You are on page 1of 3

In the Classroom

The Genius of Slater’s Rules


James L. Reed
Department of Chemistry, Clark Atlanta University, Atlanta, GA 30314; jlreed@cau.edu

More than 60 years ago Slater proposed a set of very


simple rules for the computation of the effective nuclear
charge experienced by an electron in an atom (1). Since that
time these rules have been used in applications ranging from
evaluating atomic overlap integrals (2) to establishing tables
of atomic radii (3). Recently they were used to provide new
insights into the often elusive nature of chemical hardness
(4 ). Because Slater’s procedure provides for the computation
of one-electron energies rather than orbital energies, it is es-
pecially well suited for instructional purposes and has been
included in a number of textbooks (5). The need for simple
activities that illustrate and reinforce an intuitive understanding
of atoms and their properties is needed throughout the
undergraduate curriculum. However, one difficulty with
Slater’s original procedure has been that it fails to yield the
level of detail needed to illustrate important concepts. With
relatively minor modifications, Slater’s method would provide
to both students and teachers a model of the atom that is
instructive as well as surprisingly simple and intuitive.

The Basic Concepts


Slater’s method and its implementation require only a
reasonable mastery of three concepts, which are already part of
the first-year curriculum. The first of these is that the electron–
electron interactions can be reasonably described as shielding
of the nuclear charge by each of the other electrons, or
Figure 1. A diagrammatic presentation of the modified Slater’s Rules.
Σ
all electrons

Zi*=Z – ci = Z – cij (1)


i≠ j
through barium (no f electrons). These rules are applicable
to positive, negative, and fractionally charged atoms. The rules
where Zi* is the effective nuclear charge experienced by elec-
given below are summarized in Figure 1.
tron i and cij is the shielding of electron i by electron j. Once
this is determined, the energy of the electron may be com- The Rules for Light Elements
puted, using the relationship Larger Shell. An electron is not shielded by any electron
2 in a larger shell (c ij = 0.00).
{1 Z i* Same Shell. Both s and p electrons in the same shell shield
e i = {1312 kJ mol (2)
ni each other: 0.3228.
Next Smaller Shell. The shielding of an s electron by an s
This relationship most often appears in first-year texts in
or p electron in the next smallest shell is 0.8366. The
its more complicated form as the Rydberg equation. Finally, shielding of a p electron by an s or p electron in the next
since Slater’s procedure yields one-electron energies, the total smallest shell is 0.9155.
electronic energy is the sum of the one-electron energies: Inner Shells. An electron is completely shielded by elec-
E = e1 + e2 + e3 + … + eN + epairing (3) trons in shells that are smaller than the next smallest (c ij
= 1.00).
The pairing energy was not explicitly included in Slater’s The pairing energy for p electrons is 1360.1/n2 kJ/mol.
original formulation, but its explicit consideration is necessary
for a number of applications. The Rules for Heavier Elements
Same Shell. In the same shell the shielding of a d electron
Shielding by s and p electrons is 0.8933, but the shielding of an s
or p electron by a d electron in 0.0352. The shielding of
There are potentially a very large number of shielding d electrons by d electrons is 0.3228.
constants. However, Slater found that, because many were Smaller Shells. The shielding of both s and p electrons by a
virtually equal, the number could be reduced to only three. d electron in the next smaller shell is 0.9143. A d electron
We have refined the scheme to use only three for the lighter is completely shielded by any electron in a smaller shell.
elements (no d electrons) and a total of eight for the elements The pairing energy for d electrons is 9610.7/n2 kJ/mol.

