Physical Inorganic Chemistry - A Coordination Chemistry Approach PDF
Physical Inorganic Chemistry - A Coordination Chemistry Approach PDF
Physical Inorganic Chemistry - A Coordination Chemistry Approach PDF
Physica Inorganic
Chemistry
A Coordination Chemistry Approach
S. F. A. KETTLE
Professorial Fellow, University of East Anglia, and
Adjunct Professor, Royal Military College, Kingston, Ontario
3.1 Nomenclature 24
4
Preparation of coordination
3.2 Coordination numbers 31
compounds 51
3.2.1 Complexes with coordination numbers
one, two or three 32
3.2.2 Complexes with coordination number
4.1 Introduction 51
four 33 4.2 Preparative methods 52
3.2.3 Complexes with coordination number
five 35 4.2.1 Simple addition reactions 52
3.2.4 Complexes with coordination number 4.2.2 Substitution reactions 54
six 38
4.2.3 Oxidation-reduction reactions 58
3.2.5 Complexes with coordination number
seven 38 4.2.4 Thermal dissociation reactions 61
3.2.6 Complexes with coordination number 4.2.5 Preparations in the absence of oxygen 62
eight 39
4.2.6 Reactions of coordinated ligands 65
3.2.7 Complexes with coordination number
nine 41 4.2.7 The trans effect 68
3.2.8 Complexes of higher coordination 4.2.8 Other methods of preparing
number 41 coordination compounds 69
viii 1 Contents
····························································································································································································
5 7.8 Tetrahedral complexes 148
Appendix 6 Appendix 11
Ligand tT group orbitals of an High temperature superconductors 472
octahedral complex 449
Appendix 7 Appendix 12
Tanabe-Sugano diagrams and some Combining spin and orbital angular
illustrative spectra 455 momenta 477
Appendix 8
Appendix 13
Group theoretical aspects of band
Bonding between a transition metal
intensities in octahedral complexes
atom and a en Rn ring, n = 4, 5 and 6
459
4 79
Appendix 9
Determination of magnetic Appendix 14
susceptibilities 462 Hole-electron relationship in
spin-orbit coupling 484
Appendix 10
Magnetic susceptibility of a
tetragonally distorted dg ion 466 Index 487
Foreword
GEORGE CHRISTOU
Indiana University, Bloomington
comment and information. I am particularly indebted to the Rev. Dr. lain Paul who,
in his own inimitable manner, worked through every sentence and made a multitude
of suggestions for improvement and clarification. Defects, errors and omissions, of
course, are my own responsibility.
S.F.A.K.
Introduction
geometry and electronic structure, the link between the two commonly being
provided by group theory.
Complexes are formed by both transition metal and non-transition
elements. Indeed, at the present time all compounds of transition metal
ions, with very few exceptions, are regarded as complexes. However, despite
the argument given above, the simple donor-acceptor bond approach does
not seem immediately applicable to coordination complexes of the transi-
tion metals, since the molecular geometry does not depend greatly on the
number of valence shell electrons-and, so, on the number of empty
orbitals. As will be seen in Chapter 7, in the simplest model of the bonding
in transition metal complexes electron donation is not even considered to
be involved, a molecule being regarded as held together by electrostatic
attraction between a central transition metal cation and the surrounding
anions or dipolar ligands. However, in more sophisticated discussions of
the bonding (Chapters 6 and 10) the donor-acceptor concept is largely
reinstated for these compounds. So we may conveniently (but not always
correctly) regard a coordination compound as composed of (a) an electron
donor (ligand or Lewis base), an individual atom or molecule which possesses
non-bonding lone-pair electrons but no low-lying empty orbitals; and (b) an
electron acceptor (metal atom, cation or Lewis acid) which possesses a
low-lying empty orbital. As in many other areas of chemistry, we shall often
be particularly concerned with the pair of electrons that occupy the highest
occupied molecular orbital (the HOMO) of the electron donor. This is
matched by an interest in the lowest unoccupied molecular orbital of the
electron acceptor (the LUM0). 2 The donor atom of a ligand is usually of
relatively high electronegativity and the acceptor atom is either a metal or
metalloid element.
Chapters 2-4 are full of examples of ligands and coordination com-
pounds and the reader can gain an impression of the field by quickly
thumbing through them. The field is not as complicated as it may appear,
although it will rapidly become evident that at the present time some rather
unusual organic molecules are increasingly being used as ligands and that
neither the methods of preparation nor the molecular geometries formed
need be quite as simple as for the examples given above. Indeed, part of
the current fascination of the subject lies in the elegance of many of the
complexes which are currently being studied. Complexes in which the metal
atom is totally encapsulated, as in the sepulchrates; those in which it is at
the centre of a crown (crown ether complexes, for instance); those in which
it is surrounded by two ligands which interleave each other (complexes of
catenands); those in which it is at the centre of a stockade-like ligand
(picket-fence complexes) and so on. By such means it is proving possible
to design highly metal-specific ligands, which offer the prospect of selective
ion extraction from, for example, low-grade ores or recycled materials.
The future importance of such possibilities in the face of ever-declining
natural resources can scarcely be overestimated. Similarly, the use of such
complexes in small-molecule activation will surely be of vital importance--
for instance, in the fixation of gaseous nitrogen and the synthetic use of
hydrocarbon species which would otherwise be used as fuels.
2 Because the basics of the subject were developed before use of the HOMO and LUMO
terminology became widespread these labels are scarcely to be found in the relevant literature.
4 1 Introduction
Inevitably, current research tends to focus on the unusual and the exotic
and so, since a book such as this attempts to reflect something of current
work, tends to make the subject appear less straightforward than it really
is. Perhaps it is helpful to recognize that even at a simple level, problems
of definition can occur. Thus, an uncharged compound containing a main
group metal or metalloid element bonded to a methyl group is not usually
viewed as a complex in which CH 3 functions as a ligand, although the
CH 3 group is isoelectronic with ammonia, a molecule which is frequently a
ligand. So, compounds such as Zn(C 2 H 5 h and Si(CH 3 ) 4 would not usually
be considered complexes. However, successes in the synthesis of transition
metal-methyl compounds means that there has been a change in attitude
and that these, too, are now regarded as complexes containing the CH 3
group as ligand. The question of whether they should be considered
as complexes of CH3 is not usually regarded as of great importance. A
similar ambiguity is that although the manganate ion MnOi- would be
considered a coordination complex (of Mn 6 + and 0 2 -) the sulfate anion
SOi- would not. Evidently, we have reached the point at which history
and tradition, as well as utility, colour the definition of a coordination
compound.
The father of modern coordination chemistry was Alfred Werner, who
was born in 1866 and lived most of his life in Zurich. At the time it was
known that the oxidation of cobalt(II) (cobaltous) salts made alkaline with
aqueous ammonia led to the formation of cobalt(III) (cobaltic) salts
containing up to six ammonia molecules per cobalt atom. These ammonia
molecules were evidently strongly bonded because very extreme conditions-
boiling sulfuric acid, for example-were needed to separate them from the
cobalt. There had been considerable speculation about the cobalt-ammonia
bonding and structures such as
/NH 3 -CI
Co--NH 3 · NH 3 · NH 3 · NH 2 -CI
"'-NH 3 -CI
which today look quite ridiculous (although based on the not unreasonable
hypothesis that, like carbon, nitrogen can form linear chains) had been
proposed for the cobalt(III) salt CoN 6 H 18 CI 3 (which we would now write
as [Co(NH 3 ) 6 ]CI 3 ). Werner's greatest contribution to coordination chem-
istry came in a flash of inspiration (in 1893, at two o'clock in the
morning) when he recognized that the number of groups attached to an
atom (something that he referred to as its secondary valency) need not
equal its oxidation number (he called it primary valency). Further, he
speculated that for any element, primary and secondary valencies could
vary independently of each other. The chemistry of the cobalt(III)- ammonia
adducts could be rationalized if in them cobalt had a primary valency of
three, as in CoCI 3 , but a secondary valency of six, as in [Co(NH 3 ) 6 ]CI 3 . The
term secondary valence has now been replaced by coordination number and
primary valency by oxidation state but Werner's ideas otherwise stand largely
unchanged.
Subsequently, Werner and his students obtained a vast body of experi-
mental evidence, all supporting his basic ideas. They further showed that
in the complexes they were studying the six coordinated ligands were
Introduction I5
arranged octahedrally about the central atom (Figs. 1.2 and 1.3). Werner
was awarded the Nobel prize for chemistry for this work in 1913. Some
measure of his stature and work is provided by the fact that in one field
(that of polynuclear cobaltammine complexes) there has, to this day, been
scarcely any addition to the list of compounds he prepared.
Most textbooks discuss transition metal complexes separately from
those of the main group elements. There is, in fact, much in common
between the two classes and, whenever possible, we shall treat them as one.
However, complexes of the transition metal ions may possess an incomplete
shell of d electrons which necessitate separate discussion. This character-
istic makes it particularly useful to determine the magnetic and spectral
properties of members of this class of complexes and the exploration of
these properties will require separate chapters devoted to them. In a similar
way, complexes of the lanthanide and actinide elements, with, typically, an
FJ&. 1.2 An octahedral complex ML., where M incomplete shell off electrons tucked rather well inside the atom and away
is represented by the central white atom and
the ligands L each by a shaded atom. A regular from the ligands- and so behaving rather as if they ·are in an isolated
octahedro~ne is shown in Rg. 1.3--has atom-require their own discussion.
eight faces (each an equilateral triangle) and six The water-soluble ionic species of transition elements such as chromium,
equivalent vertices. In an octahedral complex
the ligands are placed at the vertices. In Rg. manganese, iron and copper seem to exist in aqueous solution as, for
1.2, and in similar diagrams throughout this example, [Cr(H 2 0) 6 ]3+, [Mn(H 2 0) 6 ]2+ and [Fe(H 2 0) 6 ]2+ . That is, it is
book, the perspective is exaggerated (the more accurate to talk of 'the aqueous chemistry of the [Cr(H 2 0) 6 ]3+ ion'
central four ligands lie at the comers of a
square) and all ligand atoms are the same size. than of 'the aqueous chemistry of the Cr3+ ion '. Similarly, in solid FeCI 3 ,
In this example, all six ligands are identicaL the iron atoms are not attached to three chlorines but, octahedrally, to six
Even if they are not, provided the geometrical (each chlorine is bonded to two iron atoms). We have already encountered
arrangement shown in Fig. 1.2 is more-or-less
maintained, the complex is still referred to as the fact that solid PCI 5 is really [PCI 4 ] + [PCI 6 r. The lesson to be
octahedral. As molecular symmetry is important learnt from all this is that coordination compounds are much more
for the arguments to be presented in many of common than one might at first think. The colour of many gemstones and
the following chapters, it will often prove
convenient to emphasize this by including in minerals, the chemistry carried out within an oil refinery, element deficiency
structural diagrams, lines which remind the diseases in animals, the reprocessing of nuclear fuel rods, the manufacture
reader of the molecular symmetry. Commonly, of integrated circuits, the chlorophyll in plants, the colours of a television
such lines will link ligands together and , clearly,
should not be interpreted as bonds between screen-all involve complexes, though we shall not be able to cover all
ligands. of these diverse topics in the present book. Although the first example we
gave in this chapter portrayed complexes as being formed between indepen-
dently stable species, and this is often the case, there are also many
fascinating examples of molecules which are only stable when they exist
as part of a complex; even independently stable species have their chemical
as well as their physical properties drastically changed as a result of
coordination.
In summary, there is no precise and time-constant definition of a
coordination compound- at one extreme methane could be regarded as
one-and the usage of the term is extended to all compounds to which
some of the concepts developed in the following chapters can usefully be
applied. Indeed, one could argue that the value of the concept lies in its
flexibility and adaptability so that the absence of a fixed and agreed
definition is no handicap. We shall find that the study of coordination
compounds excludes few elements-the sodium ion forms complexes- and
overlaps with biochemistry and organic chemistry. Further, it will involve
some fairly detailed theoretical interpretations, although in this book the
FJt. 1.3 An octahedron, a regular figure in
which all vertices are equivalent, as are all powerful but surprisingly simple concepts of symmetry are used to reduce
faces and all edges. theoretical complexities to a minimum.
6 1 Introduction
Table 2.1 Some classical ligands which are common in the complexes of either
transition metal andjor main group elements. The names given follow the rules
to be detailed in Chapter 3. Note that some species are shown twice, when they
can coordinate in more than one way. Note, too, that some ions shown once
can, in fact, coordinate in more than one way; examples are provided by eN-, S 2 o~
and OCW
Ligand Name
A list of some simple and common ligands is given in Table 2.1. The
entries in this table are confined to classical ligands, such as could well
have been studied by Werner. There are other ligands, many also simple
and common but non-classical-such as the organophosphines, which will
be covered shortly. Inevitably, the distinction we are making is an arbitrary
one. In Table 2.2 are listed representative examples of complexes formed
by some of the ligands in Table 2.1. The detailed molecular geometries of
the complexes in Table 2.2 will not be discussed because for many of them
there are ambiguities. These problems will be dealt with in Chapter 3,
where many of the examples given in Table 2.2 will reappear.
