Sei sulla pagina 1di 5

Physics 6330

14 February 2011
D. G. Ellis
Quantum States of Electrons in Atoms
Ref. Sakurai §6.4

For the standard nonrelativistic model of a helium atom or helium-like ion, we consider two
electrons in the central Coulomb potential created by a heavy, spinless nucleus of charge Ze. To
begin with we want to enumerate the available quantum states; for that purpose we temporarily
ignore the Coulomb repulsion between the electrons and consider them to be non-interacting. So
this initial model problem has the spin-independent hamiltonian

p21 p2 Zκ Zκ
H0 = H(1) + H(2) = + 2 − − (1)
2m 2m r1 r2

where κ = e2 /(4π0 ) = 1.44 eV·nm.

1 First look only at the spin states. Each electron has two spin states, so the two-electron
system has four. We can use either uncoupled basis states { | m1 , m2 i }, or coupled basis states
{ | S, MS i }, depending on whether we use eigenvectors and eigenvalues of the individual spins
~
S(1), ~
S(2) ~ = S(1)
or of the total spin S ~ ~
+ S(2) and its z component. Recall that the states are
defined by

S 2 | S, MS i = S(S + 1)h̄2 | S, MS i
Sz | S, MS i = MS h̄ | S, MS i
S(1)z | m1 , m2 i = m1 h̄ | m1 , m2 i
S(2)z | m1 , m2 i = m2 h̄ | m1 , m2 i

Note that all these states are eigenstates of S(1)2 and S(2)2 with quantum numbers s1 = s2 = 1/2.
We know how to relate these bases by using shift operators to generate the vector coupling (CG)
coefficients. Recall the results are that we have a triplet of states with total spin S = 1 and
a singlet state with total spin S = 0. If we denote spin states with the letter χ, so that
χ(S, MS ) = | S, MS i , our four coupled basis states are

Spin triplet:
χ(1, +1) = | m1 = +1/2 i| m2 = +1/2 i
r r
1 1
χ(1, 0) = | m1 = +1/2 i| m2 = −1/2 i + | m1 = −1/2 i| m2 = +1/2 i (2)
2 2
χ(1, −1) = | m1 = −1/2 i| m2 = −1/2 i
Spin singlet:
r r
1 1
χ(0, 0) = | m1 = +1/2 i| m2 = −1/2 i − | m1 = −1/2 i| m2 = +1/2 i (3)
2 2
We also recall that the expansions of the coupled in terms of the uncoupled states show that the
triplet states are all symmetric under particle interchange, while the singlet state is antisym-
metric. Thus the coupled states have a big advantage in atomic physics because of the need to
construct wavefunctions which are antisymmetric under electron interchange.

1
2 The spatial wavefunctions
If we are using (1) we really have two independent hydrogen-like ions, with each electron having
its own wavefunction of form
φnlm (~r ) = Rnl (r)Ylm (θ, φ) .
When we combine two of these, we again have the choice between coupled and uncoupled angular
momenta. For given values of the principal and orbital quantum numbers (n1 , n2 , l1 , l2 ), we can
complete the specification of the angular dependence of the two-particle wavefunction either by
giving the two magnetic quantum numbers (m1 , m2 ), or by coupling ~l1 and ~l2 to form the total
orbital angular momentum L ~ = ~l1 + ~l2 and using the coupled quantum numbers (L, ML ). For
example, if l1 = l2 = 1, we have 3 values each for m1 , m2 , so there are 9 states. When coupled, we
can have L = 2 (5 states), L = 1 (3 states), or L = 0 (1 state), again giving a total of 9. But, just
as with the spin states, we prefer the coupled states because they have definite symmetry . If we
write out the expansions as we did with the spins, we find that the L = 2 states must be symmetric
under particle interchange, because the L = 2, ML = 2 state is obviously symmetric and the other
L = 2 states are formed from them by applying lowering operators in a symmetric way. Then we
see that the L = 1 states are anti-symmetric, because we begin by making the L = 1, ML = 1
state orthogonal to the L = 2, ML = 1 state. Finally, although it is less obvious, we find that in
order for | L = 0, ML = 0 i to be orthogonal to both | L = 1, ML = 0 i and | L = 2, ML = 0 i ,
it must be symmetric.

