You are on page 1of 7

In the Classroom

The Concept of Oxidation States in Metal Complexes


Dirk Steinborn
Institut für Anorganische Chemie, Martin-Luther-Universität Halle-Wittenberg, 06120 Halle, Kurt-Mothes-Straße 2,
Germany; steinborn@chemie.uni-halle.de

The concept of oxidation states is one of the most pow- guish between stoichiometric and electrochemical valencies
erful heuristic concepts in chemistry. It plays an important and to define oxidation numbers. Pauling described this con-
role in teaching chemistry. To assign oxidation numbers (ON) cept in the 1947 edition of his book General Chemistry (5)
of atoms in molecules, most textbooks and journal articles as follows: “The oxidation number of an atom is a number
present a set of hierarchical rules. These sets can be quite com- which represents the electrical charge which the atom would
plex (up to 16 rules to construct computer driven expert sys- have if the electrons in a compound were assigned to the at-
tems) but, in any case, are incomplete and suffer from an oms in a certain way.” This way of assignment can be de-
increasing number of exceptions (1). An alternative way is scribed for atoms in covalent compounds of known structures
to go “back to the roots” of the concept and to assign oxida- as “... the charge remaining on the atom when each shared
tion numbers of atoms in molecules on the basis of their electron pair is assigned completely to the more electronega-
Lewis structures. This system of splitting shared electrons in tive of the two atoms sharing it. A pair shared by two atoms
covalent bonds heterolytically (2), according to certain prin- of the same element is split between them” (5). Furthermore,
ciples (see below), is less frequently explained in textbooks. Pauling also states that in compounds whose structures are
Thus, from the Lewis structure of a molecule, with knowl- uncertain, the oxidation numbers may be calculated from a
edge of the electronegativities of elements, oxidation num- reasonable assignment of oxidation numbers to the other el-
bers can be assigned. In doing this it is obvious that the ements in the compounds.
concept of oxidation states can be more than bookkeeping
of electrons to balance oxidation–reduction reactions. Fur- Basic Considerations
thermore, limitations of the concept will be obvious in cases
where the electronic structure of a molecule cannot properly The oxidation number of an atom in a molecule is de-
be described by one Lewis structure. fined as the charge of that atom if the molecule would con-
sist of (monatomic) ions that may also be neutral atoms.
Historical Survey Cleavage of molecules into (hypothetical monatomic) ions
has to be based on the polarity of covalent bonds. Electrons
Originally, the valency of an atom was thought to be a in polar covalent bonds are assigned to the more electrone-
fundamental atomic property that was constant and invari- gative atom whereas those in nonpolar covalent bonds are
able (Kekulé, around 1860). The principle of constant va- equally shared between the bonded atoms. From this defini-
lence numbers was successfully applied in the second half of tion three statements follow: (i) Oxidation numbers are
the 19th century for carbon compounds in organic chemis- charges of hypothetical monatomic ions that cannot be mea-
try. However, it proved to be wrong in inorganic chemistry sured experimentally, in principle; (ii) Assignment of oxida-
(3). For instance, contrary to experimental results lower va- tion numbers of atoms in a molecule is based on its electron
lent metal halides had to be formulated as dimers (Figure 1) distribution represented by an actual Lewis structure. Ambi-
to maintain fixed valencies for metals. Thus, it was recog- guities may arise when the electronic structure of a molecule
nized that valency is not an inherent property of an atom, as cannot be properly described by only one Lewis structure or
for instance, the atomic mass is. Atoms may have variable when the polarity of a covalent bond to be split is not known;
valencies (“valence numbers”) that are indicated in the names and (iii) The concept in this simple form is limited to mol-
of compounds by terminations such as “-ous” and “-ic” (e.g., ecules whose electronic structure can be described by appro-
ferrous, ferric). Later on valence numbers were placed as Ro- priate Lewis structures. In some cases this may not be possible,
man numerals in parentheses (Stock’s method; ref 4 ), which for example, in cluster compounds and in compounds with
is still used today. electrons in highly delocalized molecular orbitals.
The progress in understanding chemical bonding and To assign oxidation numbers of atoms in molecules, the
oxidation–reduction reactions resulted in the need to distin- relevant Lewis structures have to be drawn. An integral fea-
ture of (complete) Lewis structures are the formal charges of
atoms. These are charges of hypothetical monatomic ions ob-
tained by homolytic cleavage of all covalent bonds. On the
FeCl2 FeCl3 SnCl2 SnCl4
other hand, heterolytic cleavage of polar covalent bonds and
homolytic cleavage of nonpolar bonds (as described above),
Cl Cl Cl Cl Cl Cl Cl result in formation of monatomic ions whose charges define
Fe Fe Fe Cl Sn Sn Sn the oxidation numbers of atoms in a molecule. Thus, formal
Cl Cl Cl Cl Cl Cl Cl
charges and oxidation numbers reflect the two hypothetical
borderline cases of electron distribution in a molecule, namely
Figure 1. Formulas of iron and tin chlorides according to the prin- the “pure covalent” and the “pure electrostatic” view of polar
ciple of fixed valencies for iron (trivalent) and tin (tetravalent). covalent bonds, respectively. From these definitions it follows