Complexes of most of the ligands that have so far been mentioned have
been studied for almost a century. Although one might expect the field to
be exhausted, each year there are a few new surprises: the discovery of a
method for the easy preparation of complexes of a metal in a valence state
Classical ligands, classical complexes 1 9
[
[Co(NH 3 ) 5 N3 ] 2~ ] 3+ the complexes it forms there simply is not enough space to fit very many
ligands around the central atom and so a low coordination number or
(NH3)4< >(NH3)4 unusual geometry results. It is found that metal ions in unusual coordination
geometries often have unusual reactions and/or properties and this makes
NH 2 them of particular interest. Thus, with suitable choice of ligand it is possible
[Cr04 ] 2 -
[Cr(SCN)6]3-
to make volatile compounds of sodium! Alternatively, by careful tuning of
[Mo(CN) 8] 3 - the ligand geometry it may be possible to make it highly specific for a
[Mo2S2(CN)8]6- particular metal. This produces visions of metal recovery from low-grade ore
[CuCI4] 2 - and even gold from sea water (such schemes tend to fail because the cost of
[CuCI5]3- the ligand and its recovery for reuse exceeds the value of the metal obtained).
[Fe(H2 0) 6 ] 3 +
[PtCI 2 (py)2 ]
Next, it may be possible to produce a ligand which closely mimics, in
[Pt(NH3),(0H),] its geometry and composition, that of one found in nature in a complex
[AI(OH)(H 2 0) 5 ]2+ of biological importance. The biological compound is almost certainly only
[TiCI 4 (Et,0) 2 ] available in small quantities, difficult to purify and unstable under most
[SnCisr laboratory conditions. Working with a model compound is much easier than
[AuF.r
[ZrF7 ]3- working with the real thing! This topic will be covered in much more detail
[BeF4]'- in Chapter 16.
[SbBr6 r There is one further advantage to working with ligands containing more
[Ailela] 2 - than one donor atom. This is that the (thermodynamic) stability of a
complex in which two or more donor atoms are part of the same ligand
molecule often appears much greater than if the same atoms were in
separate ligand molecules. There has been much debate on the origin of
this co-called chelate effect. Chemists like to use their imaginations and to
compare a metal ion held between two donor atoms on a ligand with a
crab holding its prey in its claws-hence chelate (Greek chelos-a claw).
It is common to talk of chelating ligands and of chelate complexes.
Complexes are often conveniently divided into two classes, labile and inert.
In the former, consisting of most complexes of main group metals and many
10 1 Typical ligands, typical complexes
of the more familiar transition metals, ligands are readily replaced. In the
latter, for example complexes of Cr111 and Co 111, ligand replacement is very
slow except under forcing conditions. The chelate effect is a phenomenon
which increases the inertness of complexes (which may mean making a
complex less labile). We shall return to the chelate and related effects in
Section 5.5. At this point all that will be added is a word of caution:
although the occurrence of a chelate effect is common it is not invariably
present. So, organic isocyanides, RNC, form many complexes (bonding to
the metal through the terminal C). The R-N-C sequence remains essentially
linear in complexes and the metal atom also bonds colinearly. The result
is that if a bidentate organic ligand containing two RNC groups is
synthesized then there has to be a sizeable number of carbon atoms (seven
CH 2 units, for instance) between the two -NC groups if they are both to
coordinate to the same metal atom and thus form a chelate. This means
a 12-membered ring system and, as will become evident in Chapter 5,
12-membered rings show no hint of a chelate effect.
In Table 2.3 are listed some of the more common, classical, polydentate
ligands. Again, imagination. The ligand is now pictured as biting, and thus
holding onto, the metal with several teeth (bidentate1 = two donor atoms,
tridentate= three donor atoms). The (minimum) distance between two
donor atoms in a bidentate ligand is sometimes referred to as the bite of
the ligand and the angle subtended at the metal atom the bite angle. In
Table 2.4 are detailed a selection of some of the more exotic ligand species
under current study. The systematic names of these molecules are usually
so horrendous that trivial, often physically descriptive, names are preferred.
Typical examples are picket fence, crown and tripod, some of which are given
in Table 2.4. Table 2.5 shows a selection of the complexes formed by the
ligands contained in the previous two tables.
Table 2.3 Some common polydentate ligands" {charges on anions are omitted)
0 0
Oxalato ox
0
,,,--,
~c-cIt·
I I
\
0
\ ';
CH2-CH2
I \
Ethylenediamine or 1,2-ethanediamine en NH2 NH2
\ I
©(As(CHal2
of>henylenenebis(dimethylarsine) or
1,2-phenylenebls(dimethyarsine)
diars
0 ~Hal2
Glycinato gly
8-Hydmxyquinolinato oxinate
w o-
1,10-Phenanthroline phen
©0© \ I
C~a /Ha
1/-c~
Dimethylglyoxlmato or 2,3-butanedione dioximato dmg O-N N-O
(see Table 2.5) /
I
\ I \
I
\ (continued)
12 1 Typical ligands, typical complexes
Tridentate• ligands
Diethylenetriamine dien
CH2-N...._
I H2
CH-N_...,.
I H2
1,2,3-Triaminopropane tap CH2-W.-
H2
CH:i'""""/\-C~
I CH2 CH2
CH2 I I
tren
I;~ c~ / -NH2
Tris(2-aminoethyl)amine NH2 NH2
CH:i'""""/\-C~
1 CH2 c=o
O=C I I
1,...1-c'- / -o
0 0 0
Nltrilotriacetato NTA
~L~
~~ ¢ ~~.
Tris(2-diphenylarsineophenyt)arsine QAS
'AsPh2
Novel ligands, novel complexes 1 13
Phthalocyanino pc
Hexadentate• ligand
o=c-o..- --o-c=o
I I
CH2 CH2
I I
Ethylenediaminetetraacetato EDTA --N-CH2-CH2-N--
I I
CH2 CH2
I I
o=c-o-- --o-c=o
a This notation is probably self-explanatory because of the examples given. If not, it is described towards the end of Section 3.1.
in some detail if not depth (that will come in the next few chapters). What
of the complexes formed by non-classicalligands and, in particular, what is
the ligand-metal bonding involved? The answer to this question has become
so important in inorganic chemistry that an outline answer will be given
here, although it will have to be refined later. The pattern is illustrated by
a discussion of the carbon monoxide molecule and the way that it bonds
to a transition metal; with not much modification a general picture emerges.
At the heart of it is the simultaneous-and interdependent- coexistence of
two distinct bonding mechanisms.
At first sight one might well expect CO to bond to a metal through its
oxygen atom-this oxygen has a lone pair of a electrons and oxygen donors
form many complexes; we would expect the oxygen to be more negatively
charged than the carbon because of the difference in their electronegativities.
Although oxygen-bonded CO complexes are known, the almost invariant
mode of bonding of CO to a transition metal is through the carbon atom.
14 1 Typical ligands, typical complexes
't
o=c\
NH
Oibenzo-l!k:rown-6
A crown ether with 18 atoms in the ring of which 6 are oxygens
A picket-fence porphyrin
(N'D
GI')Nyr;N
N
2,2 ,2-crypt
A cryptand (= Greek hidden); the 2s indicate the number of
liN)
oxygens In each N.. · N chain The sepulchrate ligand, which is hexadentate (the top
and bottom nitrogens do not normally coordinate)
tpp
Tetraphenylporphyrin
Bipyridyl groups grafted onto a cyclic hexamine
Novel ligands, novel complexes 1 15
A basket-handle porphyrin
When three bipyridine ligands are capped (twice) by a triamine
the cage ligand that results can complex three metal atoms
simultaneously (in this example the 'caps' are, essentially,
the ligand tren, tris(2-aminoethyl)amine)
CH3 /O-H···~
'c-:-N, -.r:!::-c-CH3
/,;>f 'N·---/ 'I
cH;;-c,~-?-=-:: '-.......N~c-cH3
I /
0 .... H-0
2+
(continued)
16 1 Typical ligands, typical complexes
3+
(Ntl
The con complex of tpp: one of two
equivalent structures (cannonical
GN'fj(N)
l\
forms) of the ligand is shown
Q) N
[Co(sepulchrate)] 3 +
There is also a lone pair of r:r electrons on the carbon which is larger, and
so more available for bonding, than that on the oxygen. This, then, is the
first member of the bonding partnership, a perfectly normal r:r donation from
carbon to the metal. But, given that a carbon atom is involved and that
carbon atoms seldom act as ligands in classical complexes unless there is a
negative charge around (as in CN-), it would be surprising if this r:r donation
above were to lead to a strong bond. It needs reinforcing. The CO molecule
is well able to provide this reinforcement because not only is the lone pair
r:r orbital larger on carbon than on oxygen, so too are the lobes of the orbitals
Novel ligands, novel complexes 1 17
(b)
M
which are the anti bonding counterparts of the C=O rr bonding orbitals. In
CO itself, of course, these orbitals are unoccupied but that is no reason for
ignoring them in complexes. They are LUMOs and a lesson which has
been well learnt in recent years is that LUMOs are seldom disinterested
spectators- they commonly play a key role in determining the outcome
of chemical reactions (and that means that they get involved in the bonding
somewhere along the way). It is therefore not surprising to learn that there
is both experimental and theoretical evidence that electron dona tion from
transition metal d orbitals to the CO rr antibonding orbitals is of vital
importance in the M--CO bonding. That is, the metal to carbon bonding
consists of two parts; 11 electron donation from the carbon to the transition
metal and rr electron back-donation from the transition metal to the carbon
(a) atom. This bonding is pictured in Fig. 2.1. Either bonding mechanism, on
its own, would lead to charge buildup on the metal (11) or the carbon (rr);
by acting together, the resultant charge buildup is small and each charge
transfer can proceed further than wo uld be possible in the absence of the
other. One talks of synergic bonding (synergismus is Greek for 'working
together').
Whenever one learns that a ligand bonds strongly to a transition metal
ion but much more weakly to a main group element, the involvement of
some rr-mediated back-bonding synergic mechanism is likely. Indeed, such
ligands are usually referred to as rr bonding ligands, not only because of
the existence of this mechanism but also because of its constancy, the
bonding always involving ligand empty rr orbitals of one sort or another.
In contrast, there can be more variability in the 11-bonding mecha nism.
Thus, for the latter the electrons involved can be those that are associated
with a chemical bond rather than with a particular atom. An excellent
example is provided by the ethene (ethylene) molecule, C 2 H 4 . Its mecha n-
ism of bonding (the usual pattern of 11 donation a nd rr acceptance) is shown
(b) in Fig. 2.2. The orientation of the bonded molecule is such that it is clear
tha t its orbital involved in 11 bonding is the C=C rr bonding orbital (which,
Fie. 2.2 (a) The u donation C2 H4 n ~ metal
empty orbital. (b) The rr back-donation of course, is occupied). The rr antibonding orbital is also that associated
metal ___. C2 H4 n antibonding orbital. with the C=C rr bond (a nd which, of course, is empty in the isolated ethene
18 1 Typical ligands, typical complexes
©
This is because many of the ligands listed can bond in several different ways,
each way giving rise to different complexes. In that a classic ligand such as
Cl- can bond not only to one but also to two metal ions (a bridging chloride
ligand) or to three which lie at the corners of a triangle (a face-bonding
chloride ligand) it is not surprising that a ligand such as CO should do the
Benzene same. How CO differs is that, much more than Cl-, it seems to form an
almost continuous range of intermediate bonding patterns between these
three main types. A ligand such as C 5 H 5 may have one, two, three, four or
five of its carbons bonded to a transition metal; when there are five they
may not all be equally bonded, and so on. Again, the range is enormous.
The cyclopentadienyl group CsH., The existence of such a range strongly suggests that the energy differences
usually written Cp.lt is a 5-electron between the various arrangements may be small. It seems that this, indeed,
donor (some prefer to regard it as is so. There are several possible consequences. First, the thermal vibrations
C5 H5 and as a 6 -electron donor)
- - - - - - - - - - - - - - - o f the molecule may contain sufficient energy to enable such a ligand to hop
Novel ligands, novel complexes 1 19
Table 2.7 Examples of complexes formed by some of the ligands in Table 2.6 and related ligands
©JI
The most famous complex of (1) A fulvalene complex (6)
"Q
cyclopentadiene, ferrocene
Fe
@
(C0) 2 Ru---Ru(CO),
A diphenylethyne
(7) Q~
w
(diphenylacetylene) complex
#
Dibenzenechromium; the corresponding (2)
cation [Cr(C6H6),j+ was made in 1919 but """c
C~'•
not recognized as such : ~l. . ~ _c--o
Cr ,,/co,
©
_....Co I C-....,
_....c I\ c o
' o c c I
I 1 o
~
0 0
d-
A flyover complex in which a
single hydrocarbon ligand
CH 2 Ph straddles two cobalt atoms
A complex of a cyclopentadiene derivative. (3) which are bonded to each
P(C6H5)3 is a ligand that plays a similar ,::::::- other. Each cobalt is also
H
bonded to an allyl group
role to CD __ I (these bonds are shown large
I dotted) and to a carbon atom.
Fe
CD/ ~~PPh3
A bisallyl palladium complex
(4)
'"~
G)= a
I 1,5.Cyclooctadiene (cod) (10)
~P"';r
Two of the compounds Fe( COb complex of platinum(Il)
LJ/ \r
formed by reaction of Fe(C0) 5
with alkynes
~ Cr
~
Some more exotic complexes
which, despite appearances,
are not significantly different
in kind from those above.