3 Symmetry of the combined states.


Since the two-electron system must be anti-symmetric under interchange, if the spin state is
symmetric then the space state must be anti-symmetric, and vice-versa. If the two electrons have
different n, l values, then we can simply form the needed symmetric or anti-symmetric combination
of single-particle states. But if we have equivalent electrons, that is, both in the same orbital
( n1 = n2 , l1 = l2 ), then we must look to the angular momentum coupling to provide the symmetry
or anti-symmetry. If we have two equivalent electrons with l = 1 as above, this means that if the
spin state is triplet (symmetric) we can have only the L = 1 orbital combinations, and if the spin
state is singlet (anti-symmetric) then we can have L = 0 or L = 2.

4 Spectroscopic notation
When we combine the space and spin wavefunctions, in the coupled form, we describe each
two-electron state with the quantum numbers n1 , n2 , l1 , l2 , L, S, ML , MS . This way of labeling
the states is called LS-coupling, or Russell-Saunders coupling. It is widely used in the classification
of atomic energy levels. It has the advantage of giving states of definite symmetry, as we have said.
But it also has another advantage. If relativistic effects are not important, the atomic hamiltonian
is independent of spin. Thus S ~ is a conserved quantity, and the eigenstates of S 2 , Sz can also
be energy eigenstates. Of course the same can be said for the total angular momentum of the
atom, J~ = L ~ + S,
~ and its z component MJ = M . but this means that L ~ is also conserved in
the nonrelativistic theory, so we can label our energy eigenstates by the LSJ quantum numbers
L, S, J, M .
As we have seen, the energy of an isolated atom must be independent of M , so the energy
levels for a given configuration ( n1 , l1 , n2 , l2 ) are specified by the three numbers L, S, J . To
avoid having to recite these numbers repeatedly, we have the spectroscopic notation. We denote
single-electron l values by lower case letters: l = 0 → s , l = 1 → p , l = 2 → d , l = 3 → f , · · · ,
and the total L values by the corresponding upper case letters. Then the S and J values are given
by a prefix and suffix, in the form (2S+1) LJ . Thus for example the lowest energy level of the carbon
atom is
1s2 2s2 2p2 3 P0

2
5 Example: p × p = S + P + D
First L=2: | D, +2 i = | p, +1 i | p, +1 i
r r
1 1
| D, +1 i = | p, +1 i | p, 0 i + | p, 0 i | p, +1 i
2 2
r r r
1 2 1
| D, 0 i = | p, +1 i | p, −1 i + | p, 0 i | p, 0 i + | p, −1 i | p, +1 i
6 3 6
r r
1 1
| D, −1 i = | p, 0 i | p, −1 i + | p, −1 i | p, 0 i
2 2
| D, −2 i = | p, −1 i | p, −1 i
Note all these D states are symmetric.
r r
1 1
Now L=1: | P, +1 i = | p, +1 i | p, 0 i − | p, 0 i | p, +1 i
2 2
r r
1 1
| P, 0 i = | p, +1 i | p, −1 i − | p, −1 i | p, +1 i
2 2
r r
1 1
| P, −1 i = | p, 0 i | p, −1 i − | p, −1 i | p, 0 i
2 2
Note all these P states are antisymmetric.
r r r
1 1 1
Finally L=0: | S, 0 i = | p, +1 i | p, −1 i − | p, 0 i | p, 0 i + | p, −1 i | p, +1 i
3 3 3
Note this S state is symmetric.

6 Example: np2 = 1
S + 3P + 1D
Now we can combine the above cases. To form states of the np2 configuration, we write
X
| L, S, J, M i = h L, ML , S, MS | L, S, J, M i | L, ML i | S, MS i . (4)
ML ,MS

Note that each state in this expansion is a product of a space function and a spin function,
| Ψ i = | φ(L, ML ) i | χ(S, MS ) i .
In order to satisfy the Fermi-Dirac statistics, | Ψ i must be antisymmetric under interchange of
the two electrons. Thus if | φ(L) i is symmetric, as for D or S terms, then | χ(S) i must be
antisymmetric, namely we need S = 0. But for P terms, | φ(L) i is antisymmetric, so we need the
symmetric (S = 1) spin function.
Thus we finally see that the configuration np2 has only three allowed LS-coupled terms, namely
1
S, 3P , and 1D.