1148 Journal of Chemical Education • Vol. 81 No. 8 August 2004 • www.JCE.DivCHED.org


In the Classroom

that the sum of the formal charges over all the atoms of a structures 2a–c, resulting in two different sets of oxidation
molecule and the sum of the oxidation numbers over all the numbers for sulfur atoms, ON(S): +4兾0 versus +2兾+2. An
atoms of a molecule are equal to the total charge on the mol- argument to favor the first set of oxidation numbers may be
ecule. that the Lewis structure 2a does not involve a (p–d) π bond
Neither of these charges reflect the real existing charge between third-row elements. This assignment also corre-
distributions in molecules. To get the charge distributions, the sponds with the formation of thiosulfate from sulfite and sul-
concept of partial or effective charges on atoms in molecules fur and its decay in acidic solution yielding sulfur dioxide
has to be applied. They can be derived from some spectro- and sulfur. Oxidation of thiosulfate with iodine results in the
scopic measurements, such as photoelectron and Mössbauer formation of tetrathionate, 3, with oxidation numbers of +4
spectroscopic measurements. However, usually, they are cal- and +1 for the sulfur atoms. Thus, reaction 2 → 3 is under-
culated quantum chemically by population analysis that par- stood as formal oxidation of terminal sulfur in S2O32− fol-
tition the total charge among the atoms in molecules. lowed by SS bond formation. SCN− (4), OCN− (5), and
molecules like HC(O)NH2 (6) are given in the literature (1d)
Oxidation Numbers in Nonmetal Compounds as examples where sets of oxidation-number rules failed to
assign oxidation numbers. Oxidation numbers derived from
Although several Lewis structures have to be considered the relevant Lewis structures using the heterolytic bond cleav-
for SO42− (Figure 2, structures 1a–c), the assignment of oxi- age procedure are given in Figure 2. If the electronic struc-
dation numbers is unambiguous. Owing to higher electrone- ture of a molecule has to be described by a resonance hybrid
gativity of oxygen, χ(O) > χ(S), heterolytic cleavage of of two or more (nearly) equivalent Lewis structures, then the
sulfur–oxygen bonds results in the formation of hypotheti- mean values of oxidation numbers can be given. This may
cal O2 and S6+ ions (ON(O) = 2 and ON(S) = +6). This is result in nonintegral oxidation numbers clearly indicating an
not the case in the thiosulfate anion S2O32−: Consider the extension of the original concept. Examples are the oxida-
group electronegativities (6) χ(SO3) > χ(S), the sulfur–sulfur tion numbers of oxygen in the superoxo anion O2− (ON(O)
bond should be cleaved heterolytically as shown in Figure 2, = −1兾2) and the dioxygenyl cation O2+ (ON(O) = +1/2).
Ambiguities in determining oxidation numbers from
Lewis structures may arise when adjacent atoms have small
electronegativity differences and when more than one Lewis
structure has to be taken into account, also see the discus-
ON(S): + 6 +6 +6 sion in ref 2.
2− 2− 2−
O O O
Oxidation Numbers in Metal Complexes
O S O O S O O S O etc.
O O O To assign oxidation numbers of all atoms in metal com-
plexes, the same procedure as described above can be applied.
1a 1b 1c
However, frequently the oxidation numbers of central met-
als are the only ones of interest and therefore the following
ON(S): + 4 0 +2 +2 +2 +2 discussion will be restricted to these. Owing to the highly
O
2−
O
2−
O
2− electropositive character of metals, electrons involved in
metal–ligand bonds are assigned to the ligands. Thus, ligands
O S S O S S O S S etc.
(L) are cleaved off taking electrons from the M–L bonds with
O O O them:
2a 2b 2c [MLx]n → Mz + x:Lm
The oxidation number of the metal equals the charge n of
ON(S): + 4 +1 +1 + 4 complex minus the sum of the charges of all ligands (z = n −
O O
2− xm). In doing so, heterolytic cleavage of all ligands L from a
metal complex results in a (monatomic) metal ion Mz (or
O S S S S O neutral metal atom)1 whose ionic charge z defines the oxida-
O O tion number of M in the complex [MLx]n. Examples are given
in Figure 3.
3
− 2 +1
ON: 0 + 2 −3 − 2 + 4 −3 +1 +2 −3 +1
Ligand charges: 0 −1 0 0
− − O 3+
S C N O C N H C H [Co(NH3)6] [CoH(CO)4 ] [Co(CO)4] −
N
H ON(M): +3 +1 −1
4 5 6 7a 7b 7c

