Next: About this document ...
LECTURE 6
Electronic Configurations of the Elements and the Periodic Table
(Reference: Charles E. Mortimer, Chemistry: A Conceptual Approach,
3rd ed. (1975))
A pretty good way to determine the electronic configurations of the
elements is to imagine adding one electron at a time to the energy
levels of an atom. Each electron is added in accordance with the Pauli
exclusion principle and Hund's rule. Several factors determine the
energy of an electron in an orbital. These include the number of
protons in the nucleus (Z), the average distance of the orbital from
the nucleus (larger n means larger distance), the screening of the nuclear
charge by the inner electrons, and the degree to which the the orbital
penetrates the charge cloud of the other electrons and approaches
the nucleus (given by
).
A word about screening. In a multielectron
atom, the positive charge of the nucleus is partially screened from an
outer electron by the negative cloud of intervening electrons; the
electron does not ``feel'' the full nuclear charge. The screening,
however, is not uniform for all the orbitals of a given level. The orbitals
of an atom overlap the region surrounding the nucleus, and there is a
slight probability that an electron of any type orbital will be found
close to the nucleus where the electron is more fully subject to the
nuclear charge. The more an orbital penetrates the negative cloud of
the screening electrons, the more strongly the electron is attracted
to the nucleus
and the lower is its energy. For orbitals of the same level (same
value of n), the smaller the value of
, the higher the
probability of finding the electron near the nucleus. Thus, in a given
atom a 2s electron (
) is more penetrating and has
a lower energy than a 2p electron (
). For a given
level the order by energy is: s<p<d<f.
There is no standard order of sublevels, based on energy, that pertains
to all atoms. In the potassium atom (Z=19), the 4s orbital has a
lower energy than a 3d because the 4s penetrates the screening
electrons more than the 3d. In atoms of higher Z (at approximately
Z=29, Cu), this situation reverses and a 3d orbital has a lower energy
than the 4s presumably because the higher nuclear charge makes the
penetration effect (related to
) less important than the distance
of maximum electron probability (given by n).
Aufbau ordering is one good rule of thumb for deducing the electronic
configurations.
lec6aufbau.eps
One can correlate the aufbau ordering with the periodic table.
lec6periodic.eps
The transition elements are the ones where the last electron added is
a d electron. For example the first transition element is
scandium (Z=21) which has the electronic configuration
.
The rare earth or actinide elements are the ones where
the last electron added is an f electron. The alkali metals, which
are in the first column have one s electron. Lithium, sodium,
potassium are some of the alkali metals. The noble gases (like
helium, neon, and argon) are the
ones in the last column and have a filled shell. They are very stable
chemically because they have a filled shell. Atoms like filled shells,
and their chemical reactivity with other atoms is determined by how
easy it is for them to have a filled shell. For example, fluorine
and chlorine are very reactive (electronegative)
because they are just one electron short of a filled shell. This
next-to-last column has elements called halogens.
Zeeman Effect
(Reference: Robert B. Leighton, Principles of Modern Physics,
McGraw-Hill (1959).)
In lecture 2 we said that the orbital and spin angular momenta are associated
with magnetic moments:
 |
(1) |
and
 |
(2) |
where the Landé g factor g=2 for an electron. Here
is the orbital angular momentum for all
the electrons.
is the spin angular momentum
for all the electrons in the atom. Here we are assuming that
LS coupling applies. The total magnetic moment is
 |
(3) |
Notice that
is not parallel to the net angular momentum
because the orbital and spin g factors are different.
When an atom is placed in an external magnetic field
,
the orientational energy shift
is
 |
(4) |
Let's suppose that
. Then
 |
(5) |
Notice that different values of mL and mS will have different
energies. This is loosely called the Zeeman effect. It's official name
is the Paschen-Bach effect. It's true
in strong external magnetic fields (somewhat greater than 1 Tesla) where the
external magnetic field is so much larger than the atomic magnetic field
that it overwhelms the coupling between L and S.
For external magnetic fields which are weaker than the atomic magnetic
field, one cannot ignore the spin-orbit coupling
.
In this case, with
, mJ is a good
quantum number and the level splitting is given by
 |
(6) |
where the Bohr magneton
and the Landé g factor
is given by
 |
(7) |
Notice that the magnetic field splits each energy level into 2j+1components, one for each value of mJ. This splitting is called
the Zeeman effect. To get some idea of why
coupling leads to the Landé g factor, note that since
,
 |
(8) |
So
 |
(9) |
Or
 |
(10) |
where is C is a constant. Here we are using bra and ket
notation to denote the expectation value of
with respect to a state with quantum numbers n, J, L, and S.
In case you're not familiar with bracket notation, let me go over it briefly.
Recall Schrodinger's equation:
 |
(11) |
The energy expectation value En can be written as
 |
(12) |
where the integral is over the relevant coordinates-space, spin, etc.
This can also be written as
The bra vector |n> corresponds to
and the ket vector
<n| corresponds to
.
We can also regard this as a diagonal matrix element Hn,n. Suppose
we have a basis set of wavefunctions
,
, ...,
.
The diagonal matrix elements of some operator
are the
expectation values a of
:
 |
(14) |
where
n=1,2, ...,N. If the basis set of wavefunctions
are the eigenfunctions of
the Hamiltonian H, then the diagonal matrix elements are
the energy eigenvalues and the off-diagonal matrix elements are
all zero. But if the basis set of wavefunctions are not eigenfunctions
of H, then the off-diagonal matrix elements will not be zero, in general.