7 Hund’s rule.
Before quantum mechanics it was noticed that within a configuration, the term of maximum S
usually has the lowest energy. For example in carbon and silicon ( np2 ground configuration) the
ground term is 3P , while in nitrogen and phosphorus ( np3 ), it is 4S.
The reason for this is the fact that the state of maximum S has a symmetric spin function,
and so it must have an antisymmetric space function. But this means that the space wavefunction
must be zero when the electrons are at the same point, and therefore it will be small when they
e2
are close together. But this means that the Coulomb repulsion energy has a smaller average
r12
value, and so the state has a lower total energy.

3
8 Appendix: Some Atomic Structure Terminology.
(a) Electron orbitals. An orbital is a single-electron state. Using spectroscopic notation,
a state with n = 3, l = 2 for example, is called 3d for convenience. This 3d orbital is taken to
include states with all possible values of ml , ms for the given values n = 3, l = 2. The possible
hydrogen orbitals are thus 1s, 2s, 2p, 3s, 3p, 3d, 4s, 4p, 4d, 4f, · · · .
(b) Configurations. For atoms other than hydrogen we may consider labeling the states by giving
the set of quantum numbers n, l, ml , ms for each electron. In fact, the physical N -electron state may
not be well represented in this way, but it is a reasonable starting point for a labeling system. The
Pauli Principle states that no two electrons can have the same set of these 4 quantum numbers. Thus
there is a maximum number (2(2l+1)) of electrons which can occupy each orbital. These numbers
are familiar from chemistry: s : 2, p : 6, d : 10, · · · . An electronic configuration is specified by
giving the orbital for each electron, or in other words by giving the occupation number of each
orbital. Thus the ground state of the carbon atom has the configuration (1s2 2s2 2p2 ); sometimes
filled orbitals (those containing the maximum allowed number of electrons) are omitted and this is
written simply (2p2 ).
(c) Shells. Since the principal quantum numbers determine the energy to a first approximation,
all states with the same n value are said to belong to a given shell , and as we consider increasing
values of N , the shells are filled “from the bottom up”, in order of increasing energy, and increasing
distance from the nucleus. However this approximation soon breaks down — for example the
ground configuration of potassium (Z = 19) is (3s2 3p6 4s) rather than (3s2 3p6 3d). This effect
which causes orbitals of low l to have lower energies is called penetration, since electrons in these
orbitals can penetrate closer to the nucleus. The shells can designated by upper case letters; this
notation is less common in optical spectroscopy than in chemistry or X-ray spectroscopy. The letter
K denotes the n = 1 shell; L, n = 2; M, n = 3, etc.
In summary, the atomic shell structure can be represented as follows:

SHELL ORBIT ALS ST AT ES


K 1s2 2
L 2s2 2p6 8 (5)
M 3s2 3p6 3d10 18
N 4s2 4p6 4d10 4f 14 32

and so forth.
(d) LS coupling. For atoms with two or more electrons, there are many possible quantum states,
and we need a systematic way of naming them. The most commonly used system is LS coupling ,
which can be summarized as follows. We combine all the orbital angular momenta of the individual
electrons into a total orbital angular momentum L = l 1 + l 2 + · · · , and we also combine
all the spin angular momenta of the individual electrons into a total spin angular momentum
S = s 1 + s 2 + · · · . Only after that are these two added together to get the total angular
momentum J = L + S . This is called L-S coupling or Russell-Saunders coupling . Since
the energy of an isolated atom cannot depend on the direction of the total angular momentum, the
three quantum numbers L, S, J are sufficient to determine the energy of the atom, and are thus a
convenient label for atomic energy levels. A good reference to begin the study of the structure of
complex atoms is the small book by Woodgate.
For example, in an excited state of helium, we may have the configuration 1s2p, so that there
are four angular momenta to combine: l1 = 0, l2 = 1, s1 = s2 = 1/2. We know there will be
12 possible states, since there are three possible values of ml for the p electron, and for each of
these we can have 4 possible combinations of ms1 and ms2 , namely up-up, up-down, down-up, and
down-down. We describe these 12 states in L-S coupling as follows: Adding 0 and 1 gives only one