Figure 2. Oxidation numbers for some nonmetal compounds. The Figure 3. Oxidation numbers of metals in complexes.
curved lines show heterolytic cleavage of the bonds and the straight
vertical lines show homolytic cleavage of the bonds.

www.JCE.DivCHED.org • Vol. 81 No. 8 August 2004 • Journal of Chemical Education 1149


In the Classroom

M–L interaction n -ligands π-ligands σ-ligands

M–L σ bond

Figure 4. Classification of ligands, orbital type of L n-donor (lone pair) π-donor σ-donor
L, in metal complexes. Orbital rep-
resentations are shown schemati-
cally. Arrows are directed from M–L π bond
the (filled) donor orbital to the
(empty) acceptor orbital.
orbital type of L none π-donor π*-acceptor π*-acceptor σ*-acceptor

examples OH2, OR2, O2−, F − , ... CO, NO, PF3, olefins, dienes, H H

NH3, H , ... η1-N , ... alkynes, aro- H , SiR3 , ...
2
matics, Cp −,
η2-N2, ...

Ligands may be classified according to the types of or- CO4 (8c) ligands exist in these complexes. The ν(CO)
bitals used for metal–ligand bonds (Figure 4): ligand orbit- stretching frequencies for a series of mononuclear
als forming the σ ML bonds may be non, π, or σ bonded tetracarbonyl–metal complexes having a total of 18 valence
resulting in formation of n, π, and σ complexes,2 respectively. electrons is shown in Figure 6 (7). Successive reduction gives
Additionally, n ligands may act as π donors or as π accep- rise to substantial lowering of the CO stretching frequen-
tors. To stabilize binding of π and σ ligands, strengthening cies. The ν(CO) in the “super-reduced” chromium species
of metal–ligand bonds by π and σ back-donation, respec- is in-between the typical CO double and single bonds,
tively, is usually necessary. These additional bondings may where the CO ligands may be formulated as carbyne-like
cause ambiguities in assigning the oxidation numbers of the ligands. The number of π bonds formulated in Lewis struc-
central metals. tures (on which the determination of oxidation numbers is
based) has to be in accord with the nine-orbital rule and the
n Ligands symmetry of orbitals.3 Furthermore, valence states of central
The vast majority of ligands are n ligands (e.g., NH3, atoms (ions) after (heterolytic) cleavage of all ligands have to
py, MeCN, PR3, P(OR)3, H2O, Et2O, R2S, CO, F−, Cl−, be evaluated for proper assignment of oxidation states. As-
HO−, O2−, RS−, S2−, and H−). Typical π-donating ligands are signment of d10 valence electron configurations of M and oxi-
oxo (O2−) and fluoro (F−) ligands but this type of π bonding dation numbers ranging from 0 (M = Ni) up to 4 (M = Cr)
does not affect the assigning of oxidation numbers (Figure in complexes [M(CO)4]n (Figure 6) is consistent with all
5). This does not hold for π-acceptor ligands; CO is a typi- these aspects.
cal example. Here the electrons forming the π ML bonds Further instructive examples are nitrosyl complexes of
are d electrons from the metal. As a result of the doubly de- transition metals (8). There are two types, “linear” complexes,