©
Cr
20 1 Typical ligands, typical complexes
(1) Fe 0 8 2 C5 H5 2x5 18
(2) Cr 0 6 2 C6 H6 2 X 6 18
}
(3) Fe 0 8 C5 H5 R 4
2 co 2 X 2 18
P(C6 H5 ) 3 2
(4) Fe 0 8 C5 H4 0
3 co
4
3x2 } 18
3x3 }
(5) Fe 0 8 C6 Me 4 0 4
18
3 co
(6) Consider each Ru separately: the Ru-Ru bond means that each Ru gains one electron
from the other for electron counting purposes:
(7)
Ru 0 8
c--eo
co LJ 18
(8)
Co 0 9 C2(CsHs)2
Co-Co
3 co LJ
Here, counting Co as Co 0 means that the allyl group is a three-electron donor. The
18
cobalt-carbon and cobalt-cobalt bonds each contribute one additional electron to each
Ll
cobalt:
Co 0 9 C3 H2RR'R"
Co-Co
18
Co-C
3 co
(9) Again, is it simplest if the pattern adopted in (8) is followed:
Pd 0 10 2 C3 H4 R 2 x 3 16
This is an example of a complex which does not follow the 18-electron rule; square
planar complexes of d8 ions usually deviate, giving 16 instead.
(10) In this final example it is difficult to avoid a formal charge on the metal because to
work with Pt0 would mean working with Br rather than Br-:
Pt 2+ 10 cod 2 X 2 } 16
2 Br- 2x2
Another d8 square planar complex that does not follow the 18-electron rule.
Some final comments 1 21
from one arrangement to another, perhaps equivalent to the first but perhaps
not, at quite a high frequency. Such molecular gymnastics of so-called
fluxional species have been the subject of extensive study, most notably by
NMR. This is a topic which will be covered in Section 14.8.
Second, in a chemical reaction it may happen that the most stable
orientation of a ligand in the starting material is not the most stable over
some parts of the reaction pathway, nor, perhaps, in the reaction product.
The third point stems from the second. When attached to a transition
metal, a ligand may be stable in a range of convoluted geometric orientations
that are not readily available to it as a free ligand. This may well mean that
some new, otherwise impossible, reaction pathways become available. Fur-
ther, their availability and nature are likely to be influenced by the other
ligands present, both for bonding and steric reasons. The consequence is that
the use of transition metal complexes in catalytic reactions has been an
enormous field of study, one in which it is of interest to study large numbers
of closely related compounds. With just the right, carefully tuned, catalytic
molecule one may hit the jackpot and become able to turn an otherwise
useless byproduct into a highly valuable compound! At a more realistic level,
organometallic compounds are the catalysts in several important industrial
processes. They do have a problem, though. If they involve an expensive
metal, platinum for instance, should catalyst recovery be difficult or incom-
plete, the cost of replacement catalyst may make uneconomic an otherwise
viable process. This is just one reason that heterogeneous catalysts-involving
surface complexes-are more important than homogeneous-involving dis-
solved complexes-unless the metal involved in the homogeneous catalysis
is cheap, the products volatile-and so easily separated-or the products
obtained of particular value. There will be more on this topic in Chapter 15.
Behind the content of this chapter, but not discussed within • 'Powerful new metal chelating agents developed' in Chern.
it, is the way that the information was obtained. Sometimes, Eng. News (August 1st 1988) 21.
there is a helpful story to be told. Two that may be mentioned • 'Molecules with Large Cavities in Supramolecular Chem-
are the discovery of the first complex containing coordinated istry' by C. See! and F. Viigtle, Angew. Chern., lilt. Ed. (1992)
N 2 : 'The Discovery of [Ru(NH 3 ) 5 N 2 ] 2 - ' by C. V. Senoff. 31, 528.
J. Chern. Educ. ( 1990) 67, 368 and, rather older. the nature
• 'Calixranes-supramolecular pursuits' by A. McKervey and
of the species giving rise to the intense blood-red colour
V. Bohmer, Chemistry in Britain (1992) 724.
obtained when thiocyanate, seN-, ions are added to iron(Ill)
solutions, 'The nature of iron(III) thiocyanate in solution' by • 'The Specification of Bonding Cavities in Macrocyclic Lig-
ands' by K. Hendrick and P. A. Tasker, Prog. Inorg. Chem.
S. Z. Lewin and R. S. Wagner, J. Chon. Educ. (1953) 30. 445.
Although both make reference to ideas developed later in (1985) 33, 1.
this book, this problem should not unduly inhibit under- • Supramolecular Chemistry, an Introduction by F. Viigtle, 1.
standing. Wiley, Chichester, 1991.
There are many sources of information on the exotic
A Nobel lecture by the major authority in the area, which
ligands discussed in the text. Samples which give easy-to-read
spans a wide range of topics from the history of the subject
insights are:
through to its applications in catalysis, photochemistry and
• 'Coordination Chemistry of Alkali and Alkaline-earth cat- biochemistry is 'Supramolec'.l!ar Chemistry-Scope and Per-
ions with Macrocyclic ligands' by B. Dietrich, J. Chern. spectives, Molecules, Supramoleculcs and Molecular Devices'
Educ. (1985) 62, 854. by J-M. Lehn, Angew. Chern., Int. Ed. (1988) 27, 89.
Questions • 18-crown-6
• 21-crown-7
2.1 Use Table 2.2 to compile a list of atoms (which may be
• dibenzo-12-crown-4.
part of a polyatomic ligand) which coordinate to metal ions
and relate the list to the region of the periodic table where 2.5 Cryptand ligands have been prepared in which up to
the atoms fall. If in doubt about whether an atom can half of the oxygens in 2,2,2-crypt (Table 2.4) have been
be a ligand, the answer is almost certainly 'yes', although even replaced by either S or NCH 3 . Suggest some ligands of this
so the final list will be far from comprehensive. type which might well have been studied.
2.2 Unlike Table 2.2, Table 2.3 contains no examples of 2.6 Select any three of the complexes shown in Table 2.7
ligands in which a halogen atom is the donor atom. Suggest and carry out a valence electron count on them. Compare
a reason why such species are rare. your results with those given in Table 2.8 (if any of the last
three complexes in Table 2.7 were chosen some thought will
2.3 Although Table 2.2 contains examples of ligands in which be needed in using Table 2.8 ).
nitrogen or arsenic are donor atoms there are no examples
which contain phosphorus. Suggest a few possible phosphorus- 2.7 As Table 2.6 hints, complexes are known in which 1,5-
containing ligands (if you wish, some of the ligands in Table cyclooctadiene (cod) bonds through both of its double bonds
2.2 might be appropriately modified). to a single metal atom. It may be that Fig. 2.2 can be
applied to each double bond separately. Equally, it could be
2.4 Using the ligands in Table 2.4 as a guide, suggest probable argued that although this approach may be valid for the rr
structures for: back-bonding, it is inadequate for the (J donation from the
• 15-crown-5 cod. Suggest a reason for this reservation.
Nomenclature, geometrical
structure and isomerism of
coordination compounds
3.1 Nomenclature
In order to facilitate communication between chemists it is desirable that a
convention for naming coordination compounds be followed. This section
contains an outline of the system suggested by a Nomenclature Committee
of the International Union of Pure and Applied Chemistry (IUPAC). 1
Although this convention is commonly adopted-the Russian literature
contains some variants and each language uses its own words, or spelling
(as in 'sulphate' and 'sulfate' for UK and USA, respectively, although the
UK has recently agreed to adopt the American version)-it is often simpler
to give a structural formula, e.g. [Co(NH 3 ) 4 CI(N02 )]+, than to write the
name in full and this will frequently be done in the following chapters. Apart
from this device for side-stepping the problem, in this book we generally
follow the most recent IUPAC recommendations except for those which
have yet to gain general acceptance.
The IUPAC system has both advantages and disadvantages. One major
advantage is that the most recent recommendations are just that, recom-
1 Perhaps the most generally available complete system is the American version because it
is contained in the Handbook of Chemistry and Physics published annually by the CRC
Press and which can be found in most libraries. Inevitably, there is a time lag before the most
recent recommendations appear so it is as well to check on this. The greatest detail and the
most current recommendations have been published in Nomenclature of Inorganic Chemistry,
Recommendations /990 (IUPAC) ed. G. J. Leigh, Blackwell, Oxford and it is these that have
been followed in the present text. At the same time as the IUPAC book appeared so, too, did
another, Inorganic Chemical Nomenclature by W. H. Powell and W. C. Fernelius, published by
the American Chemical Society (the publication is the outcome of the deliberations of this
Society's nomenclature committee), 1990. Whilst there are minor points of disagreement with
the IUPAC book, with its greater emphasis on the notations that will be found in the older
literature, the ACS publication may be seen as an acceptable complement to it.
Nomenclature 1 25
mendations. Previously, they had been 'rules', a situation that led one
group of writers to comment 'such rules often represent a compromise of
conflicting views and may not be completely acceptable (or convenient) to
all'. In what follows, in the spirit in which they are presented, the current
recommendations are not regarded as above criticism.
Having mentioned one advantage, let us get the disadvantages out of
the way. Names are often quite long, and involve numbers, brackets,
subscripts, superscripts and Greek letters. When reading a name all of these
details have to be remembered because it is not until the name is completed
that one can really start building up the complex. The name of the metal
is given at the end, whereas the thinking of most chemists starts with the
metal. It is not surprising that 'nickel tetracarbonyl' is in much more
common usage than the IUPAC name 'tetracarbonylnickel(O)'. So, whilst
the IUPAC notation is that generally encountered in the scientific litera-
ture, when chemists talk to each other they either greatly simplify it or use
one that has grown up in their specialist field. It is therefore only to be
expected that trivial names persist, so, although one should talk of the
hexacyanoferrate(II) and hexacyanoferrate(III) anions, most workers bow
to common usage and continue to speak of them as ferrocyanide and
ferricyanide, respectively. The practice of naming what has at some time
or other proved an important coordination compound after the person who
first prepared it is widely followed, thus: NH 4 [Cr(NH 3 h(NCS) 4 ], Rein-
ecke's salt; [Pt(NH 3 ) 4 ][PtC1 4 ], Magnus's green salt; [IrCOCl(P(C 6 H 5 h) 2 ],
Vaska's compound; [RhCl(P(C 6 H 5 hhJ, Wilkinson's compound and, one
met in Chapter 2 K[Pt(C 2 H 4 )Cl 3 ], Zeise's salt. A system which has
mercifully disappeared, but which will be found in the very old literature,
is that of naming a compound according to the colour of the corresponding
cobalt(III) complex (no matter the colour of the complex itself!). Thus,
'purpureo' salts meant ions with the general formula [M(NH 3 ) 5 Cl]"+.
The following rules summarize the more important recommendations
of the IUPAC committee and contained in their 1990 publication. It should
be recognized that when these differ from the previous rules the earlier
version is the one likely to be met by the reader~the new recommendations
are only just being accepted and used.
Although in writing the formula of a complex the central atom is given
first, in the corresponding name it is given last, thus [Fe(CN) 6 ] 3 -, hexa-
cyanoferrate(III). For anionic complexes the characteristic ending is '-ate'
(as hexacyanoferrate(III)), but for neutral or cationic complexes the name
of the central element (normally a metal) is not modified: [Fe(H 2 0) 6 ] 2 +,
hexaaquairon(II). A distinction is made for anionic complexes so that the
corresponding acids can be systematically named, the characteristic ending
for the acid being '-ic', as for H 4 [Fe(CN) 6 ], hexacyanoferric(II) acid. As
indicated in these examples, the formal oxidation state of the central atom
(Werner's primary valency) is indicated by a Roman numeral in paren-
theses after the name of the complex, but with no space between them. A
formal oxidation state of zero is indicated by (0) and a negative state by
a minus sign, e.g. (-I).
The name of the complex species is written as one word. Ligands which
may be regarded as carrying a negative charge (Cl-, SO;i- etc.) all end in
'-o' (chloro, sulfato etc.). Whereas, previously, the negatively charged
26 1 Nomenclature, geometrical structure and isomerism of coordination compounds
,u-amido-,u-nitrobis(tetraamminecobalt(lll))
When more than two metal atoms are spanned by a single ligand then
the number of centres spanned is indicated by a subscript: J1 3 , J1 4, Jls
(although the reader should be warned that some authors prefer to use
superscripts: J1 3 , 11\ etc.). In such cases, since each metal usually has its
own set of ligands, the formal name can stretch over several lines and, not
surprisingly, its use is avoided; a formula is given instead. So, the molecule
shown in Fig. 3.1 , in which P(C 6 H 11 h groups bridge four Rh(CO) units
which themselves lie at the corners of a square, the whole being capped
by a P(C 6 H 11 ) group, is written [Rh 4 (COMJ1 4 -P(C 6 H 1 d(J1-P(C 6 H 11 ) 2 ) 4 ],
although, written like this, the 1:1 relationship between the Rh and CO is
not explicitly stated. Note that J1S appear in sequence of decreasing suffix
values, although when a ligand is both bridging and non-bridging the
non-bridging is listed first (e.g. Cl - as both bridging and terminal ligand).
28 1 Nomenclature, geometrical structure and isomerism of coordination compounds
Not only can a ligand be attached to more than one metal atom, a given
ligand may be bonded through several atoms to a single metal. This is
described by the prefix '1 (eta), with a superscript to indicate the number
of atoms bonded. One talks of the hapticity (Greek: hapto to fasten) of the
ligand, thus: '7 1 monohapto, '7 2 dihapto, '7 5 pentahapto etc. This nomen-
clature is particularly important in organometallic chemistry, where, as
mentioned in Chapter 2, a given ligand may often coordinate in a variety
of ways and jump from one mode to another. So, the compound in Fig.
3.2 is tricarbonyl('1 4 -cyclobutadienyl)iron(O). Although its name would
lead one to adopt the representation shown in Fig. 3.2(b), for simplicity
that given in Fig. 3.2(a) is usually preferred, unless a specific point has to
be made (for instance, if, in a reaction, one of the four Fe-C bonds breaks).