In this case the off-diagonal matrix elements are the
transition amplitudes from one state n to another
. The
transitions are induced by the Hamiltonian H.
 |
(15) |
The probability that there will be a transition from n to
induced by the coupling H is the square of the amplitude:
 |
(16) |
(For a Hermitian operator
,
, where
* means complex conjugate and
is the hermitian
adjoint of
.)
Transition Rates Between Atomic Energy Levels
Discussing transition probabilities brings up the topic of electron
transitions between states. An electron which occupies an excited
state
in an atom can jump to a lower energy state n by
emitting a photon whose energy
. The
photon is needed for energy conservation. Similarly an electron
can jump from a lower energy state to a higher energy state by
absorbing a photon whose energy equals the energy difference.
Classically electromagnetic radiation is produced by accelerated charges.
The most efficient way to do this is with an electric dipole antenna.
The electric dipole term is the first term in a multipole expansion.
When one treats the atom quantum mechanically, there is an electric
dipole term HED in the Hamiltonian which is responsible for
virtually all of the observed
electronic transitions. In other words the transition
rate due to HED is much higher than that due to other terms (like
the magnetic dipole and electric quadrupole terms). The transition
rate
from an initial state i to a final state f is given by
Fermi's Golden Rule:
 |
(17) |
where
corresponds to photon emission and
corresponds to photon absorption. Notice that the transition rate is
proportional to the transition probability. The
function
ensures energy conservation. The units work out to give the right units
for a rate (inverse time). Fermi's Golden Rule is one of the most
useful formulas in quantum mechanics. Often there is a group of final
states that are nearly equal in energy and for which <f|H|i> is
roughly independent of f. Then we can sum over final states which
have energy very close to E and write the transition rate as
 |
(18) |
where N(E) is the density of final states with energy E.
This is a more useful form of Fermi's Golden Rule.
Not every possible transition occurs. There are other constraints
imposed by other conservation rules. This results in selection
rules which tell us which transitions are allowed. The selection
rules for electric dipole transitions are (assuming LS coupling):
- 1.
- Only one electron jumps at a time.
- 2.
- The
value of the jumping electron must change by one unit
 |
(19) |
This is because the parity of the wavefunction must change in an electric
dipole transition.
To determine the parity of a wavefunction
,
let
and see if the wavefunction changes sign:
A parity operation changes spherical coordinates in the following way:
 |
(20) |
If the orbital angular momentum
is a good quantum number, the
parity of the wavefunction is given by
.
So if
is an even integer, the wavefunction has even parity, and if
is an odd integer, the wavefunction has odd parity. Since the
electric dipole moment
,
,
and
HED is odd under parity because
changes sign when
.
Remember that
<f|HED|i>represents an integral and integrating an odd integrand
over all space gives zero. In order for the matrix element
to be nonzero,
<f|HED|i> must have even parity, so the initial
and final states must have opposite parity.
- 3.
- For the atom as a whole, the quantum numbers L, S, J, and
MJ must change as follows:
Since a photon has angular momentum 1, the total angular momentum Jfof the final state must equal the vector sum of the angular momentum
of the photon and of the total angular momentum Ji of the initial state.
Recall from lecture 2 that when we add angular momentum j1 and j2,
the total j obeys
 |
(21) |
So
unless Ji=0. If Ji=0, Jf=1.
The fact that the sum of the
angular momenta of |i> and HED equals the angular momentum
of the final state |f> in
<f|HED|i> is an example of the
Wigner-Eckart theorem. It's also just conservation of angular momentum.
Note on
.
Usually
,
and occasionally
,
even though for the electron making the
transition
.
This is due to the
way that angular momenta are added in quantum mechanics as one can see
from (21).
j-j Coupling
When we assume the opposite case to Russell-Saunders coupling-namely,
not that there is a strong interaction of the
with one
another and the si with one another, but rather that there is
considerable interaction between each
and the si belonging
to it-we obtain the so-called j-j coupling: Each
combines with
the corresponding si to give a ji, the total angular momentum
of the individual electron. The individual ji are less strongly
coupled with one another and form the total angular momentum J of the
atom. Such coupling can be written symbolically:
 |
(22) |
There is no definite L and S for this coupling. However J remains
well defined. The same holds for M.
Let us consider, as an example, the configuration ps, which gives
3P0,1,2 and a 1P1 state on the basis of Russell-Saunders
coupling. Assuming j-j coupling, however, the p orbital has
and s1=1/2 which gives
. From the
assumption of strong coupling between
and s, these two states
can have very different energies. The
s orbital has
and s2=1/2 which gives j2=1/2.
The two states may be characterized as
(j1,j2)=(3/2,1/2) and (1/2,1/2). To the same approximation,
the analogous Russell-Saunders states are 1P and 3P. When
the small j-j coupling is taken into account, a slight splitting
of each of the two (j1,j2) states occurs. For (3/2,1/2),
J is 2 or 1; for (1/2,1/2), J is 0 or 1. For Russell-Saunders
coupling, the small spin-orbit coupling splits 3P into
3 components,
.
jj.eps
The selection rules for j-j coupling are
- 1.
- Only one electron jumps at a time.
- 2.
- The
value of the jumping electron must change by one unit
or, more generally, the parity must change.
- 3.
-
for the jumping electron, and
for
all the other electrons.
- 4.
- For the atom as a whole
Most atoms follow LS coupling rather than j-j coupling, so we
won't discuss j-j coupling further.
Next: About this document ...
Clare Yu
1999-10-14