generate LUMO of CO, there is the capacity for two π-type MNO 160–180, and bent complexes, MNO 120–
back-bonds. The increase of MCO bond order is associ- 140, with sp- and sp2-hybridized nitrogen atoms, respec-
ated with the decrease of CO bond order (Figure 6, 8a → tively. Relevant Lewis structures are shown in Figure 7. Of
8b → 8c). Thus, Lewis structures 8b and 8c represent a metal particular interest is the lone electron pair on N in the bent
in an oxidation number being two and four units, respec- complexes, 10a and 10b, which have σ symmetry and may
tively, higher than that in 8a. Thus (formally) CO2 (8b) and not be included into the resonance of the π system. To choose
appropriate oxidation numbers, experimental data are nec-
essary on the bending of the MNO unit and on the ex-
tent of π back-donation. Owing to overlapping, IR
spectroscopy cannot serve as a diagnostic tool for unambigu-
ous structural assignments: Typical ranges of ν(NO) in lin-
Lx M O Lx M O
ear and bent NO complexes are 1600–1950 cm 1 and
ON(O) −2 −2 1520–1720 cm 1 , respectively (for comparison,
noncoordinated NO: ν(NO) = 1878 cm1). The FeNO
Lx M F Lx M F angle in Na2[Fe(CN)5(NO)]2H2O is 178 and ν(NO) =
1944 cm1. This clearly indicates a linear-type complex, FeΙΙ
ON(F) −1 −1 and a bound NO+ ligand. [IrCl(NO)(CO)(PPh3)2](BF4) is a
bent-type complex, IrNO 124 and ν(NO) = 1680
cm–1. Hence, it has to be formulated as an IrIII complex hav-
Figure 5. Examples of π-donating ligands. ing a bound NO− ligand. The classic “brown-ring” reaction

1150 Journal of Chemical Education • Vol. 81 No. 8 August 2004 • www.JCE.DivCHED.org


In the Classroom

Lx M C O Lx M C O Lx M C O

ON(M): 0 +2 +4
2− −
Figure 6. Lewis structures and Ligand: CO CO CO4
ν(CO) stretching frequen- 8a 8b 8c
cies of some tetracarbonyl
complexes. L is a neutral
ligand.
CO C O C O

(1000 –1300 cm–1)

νCO / cm-1 2000 1800 1600 1400


[Ni(CO)4] [Co(CO)4] − −
[Fe(CO)4] 2 [Cr(CO)4]4


[Mn(CO)4]3

to detect nitrate-ions results in formation of a bent-type com- (magnetic, Mössbauer, and ESR measurements) or on quan-
plex, [FeIII(NO)(H2O)5]2+, and not a linear-type complex, tum chemical calculations. As shown in Figure 8, at first the
[FeI(NO)(H2O)5]2+, as is usually quoted in undergraduate central atom is reduced to Fe0 and is followed by reduction
textbooks (9). of the ligand up to pc4.
Structurally similar metal complexes differing only in the
number of electrons are also of interest with respect to as-
signing oxidation numbers. Such an electron variability has
been found in an unusually broad range of phthalocyanine–
metal complexes. As an example iron–phthalocyaninato com-
z
plexes, [Fe(pc)]n (H2pc = phthalocyanine), are shown in
Figure 8 (10). The question to be answered is whether the
stepwise reduction of [Fe(pc)] results in reduction of the iron
atom or of the ligand. The decision whether the additional
electrons enters a metal-centered or a ligand-centered orbital N N N
can only be made on the basis of experimental investigations
N Fe N

