from what we have done so far?). This removes all approximations
about the relative size of B 0 and the hyperfine term, while still treating the combination of
the two as a perturbation on H 0 W f . The resulting energy diagram looks like this [C-T et al,


Ch. XII Fig. 9]:

11
12

Das könnte Ihnen auch gefallen