A detailed discussion of the bonding in such compounds will be given in
Chapter 10. The above paragraphs have been written with the intention
of providing a comprehensive overview, not detail. The latter is contained
in Tables 3.1-3.3 which give specifics and examples.
Names such as those we have given above indicate the ligands attached
to a central metal atom but do not detail the positions of the ligands
relative to one another. If this information is important and has to be
included, an extension of the nomenclature is necessary. It often happens
that the prefixes cis or trans are adequate, as in the complex cis-dichloro-
FIJI. 3.2 Alternative symbolic representations of di(pyridine)platinum(II) in which the platinum and the four atoms co-
the bonding between C4 H4 and Fe(CO)a groups.
In (a) the delocalized nature of the " electron ordinated to it are coplanar:
system of the C4 H4 group is emphasized. In (b)
Cl py
the ~· aspect of the bonding is made evident.
"'pt/ (py = pyridine)
Cl/"' py
The most recent IUPAC recommendations on this point represent a
considerable departure from those previously made. We first outline the
new and then review the old, for as already mentioned, it is the latter which
Table 3.1 Anions and their names will be encountered in all but the most recent literature and texts.
when acting as ligands
Free anion Coordinated anion Table 3.2 Examples of the nomenclature of simple coordination compounds. Some
Amide (NH:2) amido (or azanido) of these examples contain, and adequately define, points not explicitly covered in the
Azide (Nil n~rido (azido will also text
be met)
Compound Nomenclature
Bromide <Bn bromo
Carbonate (Co~-) carbonate K,[ReF8 ] potassium octafluororhenate (note: only 'potassium')
Cyanate (CNO-) cyan ato [Cu(NH 3 ) 4 ]S04 tetraamminecobalt(ll) sulfate (note: 'aa' and 'mm')
Fluoride (F-) fluoro (not fluo) [CuCI2(py),] dichlorobispyridinecopper(ll) (note: bipyridine is the present name for the
Hydroxide (OW) hydroxo (or hydroxide 2,2'-bipyridine ligand-see Table 2.3. More strictly, and as in the text,
or hydroxy) di(pyridine) should be used to give dichlorodi(pyridine)copper(ll). How-
ever, in the spoken language an ambiguey can arise)
Nitrite (N0:2) nitro or nitrito-N
(see text) [Hg(C2Hsl2l diethylmercury(ll)
Oxide (02 -) oxo (or oxido) [Ni(PPh 3 ) 4 ] tetra(triphenylphosphine) nickei(O)
Thiocyanate (SCN)- thiocyanato-N [Ru(NH 3 ) 5 (N 2)] 2 + pentaamminedinitrogenruthenium(ll) (note: similarly, 0., is dioxygen, but
beware confusion w~h 0:2, superoxo and o~-, peroxo)
(N-bonded),
thiocyanato-S K,[FeC1 4 ] potassium tetrachloroferrate(ll)
(S-bonded) (NH 4 ),[SnCI6 ] ammonium hexachlorostannate(IV)
Nomenclature 1 29
di-p-chloro-bis[diammineplatinum(II)l chloride
p-amino-p-peroxo-bis[tetraamminecobalt(lll)l perchlorate
bis[carbonyl(u-carbonyl)'l5-cyclopentadienyliron(0)1
(one could add trans at the front as the cis form also exists)
,...,cH2 ,...c1 .- c1
CH' -Pd 'AI
''cH2 'c( 'c1
rfl-allylpalladium(II)di-p-chlorodichloroaluminium(III)
the two axes containing -N02, the rules require that we choose that axis
for which the priority number difference is a maximum. We now move to
the plane perpendicular to the axis already chosen and again select the ligand
of highest priority number. We again write down the priority number of the
ligand trans to it. So, the two isomers above become (OC-6-22) and
(OC-6-21), respectively, and they are distinguished. This completes the
process so that the isomer in Fig. 3.3(a) is called:
(OC-6-22)-triamminetrinitrocobalt(lll)
Optical isomers are not distinguished by the notation so far presented but
it can be extended to cover this requirement.
In the past, the system adopted has been to number the coordination
positions. The numbering system adopted for square planar complexes was
1
I
4~M-2
I
3
so that an alternative to cis-dichlorodi(pyridine)platinum(II) is 1,2-di-
chlorodi(pyridine)platinum(II). Notice the brackets around (pyridine). This
practice is recommended and in the present case serves to remove any
confusion with the ligand 2,2'-bipyridine (previously dipyridine). For octa-
hedral complexes the cis and trans nomenclature is often simplest, but
for complicated cases the numbering systems shown in Fig. 3.4 has been
adopted. The ligands which are at the corners of the front face of the
octahedron are cyclically numbered 1 --+ 3 and those on the back face
numbered 4 --+ 6. There are just two ways of arranging two sets of three
identical ligands, [ML 3 L3], in an octahedral complex. When identical
ligands lie at the corners of a face of the octahedron, one has what
traditionally has been called the facial-denoted fac-arrangement (Fig.
3.3(a)). When the three identical ligands lie on a plane that bisects the
octahedron, the meridional-denoted mer-arrangement results (Fig. 3.3(b)).
As has been recognized, two atoms of the same ligand may coordinate
to the same metal, leading to the formation of a ring structure. As
1 mentioned in Chapter 2, the formation of such rings by coordination is
termed chelation and the ligand is called a chelating ligand. Historically,
the language used to describe chelates depended on which side of the
Atlantic you lived. Traditionally, Americans talked of bidentate, tridentate,
tetradentate, pentadentate and hexadentate, the whole class being called
polydentate ligands (ligands attached by a single atom being monodentate).
British textbooks preferred bidentate, terdentate, quadridentate, quique-
dentate and sexadentate, calling the whole set multidentate ligands; single
atom attachment being unidentate. However, the latest IUPAC recom-
mendations are for didentate, tridentate, tetradentate and so on. Whether
didentate will replace the one name common to both American and British
notations-bidentate-remains to be seen. In this book the American
6 usage will be followed, this being the more commonly met. (This notation
Fig. 3.4 Ugand numbering system for an was used in Table 2.3.) A ligand which coordinates four atoms to each of
octahedral complex, indicating how the
numbering is related to the selection of two
two metals is called bisquadridentate, thus combining both notations, and
opposite faces of the octahedron. so on. Note that a polydentate ligand is not necessarily a chelating ligand
Coordination numbers 1 31
properties are discussed in some detail. In the present section are described
some of the geometrical arrangements of ligands which have been found for
various coordination numbers and examples of each are given.
First, however, we set the scene by noting that the frequency of occurrence
of coordination numbers for some ions of the first transition metal series is
roughly as follows:
chromium(lll) 6(oct) » 5, others very rare
iron( Ill) 6(oct) > 4(tet) > 5"' 7
cobalt(ll) 6(oct) > 4(tet) > 5 > 4(planar)
cobalt(lll) 6(oct) »> 5 > 4
nickel (II) 6(oct) > 4(planar) > 4(tet) "' 5
copper( II) 6(oct)t > 4(planar) > 5t "' 4(tet)t
t usually distorted
Although not exhaustive, this series illustrates the fact that there i& no fixed
coordination number for any ion. It cannot be emphasized too strongly
that the empirical formula of a compound often has little connection with
either the coordination number or geometry of any complex species it
describes-this recognition was Werner's breakthrough.
unfamiliar with this notation should at least read Appendix 3. which includes an outline of
it, at this point..
(b) 4 Two partial exceptions should be noted. First, transition metal complexes containing
Fig. 3.9 (a) Trigonal bipyramidal (03 ") and strongly n-bonding ligands tend to adopt the trigonal bipyramidal configuration. Secondly, it
(b) square pyramidal (C.,) modes of is possible to make approximate predictions for main group complexes, although a delicate
five-coordination. interplay of factors is involved.
36 1 Nomenclature, geometrical structure and isomerism of coordination compounds
r
'
M
~
Fig. 3.10 The synchronous ligand motion
which serves to tum (a) a trigonal bipyramidal
into a square pyramidal complex and (b) a
square pyramidal complex into a trigonal
bipyramidal one. If a complex undergoes the
displacement shown in (a) but does not
become locked into the square pyramidal
geometry (c) then, as the ligands continue their
original motion a trigonal bipyramidal geometry
is reattained (d). Comparison of the geometries
in (a) and (d) might lead an observer to think
that (a) has been rotated to give (d), although
of course it has not. For this reason the
sequence is called pseudorotation.
Pseudorotation serves to interchange axial and
equatorial ligands of a trigonal bipyramidal
rM I
(o)~
complex.
(d)
pentagonal bipyramid into a square pyramid and vice versa, the bottom
half shows what happens if the amplitude of the vibration shown in the
top left hand corner is so great that the atoms carry on beyond the square
pyramid arrangement to give another trigonal bipyramid. A casual observer,
knowing nothing of the square pyramidal intermediate and seeing only the
before (a) and after (d) arrangements, might well conclude that the original
molecule had simply been rotated to give the final one. Of course, no
rotation has occurred, it just looks as if one has. Hence the use of the word
pseudorotation. Other mechanisms, some rather ingenious, have been pro-
posed as involved in equational-axial interconversions in trigonal bipyram-
idal complexes but the Berry mechanism is believed to be the most
important. Indeed, it has become quite common in crystallographic work
in which the structure of a five-coordinate complex is reported, for the
authors to comment on the position on the Berry reaction pathway that
their particular complex occupies.
The small energy difference between the two modes of five-coordination
is demonstrated in the crystal structure of the compound
[Cr(enlJJ[Ni(CN) 5 ]1.5H 2 0, where there are two distinct types of
[Ni(CN) 5 ] 3 - anions, one square pyramidal and the other approximately
trigonal bipyramidal. Were one form to be appreciably more stable than
the other, then that would be the only one present in the crystal. Although
it was not evident from the crystal structure work, it seems that the 1.5H 2 0
play a key role presumably by hydrogen bonding to the anions; on
dehydration of the compound there is spectroscopic evidence that all
the anions become square pyramidal.
Examples of trigonal bipyramidal structures are the [Co(NCCH 3 ) 5 ] +
and [Cu(bpyhl] + cations. In the latter, one nitrogen of each bipyridyl is
in an axial position. Anionic examples are [CuC1 5 ] 3 -, [SnC1 5 r and
Coordination numbers 1 37
[Pt(SnC1 3 ) 5 ] 3 -. The latter, with Pt-Sn bonds, is formed when acidic tin(II)
chloride solution is added to many platinum salts.
Some main group halides have trigonal bipyramidal structures but care
is needed-structures may well differ from gas to solid (with solutions being
different again). PF5 and SbC1 5 retain the structure in gas and solution
but PC1 5 in the solid is better described as [PC1 4 ] + [PC1 6 r;
SbC1 5 remains
a trigonal bipyramid in the solid; NbC1 5 , TaC1 5 and MoCI 5 exist as dimers
in the solid state, the two chlorine bridges making each metal atom
six-coordinate.
Perhaps the best-known example of square pyramidal coordination is
the compound bisacetylacetonatovanadyl, [VO(acac) 2 ], where acac is acetyl-
acetone5 (CH 3-c0-CH 2-cO-cH 3 ) less one of the central (acidic) hydro-
gens, and in which the oxygen atom directly bound to the vanadium
occupies the unique position. 6 In one salt of the [Cu 2 Cl 6 ] 2 - anion, bridges
between adjacent anions lead to a square pyramidal configuration about
each copper atom (Fig. 3.11); compare this example with [CuCI 5 ] 3 -
mentioned above. Among the main group elements, the [SbC1 5 ] 2 - anion
provides an example of square pyramidal coordination.
A feature of square pyramidal structures is that there is the possibility
of an additional ligand occupying the vacant axial site to produce a
six-coordinate complex. Some of the small variations that have been
observed in the electronic spectrum of [VO(acac),] in different solvents
are believed to be caused by a solvent molecule being weakly bound at
the sixth coordination position. There is evidence that good donor solvents
sometimes also introduce a ligating atom cis to the vanadyl oxygen.
5 The correct name for this ligand is pentane-2,4-dionate, a name that
is being increasingly
used. However, in the literature and most texts the name acetylacetone is the one which will
be encountered and for this reason is the one used in this book.
6 Note the use of the ending yl in vanadyl.
The vanadyl cation is V0 2 +; the vanadium
and oxygen are strongly bonded. The fact that this bond remains intact through most
reactions and that we are dealing with a cation containing oxygen bonded to a metal is
indicated by the yl. Another example is uranyl, UO~ +. These two ions are sometimes quoted
as examples of one- and two-coordination, respectively (see Section 3.2).
38 1 Nomenclature, geometrical structure and isomerism of coordination compounds
Fig. 3.14 The pentagonal bipyramidal (1:5:1, Fig. 3.15 The one-face centred trigonal Fig. 3.16 The one-face centred octahedral
05h) mode of seven-coordination. prismatic (1:4:2, C2 ) mode of (1:3:3, C3 J mode of seven-coordination.
seven-coordination.