4
possible value of L, namely L = 1. But there are two possible values of S, since we can have either
|s1 − s2 | = 0 or s1 + s2 = 1, but nothing in between. Thus we have states with (L = 1, S = 0) and
states with (L = 1, S = 1). Finally we form the J values. For S = 0 we must have J = L = 1;
but for S = 1 we can have J running from |L − S| to L + S in integer steps, namely J = 0, 1, 2.
Putting all this together we see that we expect 4 energy levels within the 1s2p excited configuration
of helium. We label them {L = 1, S = 1, J = 0}, {L = 1, S = 1, J = 0}, {L = 1, S = 1, J = 0}, and
{L = 1, S = 1, J = 0}.
For convenience, so that we do not have to write out all these quantum numbers, we extend
our spectroscopic notation as follows. We use an upper case letter to represent L according to
the previous scheme S, P, D, F, · · · ; then we use the integer 2S + 1 as a super-prefix, and J as a
sub-suffix. Without J, using only L, S, this is called a term; including J gives a level . Within the
level are 2J +1 states, which have the same energy since they differ only in M , that is only in the
direction of J . The energy is often determined mainly by L, S; the week dependence on J, is called
the fine structure. Finally we pronounce the spin state as “singlet” for 2S+1 = 1, “doublet” for
2S +1 = 2, “triplet” for 2S +1 = 3, and so forth. If the fine structure is small, this terminology
tells us to expect a grouping of 1, 2, 3, · · · closely-spaced energy levels.
So for our helium example we have 2 terms, “singlet P” 1P , and “triplet P” 3P , with the 4
levels 1P1 , 3P0 , 3P1 , 3P2 . Finally, note that there really are 12 states in these 4 levels: if we count
the possible values of M = −J, · · · , +J, we see that 1P1 has 3 states, 3P0 has 1, 3P1 has 3, and 3P2
has 5.
In more complex cases, it is important to make use of the simplicity afforded by a closed shell .
If we have the maximum allowed number of electrons in a shell, they always couple together to
give a system with L = 0 and S = 0. Thus for example, the beryllium (Z = 4) atom has a set
of excited states 1s2 2s2p 1P , 1s2 2s2p 3P , which are exactly analogous to the helium 1s2p states
described above.
(e) Equivalent electrons. Finally there is a complication for configurations containing equiva-
lent electrons, that is two or more electrons in the same orbital (same values of n, l). For example,
in the 2p3p configuration, the terms are 1S, 3S, 1P , 3P , 1D, 3D. These are determined by considering
all possible values of L, S when we add one electron with l1 = 1, s1 = 1/2 to another electron with
l2 = 1, s2 = 1/2. However for the 2p2 configuration, some of these terms are forbidden. The reason
is that electrons are fermions, that is they satisfy Fermi-Dirac statistics, which means that any
state of 2 or more electrons must be antisymmetric — it must change sign if we interchange the
quantum numbers of any pair of electrons. The Pauli Exclusion Principle is just one consequence
of the fact that electrons are fermions. Another consequence is that certain of the terms that occur
for 2p3p do not occur for 2p2 . In fact, the configuration 2p2 has just 3 terms, namely 1S, 3P , and
1D.

(f ) Brief summary of terminology. The states of an atom are traditionally named in four
steps, which we illustrate with the ground configuration of carbon.
First we specify the configuration: 1s2 2s2 2p2 (15 states)
then the terms: 1S (1 state), 3P (9 states), 1D (5 states)
then the levels and states:
1S (1 state): M =0
0
3P (1 state): M =0
0
3P (3 states): M = −1, 0, +1
1
3P (5 states): M = −2, −1, 0, +1, +2
2
1D (5 states): M = −2, −1, 0, +1, +2
2

Potrebbero piacerti anche