N N N

Lx M N O Lx M N O LxM N O
ON(M): −1 +1 +3
Ligand: NO+ NO− NO3−
9a 9b 9c
complex z ON(Fe) ligand

[FeBr(pc)] 1+ +3 pc2

LxM N LxM N [Fe(pc)] 0 +2 pc2
O +1 2−
O Li[Fe(pc)]·4.5thf 1− pc
ON(M): +1 +3 −
Li2[Fe(pc)]·5.5thf 2− 0 pc2
Ligand: NO− NO3− Li3[Fe(pc)]·8thf 3− 0 pc 3−

10a 10b Li4[Fe(pc)]·9thf 4− 0 pc4

Figure 7. Lewis structures of metal–nitrosyl complexes. L is a neu- Figure 8. Reduction of the iron–phthalocyanine complex.
tral ligand.

www.JCE.DivCHED.org • Vol. 81 No. 8 August 2004 • Journal of Chemical Education 1151


In the Classroom

π Ligands
CH2 CH2 CH2
The electron balance of π-donor–π-acceptor ligands has : Lx M
Lx M Lx M
to be handled in the same way as for n-donor–π-acceptor CH2 CH2 CH2
ligands. As an example, in Figure 9 the two relevant Lewis
structures 11a and 11b are shown for an η2-ethene complex ON(M): 0 +2
(11). Exceptionally high π back-donation results in increas- Ligand: C2H4 C2H42 −
ing the oxidation number of the metal by two units (n ver- 11 11a 11b
sus n + 2) and reducing the ligand (C2H4 versus C2H42−).
11a represents a π ethene and 11b a metallacyclopropane
complex. Real electronic structures of all η2-ethene complexes R
R R
(11) have to be described as resonance hybrids, as they are a
C C C
weighted average of these two canonical forms. To ascribe : Lx M Lx M
Lx M
oxidation numbers means to decide which formula contrib- C C C
utes most to the resonance hybrid. This decision has to be R R R
based on experimental data, especially the information gained ON(M): 0 +2
about the coordination induced lengthening of the olefinic
Ligand: C2R2 C2R22 −
CC bond and the degree of back bending of the substitu-
ents on olefinic carbon atoms (Figure 10). 12 12a 12b
Zeise’s salt, K[PtCl3(η2-C2H4)]H2O, is the classic ex-
ample of a π-ethene complex, CC 1.375(4) Å and α = Figure 9. η2-Ethene and η2-alkyne metal complexes. L is a neutral
16.2 (11), with a small degree of back-donation only as com- ligand.
parison with uncoordinated ethene, CC 1.339 Å and α =
0, shows. On the other hand, X-ray structural analyses of
[Os(CO)4(η2-C2H4)] and mer-[W(CO)3(η2-C2H4)(η4-nbd)]
(nbd = norbornadiene) exhibit metallacyclopropane structures
(12), where the CC bonds, 1.49(2) and 1.48(1) Å, respec- back
tively, in the η2-C2H4 ligands are nearly as long as that in α
bending
cyclopropane, 1.512 Å.
Analogously, η2-alkyne complexes (Figure 9, structure R2 R2
12) have to be considered as resonance hybrids between π- R R C
alkyne (12a) and metallacyclopropene (12b) complexes.4 C
M elongation
Diphenylacetylene, for example, forms complexes of both C C
types as shown by the following complexes (13): cis- R R
R2 R2
[Pt(C 6 F 5 ) 2 (η 2 -PhC⬅CPh) 2 ], CC 1.203(7) Å and
[WCl2(η2-PhC⬅CPh)L(PMe3)2], L = CO, CC 1.341(6) front free coordinated
Å; L = PMe3, CC 1.33(2) Å. Designation of these com- view side view
plexes as “π-diphenylacetyleneplatinum(II)” and “tungsta(IV)-
cyclopropene” is justified by comparison the CC bond
lengths in the complexes with those in noncoordinated Figure 10. Schematic showing of the coordination induced CC
diphenylacetylene, 1.21 Å, and 1,2-diphenylcyclopropene, bond elongation and back bending of the R groups in η2-olefin
≈1.34 Å. complexes.