\
Cut here
(a)
f!C. 3.18 The dodecahedral (02 ") mode of of square anti prismatic coordination are the [TaF8 ] 3 - and [ReF8 ] 2 - anions.
eight-coordination (a dodecahedron is per11aps Typically, bidentate ligands with relatively short separations between the
best thought of as a solid figure with 12 faces,
which can be, but are not required to be,
two coordinating atoms, i.e. with a short bite, form dodecahedral complexes.
equilateral triangular). The arrangement is Examples are [Co(N03 ) 4 ] 2 - , in which two oxygens from each nitrate
shown in (c)-note the approximate pentagon coordinate to give four-membered rings, and [Cr(02 ) 4 ] 3 - in which both
subtended by this projection of the five 'outer'
ligands. The arrangement is best understood by
atoms of the peroxy o~ - anions coordinate to give three-membered
its construction from two pieces of card (a), rings.
interleaved as in (b). The diagram (c) is drawn Both the dodecahedron and square antiprism may be regarded as
such that one of the planes in (b) is
approximately in the plane of the paper.
distortions of a cubic arrangement of ligands (Fig. 3.19). They are favoured
because a cubic configuration would involve greater interligand steric
(b)
Coordination numbers 1 41
8 Because it is the fluoride ligand which is involved we confine our discussion to u bonding.
In the square antiprism (04 ,), the set of eight ligand a orbitals spans the irreducible
representations A 1 + B2 + £ 1 + E2 + E,-a set which matches the s, p and d orbitals on
the central metal: s(A 1 ); p,(B2 ); p., p,(£ 1); dx,, dx'-y'(E 2 ); dm d,,(£ 3 ). In the cube (0,) the
eight ligand a orbitals span A 1, + T2 , + A2 , + T 1,. Only if an f orbital is included is this
set spanned by the orbitals of the metal atom: s(A 1,); dxy• d,,, d"(T2 ,); fx,,(A 2 ,); Pxo p,, p,(T1,).
In neither set is the metal dz2 involved in the a bonding; in the cube, dxl-y2 is not involved
either.
42 1 Nomenclature, geometrical structure and isomerism of coordination compounds
• •
fiC. 3.22 One mode of 10-coordination--the fiC. 3.23 Eleven-coordination is very rare. One
bicapped square antiprism (0..,). The two possible mode of coordination is the all-face
capping ligands are those at the top and capped trigonal prism (03h). This differs from
bottom of the diagram. At the present time no Rg. 3.21 by the addition of ligands at the top
complex is known which contains ten and bottom of the figure.
monodentate ligands.
(with not too many electrons) main group elements, calculations on transition
metal compounds are less accurate.
In Appendix 2 is outlined the most popular and successful simple model
for predicting molecular geometry of main group compounds, the valence
shell electron pair repulsion (VSEPR) model. However, alongside it are
presented the results of some detailed calculations which prompt the
comment 'the VSEPR model usually makes correct predictions, but there
is no simple reason why'. The problem of the bonding in transition metal
complexes will be the subject of models presented in Chapters 6, 7 and I 0;
this last chapter reviews the current situation. At this point it is sufficient
to comment that the most useful applications of current simple theory are
those that start with the observed structure and work from there. In the
opinion of the author, the general answer to the question posed at the
head of this section is that we really do not know.
would correspond to a low lying excited state at the other, and vice
versa. Although it was at first found that the suggested example, in the
compound [Mo(O)CI 2(PMe 2Phh], was erroneous (one crystal structure
determination was on an impure crystal and gave misleading results), it
has stimulated great interest in the possible existence of this form of
isomerism. A subsequent reinvestigation has revealed two (pure!) crystal
forms of the compound in which the rather asymmetrical phosphine ligands
adopt rather different conformations. The Mo-O bond length is 1.663 A
in one isomer and 1.682 A in the other. There are different bond lengths
but the name distortional isomerism, the original one, perhaps is the more
appropriate if the phenomenon is regarded as a form of conformational
isomerism. However, there are some clearly established examples of bond
length differences in some dimeric ruthenium complexes such as ['1 5 -
Cp*RuCI(JJ-CI)]2 in which Cp* is the sterically demanding ligand C 5 (CH 3 )s;
apart from its steric effects, which are currently giving rise to considerable
study,9 it behaves like C 5 H 5 • In these, one isomer is diamagnetic, has no
unpaired electrons, and has a Ru-Ru separation of 2.9 A. The other isomer
is paramagnetic, it has unpaired electrons, and a Ru-Ru distance of 3.8 A.
However, because of the magnetic differences between the two isomers,
they are perhaps better regarded as spin isomers, a type which will be
described later.
"""OH /
Note that each of these two cations exists in a number of isomeric forms.
The reader may find it a useful exercise to draw pictures of all of the forms
and to enquire into the isomeric relationship between pairs.
the two isomers differing in the distribution of ligands between the cation
and anion:
and
The same metal may be the coordination centre in both cation and anion:
and
and
[Co(NH 3 ) 3 (N0,) 3 ] n= 1
[Co(NH 3 ) 6 ][Co(N0,) 6 ] n= 2
"' "'
CH 2 -CH-CH 3 and CH 2 -CH 2 -CH 2
NH 2/ NH 2 NH 2 / NH 2
propylenediamine trimethylenediamine
(pn) (tn)
are isomers, both of which form complexes of the type shown in Fig. 3.25
(where a convenient representation has been adopted for the two isomeric
ligands which shows only the coordinated atoms). In this situation, the
two isomers, indistinguishable by elemental analysis, are termed ligand
isomers.
A special form of ligand isomerism arises when two different sites in a
ligand can be protonated. If only one proton is added then two different
~- 3.25 The bidentate ligands are shown
species result, sometimes called protonation isomers. They are important
here very schematically. If in one complex they because both in the case that the proton is replaced by a metal and in the
represent a particular ligand but in another case that the unprotonated site coordinates, different complexes result.
complex an isomeric ligand, then the two Such differences are important in some biochemical systems.
complexes provide an example of ligand
isomerism. A special case of ligand isomerism also arises when the ligands are
optical isomers-enantiomorphs- of each other. One interesting problem
is the extent to which electron absorption bands, which, as a first approxima-
tion, are supposed to be localized on a transition metal ion, acquire optical
activity because of the activity of a coordinated ligand. An example of this
is provided by the ligand mentioned above, propylenediamine (pn) which
exists in optically isomeric forms.
en~ 3+
"'
CH 3 S
/
N-C/ (.
"'"./
""'
CH 3 S
Here, one has to be careful because spin isomerism behaviour in solution
may well be different to that in the solid state. In the solid state the individual
magnetic ions couple weakly together but often sufficiently strongly for the
phenomenon to have a cooperative aspect and to show hysteresis; the
stronger the cooperativity the more abrupt the transition. The spin-crossover
may be induced not only thermally but also by application of pressure, for
small structural changes accompany the spin change.
For both of the last two classes of isomerism we have detailed-fluxional
and spin isomerism-the lifetime of individual isomers may be rather short.
Spin isomers, for instance, typically live for about 10- 7 s (but see Section
3.4.1 and the Further Reading at the end of this chapter). Some would
argue that classical isomerism refers only to species capable of physical
separation, and so of long lifetime. However, with the increasing use of
methods which explore short lifetimes-NMR and EPR, particularly, in
the present context-it seems sensible to ignore this limitation.
Section 9.12. Sometimes it is possible to photoexcite a molecule Hendrickson Angew. Chern., Int. Ed. (1994) 33, 425 (discusses
to the less stable spin state, whereupon it may be stable for the isomerization in the context of spin-crossovers and 'show
weeks! An example, which shows how the methods introduced that it is a reality'). A paper that looks back at the origins of
in later chapters of this book may be used to investigate such the controversy, and comes up with the suprising answer
systems, is in a paper by Gutlich and Poganiuch in Angew. contained in its title, is 'Studies of Distortional Isomers. 2.
Chern., Int. Ed. (1991) 30, 975. Evidence That Green [LWOC1 2 ]PF6 is a Ternary Mixture'
A modern general review on isomerism is to be found in by P. J. Desrochers, K. W. Nebesny, M. J. LaBarre, M. A.
an article by J. C. Bailar in 'Coord. Chern. Rev.' (1990) 100, I. Bruck, G. F. Neilson, R. P. Sperline, J. H. Enemark, G. Backes
For those interested in learning more of the bond-stretch and K. Wieghardt Inorg. Chern. (1994) 33, 15; see also the
isomer story there is detailed review by V. C. Gibson and M. footnote on page 2.
McPartlin in J. Chern. Soc., Dalton Trans. (1992) 947 and a A useful source is Volume I of Comprehensive Coordination
shorter one by J. M. Mayer, Angew. Chern., Int. Ed. (1992) 31, Chemistry G. Wilkinson, R. D. Gillard and J. A. McCleverty
286. Both became slightly out of date because of another (eds.), Pergamon Press, Oxford, 1987, and in particular Chap-
article: A. P. Bashall, S. W. A. Bligh, A. J. Edwards, V. C. ters 2 ('Coordination Numbers and Geometries' by D. L. Kepert),
Gibson, M. McPartlin and 0. B. Robinson Angew. Chern., Int. 3 ('Nomenclature of Coordination Compounds' by T. E. Sloan)
Ed. (1992) 31, 1607. Even more recent is 'Bond Stretch Isomers: and 5 ('Isomerism in Coordination Chemistry' by J. MacB.
-
Fact not Fiction' by P. Giitlich, H.A. Goodwin and D. N. Harrowfield and S. B. Wild).
4.1 Introduction
This chapter reviews the most common methods by which coordination
compounds are prepared. However, current research is almost invariably
aimed at producing the unusual and exotic, not the common. So, a
contemporary research journal would describe methods rather less simple
than most of those covered here. A flavour of the current has therefore been
included, although the reader is unlikely to meet some of the compounds
outside the research laboratory. In the reactions described in this chapter,
there are two important variables-coordination number and oxidation
number (the latter is often called the valence state). In principle, either may
increase, decrease or remain unchanged in a reaction, and the reader may
find it helpful to classify the preparative methods described according to
changes in these two numbers. In practice it is not always possible to be
certain of either without more information than that contained within a
chemical equation or chemical formula. A ligand which is potentially triden-
tate may, for example, act as a bidentate ligand and so the coordination number
differs from that expected. Similarly, is the complex ion [Co(NH 3 )sN0]2+
a complex of cobalt(II) or one of cobalt(III)? It depends on whether you
believe that the NO is better represented as NO" (where, in the complex, the
odd electron is paired with a cobalt electron) or as NO- . This problem of
formal valence states will reappear later in this chapter and again in
Chapter 6.
Complications apart, reactions in which the coordination number of an
electron acceptor is increased are called addition reactions, and when it is
unchanged they are called substitution reactions. The coordination number
decreases for dissociation reactions. Reactions involving valence state changes
are called oxidation or reduction reactions, as appropriate.
An important classification of complexes depends on the speed with which
52 1 Preparation of coordination compounds
When one reactant is a liquid and the other a gas at room temperature, a
different technique is usually followed. In the preparation of [BF3 (0Et 2 )],
for instance, the diethylether and boron trifluoride, stored in separate bulbs
Preparative methods 1 53
Reactions between liquids or solids are best carried out by mixing solutions
of them in a readily removable inert solvent, e.g.
40-60 ·c bp
SnCI 4 + 2 NMe3 --pe-tr-ol-eu_m_e-th_e_r-+
Solids may dissolve in complexing agents with a change in valence state (so
it is debatable whether such reactions should be classified as simple addition).
In the case of the dissolution of metallic silver or gold in water in the
presence of cyanide ion, the oxygen of the air acts as the oxidizing agent:
2M(s) + 4CW(aq) + ~0.., + H2 0 .... [M(CN 2 lr + 20W (M = Ag or Au)
from pink aqueous solutions containing octahedral Co 11 and CsCI. The third
point, that some complex ions display incongruent solubility, as will be seen,
is related to the second.
If an aqueous solution containing iron( II) sulfate and ammonium sulfate
in a 1:1 molar ratio is allowed to crystallize, then a compound which
historically is variously known as Mohr's salt and as ferrous ammon-
ium sulfate, [Fe(H 2 0) 6 ]S04 (NH 4 lzS04 , is obtained. Mohr's salt is said to
show congruent solubility. On the other hand, if an aqueous solution
containing a 2:1 molar ratio of potassium chloride and copper(II) chloride
crystallizes, crystals of potassium chloride are obtained first. Only later does
the complex K 2 [Cu(H 2 0lzCl 4 ] crystallize. Similarly, attempts to recrystal-
lize the salt will lead to the initial deposition of potassium chloride. The
complex is said to display incongruent solubility; it can only be obtained
from aqueous solutions containing excess of copper(II) chloride. A system
which displays incongruent solubility at one temperature may display
congruent solubility at another.
Examples of the formation of complex ions by substitution reactions
of labile complexes are the following.
1. The action of excess of ammonia on aqueous solutions of copper( II) salts:
[Cu(H 2 0) 4 ] 2+ + 4NH 3 (aq)--+ [Cu(NH 3 ) 4 ] 2+ + 4H 2 0
Although this equation 1
shows the complete substitution of coordinated water
by ammonia all such reactions occur in steps and the species [Cu(H 2 0) 4 ]2+,
[Cu(H 2 0JJNH 3 ]2+, [Cu(H 2 0h(NH 3 lz]3+, [Cu(H 2 0)(NH 3 JJ]2+, and
[Cu(NH 3 ) 4 ] 2 + are all present in the solution, although the concentrations of
some are low. By a suitable choice of concentration (using stability-constant
data of the sort discussed in Chapter 5) it is possible to ensure that the
concentration of one particular component, [Cu(H 2 0) 2 (NH 3 ) 2 ] 2 + say, is a
maximum in the solution. However, it does not follow that if crystallization is
induced (for example, by adding ethanol to the solution and so decreasing
the solubility of the complex species) the complex which crystallizes will
contain the [Cu(H 2 0lz(NH 3 lz] 2 + cation. There are many labile complexes
which may be studied readily in solution but which are very difficult to
obtain in the solid state. The converse is also true. Copper(l) bromide reacts
in ethanol with Br- to give solutions in which only the [CuBr 2 anion has r
been identified. From such solutions, crystals of salts containing anions such
as [Cu 2 Br 5 ]3+ and [Cu 4 Br 6 ] 2 -, as well as [CuBr 2 r,
have been obtained.