802 Journal of Chemical Education • Vol. 76 No. 6 June 1999 • JChemEd.chem.wisc.edu


In the Classroom

Evaluation of the Shielding Constants This is one of the more involved computations, but it will
These shielding constants are the ones which yield the serve to illustrate one of the subtleties involved in the aufbau
best least squares (5) agreement between the experimental (6 ) process. The ordering of the energy levels of the fourth-period
and computed first ionization energies for elements hydro- elements has been the subject of some controversy (8–11).
gen through xenon. Thus for atom M the ionization energy The ordering of the one-electron energies that the model
was computed via yields is 3s < 3p < 3d < 4s, which would seem to suggest that
potassium should be the first transition metal. However, in
IE(M) = E(M+) – E(M) (4) spite of this ordering, the configuration that makes potassium
where the energy is evaluated using eqs 1–3 (see the sample an alkali metal is the lower energy configuration. The accepted
calculation below for details). The first ionization energies configuration is favorable not because of the lower energy of
are known very accurately for hydrogen through xenon, which the outermost electron, but rather because of the poorer
are the elements having only s, p, and d electrons. For most shielding of the 3s and 3p electrons by the 4s electron. This
elements the ionization of d electrons occurs only for the third lowers their energies and thus lowers the total energy. As one
and subsequent ionizations of the transition metals. These were can see, the computations are quite simple both in concept
used to determine the shielding constants for d electrons. The and in execution. They are actually simpler than many of the
validity of both the model and the shielding constants has computations that general chemistry students are expected
been evidenced by their ability to reproduce the experimental to do routinely.
first ionization energies and successive ionizations of the same
atom (7). Applications
Sample Calculation Slater’s rules find application throughout the undergradu-
A potassium atom has been selected for the sample cal- ate curriculum. It should be emphasized that the activities and
culation. The configuration we have selected is presentation format should be tailored to the audience. For
this reason, in addition to the text form of the rules, an
1s22s22p63s23p63d1 equivalent graphical presentation has been included (Fig. 1).
For processes involving only valence electrons the core electron Furthermore the rules themselves have been divided into
energies do not change. Thus for most exercises the number those for light elements and those for heavier elements with
of computations is greatly reduced, and students need concern this same idea in mind. Finally, students need not be asked
themselves with only one shell. Of course to illustrate the to do all the types of computations done in the sample cal-
shell structure of an atom or core ionizations (XPS or ESCA), it culation. The instructor may choose to give the students the
would be necessary to do computations for core electrons also. effective nuclear charges or one-electron energies.
The determination of the shielding of a 3s electron is Aufbau
illustrated diagrammatically below along with the computa-
tions for all the third-shell electrons. Generating electron configurations is the first activity
that students undertake after learning about quantum numbers.
1.00 0.0352 Because students have already been introduced to the basic
concepts of thermochemistry, they can be shown or asked to
1s2 1s22p6 3s2 3p6 3d1
carry out activities which show that the ground state con-
figuration simply results in the maximum exothermicity for
the aufbau or filling process. Because pairing energies can be
0.8366 0.3228 computed, it can be easily shown that pairing electrons in
the same orbital may be energetically unfavorable.
Z *3s = 19 – [(1)(0.0352) + (7)(0.3228) + (8)(0.8366) + (2)(1.00)]
Z *3s = 8.0124 e3s = { 9,358.7 kJ/mol Ionizations and Energy Level Diagrams
*
Z 3p = 19 - [(1)(0.0352) + (7)(0.3228) + (8)(0.9155) + (2)(1.00)]
Energy level diagrams are frequently used to facilitate
Z *3p = 7.3812 e3p = { 7,942.3 kJ/mol
learning. Students may be provided with or asked to prepare
Z *3d = 19 - [(8)(0.8933) + (8)(1.00) + (2)(1.00)]
them. The energy level diagrams for sulfur, its cation, and its
Z *3d = 1.8536 e3d = { 500.9 kJ/mol
anion are given in Figure 2. The ionization process involves
epairing = 1360.1/9 = 151.1 kJ/mol
moving an electron from the highest occupied orbital to the
The electronic energy (valence electrons only) is thus zero-energy state. Although it would seem that the ionization
E[3s23p63d1]= (2)({ 9358.7)+ (6)({ 7942.3) + (1)({ 500.9) + (3)(151.1) energy should equal the difference in these two energies,
E[3s23p 63d1 ]= { 66,418.8 kJ/mol except for the group 1 elements, the experimental ionization
energies are much smaller. However, comparing the energy-
To allow for comparison, the effective nuclear charges and level diagrams for the atom and its cation reveals that the
one-electron energies for the configuration very endothermic ionization is accompanied by an exothermic
1s22s22p63s23p64s1 relaxation of the remaining valence electrons. Similarly, one
might expect that the ionization energy and electron affinity
have been computed to yield of an atom would have similar values. The small values of
e3s = { 9,441.1 Z *3s = 8.0476 the experimental electron affinities can be understood to arise
e3p = { 8,018.2 Z *3p = 7.4164 from not only the increase in the energy of the remaining
e4s = { 436.5 Z *4s = 2.3072 electrons, but also the increase in the acceptor orbital energy
E[3s23p 64s1] = { 66,974.9 during the process.

JChemEd.chem.wisc.edu • Vol. 76 No. 6 June 1999 • Journal of Chemical Education 803


In the Classroom

which is simply the average of the one-electron energies of


the component orbitals. The promotion energy is then

E(promotion) = E(ground state) – E(valence state) (6)

The following are just a few examples. Although the


atomization energies of sets of compounds of the type MX n
suggests that the bonds become weaker as n increases, this
apparent difference can be shown to arise primarily from
differences in the promotion energies, and the bonds are
essentially the same. In addition, it can be shown that the
instability of noble gas compounds is not the result of weaker
bonds, which would be counter to the periodic trend in bond
strength, but rather the result of very large promotion energies.