σ Ligands
Oxidative addition reactions of dihydrogen to low-va-
lent metal complexes yielding dihydridometal complexes are
very common and may proceed in a concerted mechanism H H
with η2-H2 complexes as intermediates (Figure 11, structures Lx M + H2 Lx M Lx M
13–15). Such complexes have been isolated and fully char- H H
acterized spectroscopically and structurally (14). These are ON(M): 0 +2
the prototypes of σ complexes. As a result of the exception- 13 14 15
ally low donor strength of H2, stable dihydrogen complexes
will not exist without a substantial degree of donation of d H H
electrons of the metal atoms into the antibonding σ* HH Lx M Lx M
orbital. However, balance is important; a too great extent of H H
back-donation results in breaking the underlying σ HH ON(M): 0 +2
bond and oxidative addition occurs, yielding 15.
14a 14b
η2-H2 complexes 14 have to be represented by the two
Lewis structures 14a and 14b.5 The first isolated stable com-
plex was [W(CO)3(η2-H2){P(i-Pr)3}2] (Kubas, 1984) exhib- Figure 11. Oxidative addition of dihydrogen to low-valence metal
iting a significantly longer HH bond than in free H2, complexes via η2-dihydrogen complexes. L is a neutral ligand.

1152 Journal of Chemical Education • Vol. 81 No. 8 August 2004 • www.JCE.DivCHED.org


In the Classroom

VSEPR
model geometric
structure

electronic
structure
molecular topological Lewis
formula formula structure(s)
formal
charges
Figure 12. Concepts of Lewis
oxidation
structures and oxidation num- electro- numbers
bers in teaching molecular negativity
chemistry (above) with SO2
as an example (below). S
O O trigonal planar (C2v symmetry)
VSEPR: O-S-O ca. 120°
exper.: O-S-O 119.5°

S sp2 hybridized
S
SO2 O S O O O - delocalized (O-S-O) π system
- lone pair of S has σ symmetry