When a salt such as [N(CH 3 ) 4 ] 3 [Cu 2 Br 5 ] is dissolved in nitromethane, the
dominant species in solution is again [CuBr 2 r.
2. The reaction between aqueous solutions of thiourea and lead nitrate:
[Pb(H 2 0) 6 ] 2+ + 6SC(NH 2 ) 2 --+ [Pb(SC(NH 2 ) 2 ) 6 ] 2+ + 6H 2 0
The lead(II) ion in aqueous solutions exchanges water between its coordina-
tion sphere and the bulk very rapidly but is probably best regarded as six-
coordinate (although some evidence indicates that the coordination number
1 A common coordination geometry for the copper(II) ion is to be surrounded by four ligands
in a plane which, together with two ligands one above and one below this plane but further
from the copper atom, form a tetragonally distorted octahedron. In this discussion these two,
more weakly bonded, ligands have been neglected.
56 1 Preparation of coordination compounds
The contrast between these reactions and those of labile complexes, where
reaction is complete almost as soon as the reactants are mixed, is very
evident.
Preparative methods I 57
2. As has been seen, refluxing thionyl chloride reacts with water and may
be used to prepare anhydrous metal chlorides from the hydrates. Addition-
ally, it is a suitable solvent for the preparation of the chloro anions of
metals:
4. As will be discussed in more detail in Chapter 15, there are three distinct
bonding mechanisms which contribute to the metal-metal bonding in
[CI 4 Re--ReCI4 ] 2 -, u, n and <5 (the chlorines are all terminal and eclipsed),
and so it is an anion of particular interest. It is readily prepared from the
commercially available ReCI 3 , which consists of molecules containing a
triangle of rhenium atoms, by fusion in molten (220 oq diethylammonium
chloride:
The success of this general preparative method rests on two factors. First,
although the product is an inert complex, the starting material is one which is
relatively labile. Other things being equal, concentrations used in the
preparation approximate to those which maximize the concentration of a
complex species identical in composition with the desired product but
differing from it in charge. Electron addition or removal (i.e. reduction or
oxidation) then gives the product. Secondly, as has been mentioned earlier,
there will be several labile complexes in equilibria, each of which can undergo
oxidation (or reduction) to give an inert product. In general, the product
actually obtained will be derived from that labile complex which is the most
readily oxidized (or reduced).
Many other complexes behave similarly and heating (usually under vacuum)
to a carefully controlled temperature is a useful preparative method.
Hydrogen halide elimination, for example, is a reaction which occurs readily
for almost any complex which has the electron-donor atom attached to a
hydrogen (e.g. H 2 0, ROH, NH 3 , R 2 NH) and the electron acceptor attached
to a halogen (e.g. BF3 , SnBr4 , FeCI 3 ). So, in the preparation of a complex
such as the first given in this chapter, [BF3 (NH 3 )], the product readily
eliminates hydrogen fluoride to give a series of compounds and ultimately
the polymeric solid BN. Another very common thermal reaction is the
expulsion of one or more neutral ligands (as in the case of copper(II)
sulfate), with a consequent reduction in the apparent coordination number
of the central atom. In fact, quite often, some previously monodentate ligand
(generally an anion) becomes either bidentate or a group that bridges two
metal atoms (the acetate and halide anions, respectively, exemplify these two
cases). Another possibility is that an anion, initially non-coordinated,
becomes attached to the metal. The actual coordination number of the
central atom is seldom reduced. Examples have already been given in Section
3.4.6.
Heating to a relatively high temperature can lead to the complete
dissociation of the complex species. For example
K[BF4 ] -+ BF3 + KF
and
BrF2 [RuF6 ] -+ BrF3 + RuF5
In the absence of some other suitable cationic species, complex fluorides
62 1 Preparation of coordination compounds
containing the [BrF2 ] + cation are formed when transition metals are
dissolved in BrF3 • Thermal decomposition of these salts is a convenient way
of making small quantities of many fluorides.
Another example of a complex prepared by a thermal dissociation
reaction is the preparation of cis-[Cr(en)zC1 2 ]+ by heating [Cr(en),]Cl 3 :
210'C
[Cr(en) 3 ]CI 3 - cis-[Cr(en) 2 CI 2 ]CI +en
possible and common, but the volume of a typical box is such that it is rather
difficult to reduce and maintain the oxygen concentration at an acceptable
level. There exist sophisticated boxes with recirculation of the inert gas
through oxygen-removing trains and entrance ports which can be thoroughly
evacuated, but they are very expensive. At the other extreme, it has been
pointed out that it is possible to work with a cheap transparent plastic glove,
in which the fingers are used to store and mix reagents, nitrogen being passed
in through the sealed-off wrist.
Many workers make use of so-called Schlenk tube techniques. This is the
name given to a whole series of simple, but versatile, devices that enable
reactions to be carried out in an essentially closed apparatus of low volume.
Examples include the use of septum caps (those used to close the vials
containing materials used for medical injections). Solvents can be taken in
and out of the apparatus using hypodermic needles; nitrogen can be passed
in through such a needle and allowed to escape through another; on removal
of all needles the apparatus is automatically sealed. Filter sticks are
used-glass tubes with a glass sinter sealed halfway down. If one is on top
of the solution to be filtered, inversion of the apparatus (perhaps with gentle
use of nitrogen gas pressure) leads to filtration. Solids can be placed in a
limb of a tube which has a bend of ca. 90° in its middle and which is held
with the angle at the top. When the solid is needed for reaction then, if the
solid is in the left-hand limb, rotation by 180° about the right-hand limb
causes the contents of the two limbs to mix. Tubes are interconnected by
taps (often greaseless) so that alternative routes exist for gas and/or liquid
flow and add to the versatility.
An example of a series of reactions involving Schlenk tubes is the
preparation of (CH 3 )NGaH 3 . GaC1 3 is dissolved in diethylether and slowly
added to a slurry of LiH in diethylether through a greaseless valve. After
reaction, the product is filtered through a filter stick. To the filtrate
(Li[GaH 4 ] in diethylether) is added [(CH 3 )eNH]Cl by the rotating arm
technique. After vacuum removal of the solvent, the product is separated by
vacuum sublimation from the reaction mixture.
Reactions involving metals, either bulk or, more commonly, finely divided,
are an entry point for organometallic complexes of transition metals.
Examples of direct reaction of a metal are
and
Fe + 5CO --+ Fe(C0) 5
(it is found to be important to use a 3:1 reactant ratio; the yield is only ca.
50%). Closely related are reactions in which a hydrocarbon, such as
cyclopentadiene, C 5 H 6 , which has an acidic hydrogen (loss of a proton gives
the C 5 H5 anion, with an aromatic n system), reacts either with a metal or
strong alkali to give the anion. This is then commonly reacted with a metal
halide or metal halide complex:
THF
2C5 H6 + 2K - 2KC5 H5 + H2
4KC5 C5 + UC1 4 --+ U(C5 H5 ) 4 + 4KCI
Sometimes, the steps are contained within one reaction mixture, as in the
preparation of ferrocene:
CH3 CH3
I I
0-C 0-C
Fig. 4.4 Bromination at the ligand in
trisacetylacetonatochromium(lll).
cr ~Q~c-H
0-C
10'
Cr \
0-C1
C-Br
I I
CH3 CH3
3 3
are given, firstly, by the reaction of the cobalticinium cation with (boro )·
hydride,
[(~ 5 -C 5 H 5 ) 2 Co]+ + W-+ [(~ 5 -C 5 H 5 )Co(~ 4 -C 5 H 6 )]
~ ;;;; -CH2-CH2-
way to an extensive study of the chemistry of one group of coordinated
ligands. Syntheses of the type just described are often called template
reactions for an obvious reason-the metal together with the ligands already
in place form a template for the ligand to be created.
A rather important group of complexes which may also be made by
template synthesis are those of imines. lmines are formed by condensation
of an amine and a carbonyl (Fig. 4.7). The amine can be coordinated to a
metal and the above reaction still proceeds, the amine (or, more usually,
diamine) remaining coordinated. Typical is the reaction between bisethylene-
diaminenickel(II) chloride and acetylacetone. An aqueous solution of the
mixture, to which a few drops of pyridine have been added, is refluxed
R
Fig. 4. 7 Condensation of an amine and a \
carbonyl to form an imine. C=NR"
I
R'
68 1 Preparation of coordination compounds
Me
Py I
C=O
NH2
HrnN I
Ni +2
I
CH2 + 2py +en
H21~ ! NH2
I
C=O
Py I
Me
FI,C. 4.8 Reaction between for 2 h to give the product. The reaction is probably between [Ni(enlz(pylz]2+
bisethylenediaminenickel(ll) chloride and and the acetylacetone (Fig. 4.8). Template syntheses are finding particular
acetylacetone.
applicability to the synthesis of large, sometimes complicated, ligands. The
improvement in yield of ligand, compared with an organic synthesis in the
absence of metal, is sometimes quite spectacular.
Many reactions are known in which the presence of a metal ion influences
the products of a reaction and the explanation for this may well lie in the
different reactions of free and coordinated ligands. Peptide chemistry is one
field in which this may prove to be of great importance.
One can see from this list the difficulty of proposing a general explanation
Preparative methods 1 69
CI--NH31- CI--NH31-
I Pt \ I Pt \
Cl Cl Cl N02
cis
~ NH3
CI--CI 12-
Flil. 4.9 The sequence of reagent addition can I Pt \
be altered to selectively provide cis and trans Cl Cl
products from the same starting complex.
1 N02
for the trans effect. Almost all those ligands exerting a strong trans effect are
n bonding, but a n bonding explanation does not explain the position of H-
(a polarization model, based on the unique characteristics of the H atom is
usually added to cover this). However, there seems no explanation of why
the effect is largely confined to Pt11 , perhaps along with Pd 11 and Aum-
certainly, neither the n bonding nor the polarization model is metal-specific.
It could be that the notion that the trans effect is largely confined to Pt 11 is
incorrect; it is just that other elements have not been so extensively studied.
Although this point has substance as far as Aum and Pd" are concerned, it
does not seem to have general validity. Above all, it must be remembered
that the trans effect is a kinetic effect, associated with bond breaking and
formation. It could be more a phenomenon of the reaction pathway (an
activated complex pathway, an activated complex or transition state) than
the ground state. In Chapter 14 the kinetics of the reactions of square planar
Pt11 complexes will be the subject of some discussion, a discussion that will
include the trans effect.
vacuum chamber that holds the cooled surface, typically cooled with liquid
nitrogen. Many heating techniques-resistive heating, induction heating,
laser ablation are a few-may be used to generate the metal atoms. Along
with the metal, but perhaps in alternate pulses, are condensed the chosen
ligand(s). By such techniques tin carbonyls, compounds unstable at tempera-
tures well below room temperature, have been prepared and characterized.
Paradoxically enough, such metal atom synthesis methods can also provide
a convenient high yield route to compounds which are stable at room
temperature; commercial equipment is available for those wishing to work
on a large scale. A simple version is also available for use in student
laboratories. It is methods such as these which are being used to prepare
fullerene, C 60 , and related species. This soccer-ball-like molecule can encap-
sulate some ions-Sr2+ and LaH are two examples-if suitable sources for
them are present in the carbon rods from which the fullerene is prepared
(they are evaporated by an electric arc struck between them to give a low
yield of fullerenes). The bonding is these so-called endohedral molecules is
discussed in Chapter 10.
Photochemical methods, usually irradiation with ultraviolet light from a
mercury discharge lamp, but sometimes visible light too, have long been used
in the preparation of coordination compounds. They depend on the fact that
the chemistry of an electronically excited molecule is different from that of
the same molecule in its electronic ground state. However, this can only
be exploited synthetically if the excited state lives long enough for chemistry
to be performed on it-and most excited states have very short lifetimes
because excited molecules give up their extra energy to other molecules in
a wide variety of processes. It is likely, therefore, that the success of
photochemical methods depends, in part, upon there being an insulated step
down the ladder of energy changes that lead to deactivation; a level of long
lifetime, from which deactivation is slow. Such levels exist-lasers depend
on them for their action-but it is difficult to predict them and so to predict
whether photochemical methods will lead to new compounds or just to the
destruction of those already present. One compound which is always made
photochemically is the carbonyl Fe 2 (C0) 9 , by the reaction
h,.
2Fe(C0) 5 - Fe2 (C0) 9 + CO
In recent years it has become clear that there exist complexes which
contain the H 2 molecule as a ligand; not two separated H atoms, but H 2 •
One method by which such complexes might be prepared is by the action
of H 2 on suitable coordination compounds. Unfortunately, there is a
problem-the solubility of H 2 in most solvents is rather low, so high
pressures of H 2 are needed. Not surprisingly, most workers would prefer to
work with low pressures of H 2 than with high. It is here that the ingenuity
of experimentalists becomes apparent. Supercritical fluids have properties
which in some ways resemble those of liquids. For example, they can act as
solvents for coordination compounds, and in some ways they resemble gases
in that they are miscible with gases, usually over the entire concentration
range. Here, then, is a solution to the problem-study the reaction of
coordination compounds with H 2 using supercritical fluids as solvents. The
pressures needed to maintain suprecriticality are often modest-a few tens of
Further reading 1 71
atmospheres for carbon dioxide or xenon, for instance. At the time of writing,
this is an area in its infancy but it could lead to important new developments
in the preparation of coordination compounds. 2
Finally, the solid state. The preparation of solid state compounds by high
temperature synthesis is a long established method. Unfortunately, the
available techniques have been rather limited-grind the reactants together
to a fine powder, heat, regrind, reheat and so on until uniformity is
reached-is a typical procedure. But interest in the solid state is growing
rapidly. For instance, some simple inorganic materials show long-range
structural correlations which are difficult to understand or reproduce (for
instance, despite the simple picture presented in most introductory inorganic
textbooks, ZnS has been found to crystallize in several hundred different,
but related, crystal structures, although there is no known method of
controlling which form is produced). Again, the discovery of ceramic
high-temperature superconductors, (a topic which is dealt with in Appendix
10) incorporating Cum, a rather unusual valence state, has sparked off a
search for similar novel properties of solid state materials containing metal
ions in unusual valence states.