Figure 2. Energy level diagrams for sulfur, its cation, and its anion. In Closing
I feel, as do many others, that most general chemistry
Spectroscopy courses attempt to include far too much diverse material.
The optical transitions among the configurations of atoms On the other hand, simple computations are currently used
cannot be observed experimentally, but can be computed to enhance student understanding of topics such as stoichi-
from the spectroscopic term energies (12). The energies for the ometry, kinetics, and equilibrium but not atomic structure.
transitions among the configurations computed using Slater’s It would seem that a few very simple computations would
method agree quite well with the experimentally computed be appropriate for beginning students. It should also be noted
transition energies. The ionization of core electrons by high- that all the concepts (eqs 1–3) are already among those
energy radiation has been used to probe the electronic structure beginning students are expected to master. Finally, once having
of atoms, molecules, and atoms in molecules. The XPS and developed this simple skill, both students and teachers can
ESCA chemical shifts computed for core ionizations agree use these “back of the envelope” calculations to enhance
quite well with those observed experimentally (13). It may not, instruction in other topics in later courses.
however, be immediately obvious that the XPS and ESCA
Acknowledgment
chemical shifts should correlate with atomic charge (14). The
XPS and ESCA energies computed for fractionally charged I wish to acknowledge the support of the Army Research
atoms not only yield this same type of correlation, but the Office, contract DAAA15-94-K0004.
exercise actually reveals that the XPS and ESCA chemical
shift–atomic charge dependence arises from the differences Literature Cited
in the relaxation energies of the valence electrons during
photoionization. 1. Slater, J. C. Phys. Rev. 1930, 36, 57.
2. Mulliken, R. S.; Rieke, C. A.; Orloff, H. J. Chem. Phys. 1949,
Periodicity and Periodic Trends 17, 1248.
Many of the properties of isolated atoms as well as atoms 3. Slater, J. C. J. Chem. Phys. 1964, 41, 3199.
in molecules exhibit periodic behavior and trends. All these 4. Reed, J. L. J. Phys. Chem. 1997, 101, 7396.
can be related directly or indirectly to trends in effective 5. Douglas, B. E.; McDaniel, D. H.; Alexander, J. J. Concepts and
nuclear charge and principal quantum number. One of the Models in Inorganic Chemistry, 3rd ed.; Wiley: New York, 1994;
strengths of Slater’s model is that the effective nuclear charge p 40. Sharp, A. G. Inorganic Chemistry; Longman: New York,
is directly interpretable in terms of the configuration, nuclear 1981; p 70. Porterfield, W. W. Inorganic Chemistry: A Unified
charge, and types of electron–electron interactions. Thus the Approach, 2nd ed.; Academic: San Diego, 1993; p 23.
model can assist in the interpretation of the periodic behavior 6. Nelder, J. A.; Mead, R. Comput. J. 1965, 7, 308.
of such properties as ionization energies, electron affinities, 7. Huheey, J. E. Inorganic Chemistry, 3rd ed.; Harper-Row: New
electronegativities, various atomic radii, bond distances, overlap York, 1983; p 43.
integrals, and even chemical properties such as acid–base 8. Carlton, T. S. J. Chem. Educ. 1978, 55, 2.
strengths and oxidation–reduction potentials. 9. Carlton, T. S. J. Chem. Educ. 1979, 56, 767.
Promotion Energy 10. Pilar, F. L. J. Chem. Educ. 1979, 56, 767.
11. Melrose, M. P.; Scerri, E. R. J. Chem. Educ. 1996, 73, 498.
The energetics of the promotion of atoms to their valence
12. Brink, C. D. J. Chem. Educ. 1991, 68,376.
states or to hypervalent states (expanded octet) is often
13. Tudela, D. J. Chem. Educ. 1993, 70, 956.
determinant in both structural and reaction chemistry. In
14. Moore, C. E. Atomic Energy Levels; National Bureau of Standards
the central field approximation it is easily shown that the
Circ. 467, Vol. I; National Bureau of Standards: Washington, DC,
one-electron energy (expectation value) of any hybrid orbital,
1948.
spx d y, is given by
15. Ghosh, P. K. Introduction to Photoelectron Spectroscopy; Wiley: New
espxdy = 1 e + x ep + y ed York, 1983; p 10.
(5)
1+x +y s 16. Ibid., p 60.

804 Journal of Chemical Education • Vol. 76 No. 6 June 1999 • JChemEd.chem.wisc.edu

You might also like