ON(S) = +4
etc. ON(O) = – 2

0.82(1) versus 0.74 Å. The ν(HH) stretching frequencies is, the connectivity of atoms in molecules) the Lewis formu-
in dihydrogen complexes are typically lowered by more than las are derived. They show the pattern and numbers of bonds
1000 cm1 compared with free H2. Furthermore, there are and nonbonding electrons or electron pairs and play a cen-
η2-H2 complexes with elongated HH bonds, 1.2–1.4 Å; tral role to the understanding of molecular chemistry. Lewis
for example, [ReH5(η2-H2)(PR3)2], that have to be described formulas give a first insight into geometric structures (by ap-
in terms of resonance theory by a higher contribution of ca- plying the VSEPR model) and electronic structures of mol-
nonical form 14b. ecules as well. Homolytic and heterolytic cleavage of covalent
bonds are indicated via the formal charges and the oxidation
Concluding Remarks numbers the (hypothetical) borderline cases. These are the
“pure covalent” and “pure electrostatic” views of covalent
To apply the concept of oxidation numbers means elec- bonds, respectively.
trons that are shared between atoms in molecules are assigned From the didactical point of view this is the most infor-
to a specific atom. From this it follows that oxidation num- mative way for students to understand the underlying prin-
bers, in general, do not reflect the “properties” of atoms in ciples of the concept of oxidation states7 and the model
molecules but the concept is used to systematize chemistry. character of oxidation numbers. This view neither overesti-
Furthermore, inherent in this concept is that there may be mates the meaning nor reduces the concept to a formal nu-
alternative assignments of oxidation numbers to atoms of a merical tool for bookkeeping electrons in reduction–oxidation
molecule (as it happens quite often in metal complexes). To reactions. This is especially true when the electronic struc-
select the “best” set of oxidation numbers may be—within ture of a molecule can only be properly described by a reso-
certain limits—arbitrary and may also depend on the appli- nance hybrid. By evaluating the oxidation numbers in all
cation one is interested in. If need be, experimental infor- relevant resonance forms, the students will recognize directly
mation (structural data, results from IR, Raman, or the model character of the concept and get a sense of the
Mössbauer measurements) and theoretical calculations have limitations of the concept. Thus students will learn where
to be considered. To do so, the concept of oxidation states is no oxidation numbers should be given at all because they
complementary to other “rules” (like the 18-electron rule) are meaningless or where alternative sets of oxidation num-
already in use. Furthermore, the concept proved to be a valu- bers should be discussed to understand different aspects of
able framework to systematize and to analyze coordination electronic structures of molecules under consideration. For
compounds structurally and electronically (15). the formal bookkeeping of electrons in redox equations, it
The concept of oxidation states is embedded in the teach- may also be enough to give a reasonable set of oxidation num-
ing process of molecular chemistry for early chemistry stu- bers without detailed analysis of small differences in electrone-
dents as shown in Figure 12. Starting from molecular gativities of bonding partners or of π-bonding effects. This
formulas (giving the composition of molecules) via topologi- also holds for all molecules that cannot be properly described
cal formulas6 (representing the constitution of molecules, that by Lewis structures.

www.JCE.DivCHED.org • Vol. 81 No. 8 August 2004 • Journal of Chemical Education 1153