There have been developments in synthetic methods. First, related to the
'heat and grind' method, gel/colloid methods of producing reaction pre-
cursors are giving much more control and more reproducibility of the final
product. Secondly, hydrothermal methods are finding utility. In these,
reactions are carried out at high temperature and pressure conditions
in the presence of a solvent (not necessarily water, although this has been
most commonly used). For instance, Fe"Fe~'F8 • H 2 0 was prepared in this
way using liquid HF as a solvent.
2 SeeM. Po!iakoff, S.M. Howdle and S. G. Kazarian, Angew. Chern., Int. Ed. (1995), 34, 1275.
5.1 Introduction
The statement that a compound is stable is rather loose, for several different
interpretations may be placed upon it. Used without qualification it means
that the compound exists and, under suitable conditions, may be stored for
a long period of time. However, a statement such as ' a compound is stable
in water' may mean one of two things, either that there is no reaction with
water which would lead to a lower free energy of the system (thermodynamic
stability) or that, although a reaction would lead to a more stable system,
there is no available mechanism by which the reaction can occur (kinetic
stability). For example, there may not be enough energy available to break
a strong bond, although once broken it could be replaced by an even stronger
one. As we have seen, boron trifluoride forms a stable complex with
trimethylamine, [BF3 (N(CH 3 h)]. A similar complex is formed with trisilyl-
amine, [BFiN(SiH 3 lJ)], which is thermodynamically unstable with respect
to the reaction
M + L.,tML
the equilibrium constant K 1 for the complex containing a single ligand will
be
[ML]
K1 = [M][L]
where, for the moment, activity coefficients of unity have been assumed. If
ML adds a further molecule of L,
ML + L+t ML,
then the equilibrium constant for the complex containing two ligand
molecules is
In general, the equilibrium constant for the formation of the complex ML.
from ML._ 1 will be
Kn = [Ml.,]
[Mt.,_ 1 )[L]
as
[Ml.,]
Pn = [M][L]"
Pn = K1 X K, X · · • X Kn
That is
1 Throughout this chapter we shall often not specify the charges on the species in reactions
or equilibria. Square brackets are used to indicate both the concentrations of complex species
and the species themselves. It will be clear from the context which is intended.
Determination of stability constants 1 75
Table 5.1 Typical stability constant data for monodentate ligands. All values are logarithmic so, for the Sn 2 + fCI- system,
logK1 = 1.51
These data refer to 25 'C unless otherwise stated and to zero ionic strength. As a comparison, the data for Ni 2 • /NH 3 (30 'C) in 2M NH4 NO:, are:
K 1 = 2. 78 K, = 2.27 K3 = 1.65 K4 = 1.31 K5 = 0.65 K6 = 0.08
Notice, particularly, the effect of the change on Ks· The entries included in this table have been chosen to illustrate the variety of formats that are
encountered and yet to be internally self-explanatory. As an example, the statement that K1 = 1.51 for the Sn 2 + ;cl- system is to be interpreted as
For the more general case where more than one formation constant is to
be determined, the problem is usually more difficult. For inert complexes it
may be possible to separate, and separately obtain the concentrations of,
the various complex species. In this way Bjerrum was able to determine the
six stability constants within the [Cr(H 2 0) 6 ]3+, [Cr(H 2 0) 5 SCN] 2 +, ... ,
[Cr(SCN) 6 ] 3 - series. However, this is a potentially unreliable method and
has been little used. Some methods of tackling the general problem will now
be indicated. The variants are many for this is a field in which considerable
ingenuity has been used in the design of experiments and in the analysis of
experimental data. As an example consider a ligand for which the species
H 2 L, HL- and L 2 - all exist. The metal ion M and protons may be regarded
as being in competition for L 2 -. If we titrate H 2 L against standard NaOH
solution we obtain a pH-volume curve of the form shown dotted in the
upper part of Fig. 5.1. Now add a known amount of M. Some H+ will be
displaced. If the mixture is again titrated with NaOH the curve shown dotted
in the lower part of Figure 5.1 is obtained. Compare the titres [NaOH] 1
and [NaOH] 2 at the same pH (shown dashed in Fig. 5.1). Now, the H+
liberated by the added M is dependent on the amount of ligand bound to
M, ligand which was previously protonated, the average number of H+ ions
per ligand being nH (nH = 2 for H 2L). We have
[W ion liberated] = [L bound to M] x nH = [NaOH] 2 - [NaOH] ~
So,
[L bound toM]= [NaOH],- [NaOH]~
nH
NaOH--
Determination of stability constants I 77
where the complex species are ML, ML 2 , ..• , MLN and [M], is the total
concentration of metal ion, complexed and uncomplexed, in the solution.
Now,
N N
[M], = I [ML,] = [M] + I [ML,]
n=O n=1
and
[Ml.,] = fln[M][L]"
[M],
[M]
= 1 + I
n=O
/Jn[L]"
78 1 Stability of coordination compounds
If both [M] and [L] can be measured, keeping [M], constant, but changing
the total ligand concentration for each measurement, this relationship leads
to a set of simultaneous equation in the f3s (one equation for each
measurement) which provide a quick way of determining f3 values. Alterna-
tively, if [L] is made large, so that [L] » [M], it is essentially constant. By
varying [M], and measuring [M], the f3s may similarly be found.
A variety of optical methods is used to determine both complex formation
and stability constants. Job's method of continuous variations is the best
known. A wavelength is chosen at which the complex in question absorbs
(usually in the visible or near-ultraviolet region). Several solutions are
examined, for all of which ([MJ. + [L],) is a constant, C, although each has
different values for [M], and [L],. Here, [L], is the total concentration of
ligand, free and complexed. Some measure of the intensity of absorption
(absorbance or optical density) is plotted against composition (usually
against the ratio [LJ./([M], + [L],); that is, against [L],/C, a ratio we shall
call a. As a simple example of the application of this method consider the
case where only a single complex is formed. If a single complex is formed
then we have an equilibrium
M + nL <=> ML,
Initial concentration (1- ~)c ~c o
Final concentration (1- ~- y)C (~- n,)C yC
The values of [M], [L] and [ML.J all change with a; we wish to find the
relationship that holds when [ML.J is a maximum. To do this we differen-
tiate each of the above equations with respect to a. We obtain
d[M] = -C
d~
d[L] = C
da
To find the maximum in [ML.] we set the last equation to zero and
substitute the first two. The result is
[L]( -C) + [M](nC) = 0
Introducing the explicit expressions for the concentrations [L] and [M] we
obtain
-(~max- ny) + n(1- ~max- y) = 0
Determination of stability constants 1 79
1l
!6
FJC. &.2 The Job's plot obtained when only one €
complex species, ML,, is formed between M <~
and L
that is,
n
"'max=l+n
It follows that since ac is known (from the composition with which the
solution was made up) then, if the visible or ultraviolet wavelength chosen
corresponds to absorption by the complex, then n can be determined from
the maximum in plot of absorbance against cc. Conversely, if the absorption
at the chosen value is primarily due to either M or L, the minimum of the
absorbance versus ac plot gives n. Figure 5.2 shows a typical Job's plot for
a ML2 complex (ccmax = i). The rounding at the peak is due to the fact that
we have chosen to show the case of a not-very stable complex. The more
stable the complex the less rounded the peak; indeed, in favourable cases
the peak shape may be used to determine P•. When more than one complex
is formed there will usually be a corresponding number of peaks in a Job's
plot, but stability constants may only be determined from such a plot under
very special circumstances.
There is another method for the determination of stability constants
which, although not much used for this purpose, is of some value in reaction
kinetics as a method of estimating rate constants which cannot be measured
readily (see Chapter 14). Suppose the reaction
and the rate of disappeance of ML.+ 1 is kb[ML.+ 1], where kb is the rate
constant of the backward reaction. At equilibrium,
SO 1 Stability of coordination compounds
Therefore,
[Ml,+1] - K - kt
[Ml,][L]- n + l - if;;
That is, if the rate constants kc and kb can be independently determined, the
corresponding stepwise stability constant is given by their quotient.
So far, activity coefficients of unity have been assumed, but they seldom
have this value. Consequently, the stability constants determined above
are not constants at all, but are a function of concentration. The simplest
way out of this difficulty is to determine the stability constants over a range
ofreactant concentrations and then extrapolate to zero concentration (where
activity coefficients are unity). K. and fJ. at zero concentration are true
thermodynamic equilibrium constants and are distinguished by the super-
script T, thus TK. and TfJ. (although, to avoid complexity early in the
chapter, they were not designated as such in Table 5.1). More commonly,
activity coefficients are kept essentially constant by carrying out measure-
ments in the presence of a backing electrolyte which keeps the ionic strength
of the medium constant. Stability constants obtained in this way are
proportional to the thermodynamic stability constants, the constant of
proportionality being a function of the activity coefficients of the species
involved in the complex. Such stability constants are referred to as stoichio-
metric stability constants. Strictly, it is the determination of these which we
have discussed-they are given the symbols K. and fJ. which we have been
using.
and
Table 5.2 Classification of acceptor
species {after Ahrland, Chatt and Davies) N«P>As>Sb>Bi
In addition, there is a third class of electron acceptor for which the stability
Class-a behaviour
H, the alkali and alkaline earth metals, the constants do not display either class-a or -b behaviour uniquely. The
elements Sc -+ Cr, AI -+ Cl, Zn -+ Br, In, Sb class-ajclass-b classification of some metal ions is given in Table 5.2 (normal
and I, the lanthanides and actinides valence states are assumed).
Class·b behaviour Although not included within the above classification, there are some
Rh, Pd, Ag, lr, Au and Hg other useful gradations which have been noted and are conveniently included
Borderline behaviour
at this point. For a given ligand, corresponding stability constants of
The elements Mn -+ Cu, Tl -+ Po, Mo, Te, complexes of bivalent ions of the first transition series are usually in the
Ru, W, Re, Os and Cd natural order (sometimes called the Irving-Williams order):
provided that the ligand is not changed from one ion to the next and subject
to the absence of size-matching conditions that will be discussed later in this
chapter. Similarly, for approximately constant ionic radius, the stability
constants are in the order of decreasing charge, thus,
Th 4 + > Y3 + > Ca 2 + > Na+
and
La 3 + > s,.:<+ > K+
Table 5.3 Classification of some (formally) ionic species as hard and soft acids and
bases (after Pearson)
Borderline Zn 2 + Sn 2 + Pb2 +
Fe2 + Co2 + Ni2 + Cu 2 + Ru 2 + Os2 + Sb3 + Bi3 + Rh3 + lr"+
2 R. G. Pearson, J. Chern. Ed. (1987) 64, 561 and Inorg. Chern. (1988) 27, 734.
Stability correlations I 83
A
___ j ___ _
----------2-
Number of electrons ------
Table 5.4 Absolute hardness values for common metals and ligands
species present will include foreign anions-ones not involved in the reaction,
but arising from the salt used to maintain a constant ionic strength-and
some of these might well be coordinated to M. The addition reaction
Ml., + L-+ Ml.,+1
should more properly be written as a substitution reaction (L' is usually
H 2 0 but in this section is to be generally taken to mean 'solvent molecule'):
ML;,Ln + L-+ ML;,_1Ln+ 1 + L'
In this reaction it has been assumed that the number of ligands around M
is constant, and equal to (m + n). If we let this number (usually four or six)
be N, then the reaction may be written as
Ml.N-nl., + L-+ Ml.N-n-1Ln+1 + L'
As has been seen in Table 5.1, it is commonly found that for a given M and
L there is a decrease in successive formation constants, K 1, K 2 .•• or /3 1,/32
... This is largely a statistical effect. It is easier to attach L to ML~ than to
ML~_ 1 L because there are N reaction sites on the former but only
N- 1 on the latter (replacement of L in ML~_ 1 L by L gives an identical
molecule).