In the Classroom

Acknowledgments Literature Cited

I am indebted to Rudolf Taube (Halle) for stimulating 1. (a) Eggert, A. A.; Middlecamp, C.; Kean, E. J. Chem. Inf.
discussions, to Amanda Elliott (Loghborough) for language Comput. Sci. 1990, 30, 181. (b) Birk, J. P. J. Chem. Educ. 1992,
polishing, and to the Fonds der Chemischen Industrie for 69, 294. (c) Calzaferri, G. J. Chem. Educ. 1999, 76, 362. (d)
financial support. Holder, D. A.; Johnson, B. G.; Karol, P. J. J. Chem. Educ.
2002, 79, 465.
Notes 2. (a) Kauffmann, J. M. J. Chem. Educ. 1986, 63, 474. (b) Woolf,
A. A. J. Chem. Educ. 1988, 65, 45. (c) Packer, J. E.; Woodgate,
1. When we do not allow a rearrangement of electrons, then S. D. J. Chem. Educ. 1991, 68, 456.
we get the valence state of Mz in the complex. If there is more than 3. Werner, A. Neuere Anschauungen auf dem Gebiete der Anorgan-
one reasonable possibility of (formal) cleavage of the ligands then ischen Chemie; Vieweg: Braunschweig, Germany, 1920.
the valence states of M in the alternatives may be inspected to find 4. Stock, A. Angew. Chem. 1919, 32, 373.
out what might be the more plausible oxidation number in this 5. Pauling, L. General Chemistry; Freeman: San Francisco, CA,
instance (16). 1947; p 173.
2. Of particular interest is the designation of the σ complex. 6. Mullay, J. Struct. Bond. 1987, 66, 1.
This is a complex where a ligand donates electrons from a σ bond 7. (a) Ellis, J. E. Adv. Organometal. Chem. 1990, 31, 1. (b) Beck,
into an empty metal orbital, as in the case of complexes of dihy- W. Angew. Chem. 1991, 103, 173.
drogen (see Figure 11 for an example). Before complexes of this 8. (a) Mingos, D. M. P.; Sherman, D. J. Adv. Inorg. Chem. 1989,
type were known, one could only distinguish between n and π com- 34, 293. (b) Johnson, B. F. G.; Haymore, B. L.; Dilworth, J.
plexes; typically in this case the former ones were frequently named R. In Comprehensive Coordination Chemistry; Wilkinson, G.,
as σ complexes. Gillard, R. D., McCleverty, J. A., Eds.; Pergamon: Oxford
3. Transition metals have nine valence orbitals [5(n – 1)d + 1987; Vol. 2, p 99.
1ns + 3np; n is the principal quantum number]. Hence, Lewis struc- 9. Wanat, A.; Schneppensieper, T.; Stochel, G.; van Eldik, R.;
tures that make use of more than nine orbitals are not valid. Fur- Bill, E.; Wieghardt, K. Inorg. Chem. 2002, 41, 4.
thermore, the number of π bonds has to be in accord with orbital 10. Taube, R. Pure Appl. Chem. 1974, 38, 427.
symmetry. 11. Love, R. A.; Koetzle, T. F.; Williams, G. J. B.; Andrews, L.
4. This is a simplified picture in that owing to the orthogo- C.; Bau, R. Inorg. Chem. 1975, 14, 2653.
nal π bonds alkynes may act as four-electron donors and as four- 12. (a) Bender, B. R.; Norton, J. R.; Miller, M. M.; Anderson, O.
electron acceptors. P.; Rappé, A. K. Organometallics 1992, 11, 3427. (b) Grevels,
5. Taking into consideration that resonance structures are de- F.-W.; Jacke, J.; Betz, P.; Krüger, C.; Tsay, Y.-H. Organometal-
fined as alternative representations of the electronic configuration lics 1989, 8, 293.
of a fixed set of nuclei, it will be evident that formula 14b (repre- 13. (a) Usón, R.; Forniés, J.; Tomás, M.; Menjón, B.; Fortuno,
senting one of the two canonical forms of an η2-H2 complex 14) C.; Welch, A. J.; Smith, D. E. J. Chem. Soc., Dalton Trans.
has another meaning different from Lewis structure 15 (represent- 1993, 275. (b) Clark, G. R.; Nielson, A. J.; Rae, A. D.;
ing a real existing dihydrido complex). Rickard, C. E. F. J. Chem. Soc., Dalton Trans. 1994, 1783. (c)
6. In topological formulas two atoms are connected by a line Nielson, A. J.; Boyd, P. D. W.; Clark, G. R.; Hunt, P. A.;
when a bond between them exists, irrespective of the bond type. Hursthouse, M. B.; Metson, J. B.; Rickard, C. E. F.;
Hence, these lines do not represent electron pairs as it is the case Schwerdtfeger, P. A. J. Chem. Soc., Dalton Trans. 1995, 1153.
in Lewis structures. 14. (a) Kubas, G. J. Acc. Chem. Res. 1988, 21, 120. (b) Crabtree,
7. Experience shows that students who have mastered writ- R. H. Angew. Chem. 1993, 105, 828.
ing a Lewis formula of a molecule are capable of assigning the oxi- 15. Cotton, F. A.; Wilkinson, G.; Murillo, C. A.; Bochmann, M.
dation numbers correctly in most cases. On the other hand, students Advanced Inorganic Chemistry; Wiley: New York, 1999.
who fail to draw a Lewis formula of a molecule successfully often 16. (a) Nyholm, R. S.; Tobe, M. L. Adv. Inorg. Chem. Radiochem.
fail to correctly use a set of hierarchical oxidation number rules. 1963, 5, 1. (b) Taube, R. In Internationales Döbereiner-
Instructors can check tutorials for the treatment of S2O82− as a useful Kolloquium; Bolck, F., Ed.; University Press: Friedrich-Schiller
example . University of Jena, Germany, 1981; p 73.

1154 Journal of Chemical Education • Vol. 81 No. 8 August 2004 • www.JCE.DivCHED.org

You might also like