Consider the equilibrium:
Ml.N-nl., + L <" Ml.N-n-1l.,+1 + L'
The rate of formation of ML~-.- 1 L.+ 1 is, for a simple one-step reaction,
k,[ML~-.L.][L]. If the statistical effect is the only one which varies as n
changes, the forward rate constant will vary with n in proportion to the
number of reaction sites, that is, the number of L' ligands in ML~-.L•. We
therefore rewrite the above rate as k((N- n)[ML~_.J[L'], where k( is
expected to be constant for all values of (N- n). Similarly, for the rate of
disappearance ofML~-.- 1 Ln+ 1 , kb[ML~-.- 1 L.+ 1 ][L], the backward reac-
tion rate constant, kb will vary with n in proportion to the number of L
ligands in ML~-.- 1 L.+ 1 • Hence, in the above cases kb = k~(n + 1), where
k~ is constant for all values of (n + 1). It follows that
k, ki(N -n)
Kn+1 = !(.; = kt,(n + 1)
In a similar way it is found that
K =kf(N-n+1)
n kt,n
Table 5.5 Statistical predictions and experimental ratios for the equilibrium con-
stants of the system
[Ni(H20l6-n(NH3ln] 2 + + NH3 :;:=:: [Ni(H20ls-n(NH3)n_ 1 ] 2 + + H20
Ratio n Experimental ratio Statistical
(data from Table 5.1) prediction a
Equilibrium• logK
Non-chelated complex
Ni 2 + + 2NH 3 :;:=:: [Ni(NH 3 )2]h 5.05
[Ni(NH3) 2 ] 2 + + 2NH3 :;:=:: [Ni(NH 3 ) 4 ] 2 + 2.96
[Ni(NH 3) 4 ] 2 + + 2NH 3 :;:=:: [Ni(NH 3 )6 ] 2 + 0.73
Chelated complex
Ni 2 + +en:;:=:: [Ni(en)] 2 + 7.51
[Ni(en)] 2 + +en:;:=:: [Ni(en) 2 f+ 6.35
[Ni(en) 2] 2 + +en""' [Ni(en) 3 ] 2 + 4.32
a Coordinated water is omitted for simplicity.
and mol- 1 L, respectively). How does one compare quantities with different
units? This is not the end. When we write RT InK (as we shall do shortly)
we have to recognize that we can only take the logarithm of a number.
Where have the units of K gone? The answer here is that we really evaluate
RT ln(K/1), where the I refers to a standard state which has the same units
as K. This leads to a discussion of the characteristics of the standard
state and this, in turn, to the recognition that there are three different
concentration scales in use: molarity, molality and mole fraction. These
concentration scales are not directly proportional to each other so that, for
example, it is theoretically possible for a given solution to have activity
coefficients of unity on one scale but not on the others. The chelate effect is
not enormous and so it is not surprising that such fine points can assume
undue prominence and that controversy and apparent misunderstanding
exist! A very readable account of the general problem can be found in
Chapter 2 of the book by Connors and in the articles given at the end of
the chapter. Fortunately, we can side-step most of the pit-falls and learn
more about the chelate effect from a more detailed analysis of stability
constant data. To do this, we note the relationships
which show that the chelate effect could originate in either the heat term,
!.lH 0 , the entropy term, !.lS0 , or both.
To proceed further, we could analyse the equilibrium constants in Table
5.6, in pairs, using these thermodynamic relationships. We can simplify
matters by considering, instead, the equilibrium
These results are qualitatively general: the chelate effect is largely an entropy
effect; for non-transition metal ions the heat term is particularly small.
There are two important contributions to the !.lH 0 term when it makes
a significant contribution. First, when two monodentate anionic ligands
are brought together to occupy adjacent coordination sites in a complex,
there will be an electrostatic repulsion between them against which work
has to be done. The same is true for uncharged monodentate ligands because
such ligands are always dipolar. For chelating ligands the coordinating
centres do not have to be brought together and most of this repulsive
energy is 'built in' (it makes a contribution to the enthalpy of formation of
the ligand). That is, in the above equilibrium this contribution to the heat
term would be expected to favour chelated species because the oriented
ammonia molecules repel each other. The second, more variable, contribu-
88 1 Stability of coordination compounds
tion to the 11H0 term comes from solvation energies. Each of the species in
the equilibrium will be solvated (by hydrogen bonding, for example). Further,
there will be anions closely associated with the cationic species. We shall see
in Chapter 14 that there is kinetic evidence that counterions, although not
directly bonded to the coordination centre--they are not in the first
coordination sphere--are often present in an outer sphere (the so-called
second coordination sphere). It does not seem possible to predict, in general,
whether this term makes a positive or negative contribution to 11H0 , but it
appears to make a positive contribution when the coordination centre is a
main group element.
The entropy effect is readily understood. An increase of randomness
is associated with a concomitant increase in entropy. A complex molecule
has a lower entropy than its separated and therefore independent com-
ponents. In the equilibrium above there is a 50% increase in the number of
independent molecules in the right-hand side compared with the left and,
so, the right-hand side is favoured. Another aspect of the entropy effect is
the following. When one end, and only one end, of an ethylenediamine
molecule is coordinated, the effective concentration of the other end in the
system, and the probability that it will coordinate, is high, because it is
constrained to stay close to the cation. This means that it is easier to form
a chelate ring than to coordinate two independent molecules because the
two acts of coordination are related for the former, whilst for the latter they
are entirely independent of each other. All these, then, are potential
contributors to the chelate effect. They are difficult to relate to the kinetic
data on the effect. These show that it occurs because the ring opening
is slower than expected for the dissociation of the first of two independent
ligands, not because the formation of the ring is faster (it is the ratio of these
two that is the equilibrium constant K ). It is understandable that the topic
continues to be the subject of debate.
The chelate effect varies with the size of the ring formed on coordination. It
is usually a maximum for five-membered rings and only slightly smaller for
six-membered rings. It is not difficult to see one reason why this should be.
For an octahedral complex, the angle subtended by a chelate ring at the
metal, no matter the size of the ring, will be close to 90°. For simplicity, we
fix it at this value. What will be the bond angles at the other atoms in the
ring? Assuming planar rings, the average values will be:
4 90'
5 112'
6 126'
7 135'
So, if in a saturated ring system, such as that formed by en, the five-membered
ring system will have angles that are closest to the natural, tetrahedral, value
(109.SO). Similarly, for a ring that is conjugated, as in acac-, which may be
drawn as shown in Fig. 5.4, the six-membered ring has an angle closest to
the natural trigonal value of 120°. Of course, ring puckering can and does
occur, but can only be of a large magnitude at a cost of either steric
Solid complexes I 89
5. 7 Sterle effects
As has been indicated above, steric effects can play a part in determining
the stability of complexes. Indeed, this has become an area of much research.
If a very bulky ligand coordinates it may well form a weak bond, in which
case molecules containing it may well be rather reactive. If, on the other
hand, despite its bulk it coordinates strongly, then it may well cause the
bonding of other ligands to be weaker than normal. Again, an enhanced
reactivity is likely. Enhanced reactivity, particularly in the field of catalysis,
in which, often, a ligand dissociates from a metal ion at one point in the
mechanistic sequences and recoordinates at another, is always interesting.
A question at once arises: can we classify steric effect? The answer is that we
can, to some extent. Before attempting this, however, there is an important
point which has to be made. Consider a ligand such as ammonia but consider
it as a free molecule. The energy required to make it rotate, to spin, about
its threefold axis is low-it falls in the microwave region of the spectrum.
When the ammonia molecule is coordinated to a metal this rotation becomes
less free--it may well become a libration, in which it oscillates about its
threefold axis as a solid unit but does not overcome the barrier preventing
it reaching an equivalent position-but it is still of low energy. Everything
depends on the thermal energy available compared to the barrier height. At
room temperature the NH 3 groups in ammine complexes are commonly in
a state of large-amplitude oscillation, if not almost free rotation. This is
important because it means that the 'bumps' on the ammonia molecule--the
hydrogen atoms-may well behave as rather smeared out as far as steric
effects are concerned. It may not be too bad an approximation to regard
each ligand as effectively acting as a circular cone, with its apex at the metal
atom. This immediately leads us to some qualitative predictions. As far as
steric effects are concerned, we would expect the linear M-N=<:~S and
M-N=<:~e systems to be more stable than their bent isomeric counterparts
M-S M-Se
"'-c "'-c
~N ~N
A case where this may well be evident is in the square planar complex formed
between the tridentate ligand diethylenetriamine, NH 2-CH 2-CH 2-NH-
CH2-CH2-NH2, dien, and palladium( II), [Pd(dien)(SeCN)]. In this complex
the SeeN- anion is coordinated through Se and is bent. When the terminal
Steric effects 1 91
cone angle
Table 5.7 Cone angles for some simple systems, see Fig. 5.5
(ligand charges are not shown
H 75 PH 3 87
Me 90 PF3 104
2.2aA F 92 P(0Me) 3 107"
CO,CN,N 2 ,NO ~95 P(0Et) 3 108"
Cl, Et 102 PMe 3 118
Br, C6 H5 105 PCI 3 124
FIJI. &.& The definition of the cone angle tor a
symmetrical phosphine. PR3 • It is necessary, of I 107 PEt3 132
course, to define the cone angle with respect to i-Pr 114 PPh3 145
the metal atom to which the phosphine is t-Bu 126 P(i-Pr) 3 160
coordinated. It is assumed in the definition that CsHs 136 P(t-Bu) 3 182
this atom is 2.28A from the phosphorus. For a
a In obtaining these values, Tolman used a conformation which has yet to be
different coordinated atom a different
observed.
metal-ligand separation would have to be
assumed.
NH 2 groups of the dien ligand are replaced by NEt 2 groups, the SeeN-
anion is coordinated through nitrogen and is linear. Steric effects seem to
be the factor causing the change. Can the effect be treated in a reasonably
quantitative manner? The answer is yes. By a study both of crystal structures
and of molecular models, Tolman3 has compiled data for steric effects,
described in terms of cone angles, defined as in Fig. 5.5, for a large number
of ligands, particularly for those in which phosphorus is the donor atom. A
selection of his data is reproduced in Table 5.7. There have been many
variants in this field-the use of X-ray data to determine cone angles, the
use of detailed theoretical calculations, methods of averaging when very
different substituents are present, particularly on a phosphorus atom, and
so on. Generally, these other approaches lead to cone angles which are
somewhat smaller than those given by Tolman.
A rather different form of size effect occurs for those ligands which are
cage-like and which largely envelop a cation at their centre. For these, it is
the matching of the size of the central cavity of the ligand to the size of the
cation which determines stability. In Chapter 3 the cryptand ligands were
mentioned; these are shown again in Fig. 5.6. They are given shorthand
names which simply list the number of oxygen atoms in each bridge between
the two nitrogen atoms. Also in Fig. 5.6 are given the approximate radius
of a hypothetical spherical cavity at the centre of each ligand. These values
are to be compared with the approximate ionic radii of the alkali metal
cations, as determined experimentally by X-ray data diffraction,4 of Li +
0.76A, Na+ 1.02A, K+ 1.38A, Rb+ 1.52A and cs+ 1.67A. The relative
stabilities of the alkali metal cryptand complexes are
[211]
o.sA [221]
Fig. 5.6 Four cryptand ligands. The three 1.1A
numbers within the square brackets correspond
to the three chains linking the two nitrogen
atoms and give, in descending numerical order,
the number of oxygen atoms in these chains.
Below each ligand designation is given the
approximate size of the cavity within that ligand.
[222]
[322]
1.4 A
1.8A
5.8 Conclusions
Although in this chapter many aspects of stability have been discussed, there
are many others that have not. It is now accepted that for some ligands,
metal-ligand double bonding occurs. 5 Just how are bond order and stability
related? Is the distinction between inert and labile transition metal complexes
related in any way to the d-electron configuration of the transition metal
(we shall have more to say about this in Chapter 14)? In addition to the
kinetic trans effect, discussed at the end of Chapter 4, there is a static trans
effect, sometimes called the trans influence. In this, a metal-ligand bond
5 See W. A. Nugent and J. M. Mayer, Metal-Ligand multiple honds. Wiley, New York, 1988.
Further reading I 93
• 'Steric Effects of Phosphorus Ligands', C. A. Tolman, Chem. • 'Hard and Soft Acids and Bases-the Evolution of a Chem-
Rev. (1977), 77. 313. ical Concept', R. G. Pearson, Coord. Chem. Rev. (1990), 100,
• 'Ligand Interactions in Crowded Molecules', H. C. Clark and 403.
M. J. Hampden-Smith, Coord. Chem. Rev. (1987) 79, 229. • 'Absolute Electronegativity and Hardness', R. G. Pearson,
• 'Ligand Steric Properties', T. L. Brown and K. J. Lee, Coord. Chem. in Britain (1991) 444.
Chem. Rev. (1993) 128. 89. • 'Molecular Organization, Portal to Supramolecular Chem-
• 'The Specification of Bonding Cavities in Macrocylic Lig- istry. Structural Analysis of the Factors Associated with
ands', K. Hendrick, P. A. Tasker and C. F. Linday, Prog. Molecular Organization in Coordination and Inclusion
Inorg. Chem. (1985) 33, I. Chemistry, including the Coordination Template Effect',
• 'The Chelate Effect Redefined', J. J. R. Frausto da Silva, D. H. Busch and N. A. Stephenson, Coord. Chem. Rev. (1990)
J. Chem. Educ. (1983) 60, 390. 100, 119.
• 'Misunderstandings over the Chelate Effect', D. Munro,
Chem. in Britain (1977) 100.
Questions 5.3 Use the data in Fig. 5.2 to show that, as claimed in the
caption, it represents the formation of the ML2 complex.
5.1 In Fig. 3.27 was shown an example of a compound with
a low barrier to internal rotation. There are many similar
compounds, some of which spontaneously ignite when exposed 5.4 The experimental data used in Table 5.5 are those for
to air. Comment on the applicability of the terms kinetic zero ionic strength in Table 5.1. Calculate the corresponding
stability and thermodynamic stability to such species. experimental data for 2 M NH4 N03 (given in the caption to
Table 5.1 ). Comment on the relative agreement of the two sets
5.2 (Note that in this question square brackets indicate con- of experimental data with experiment.
centrations.) The consequences of the addition of base to
aqueous Cr 111 was followed for four years. [Cr 2 (0H),] 4 +
maximized after a few days whilst [Cr 3 (0H) 4 ]<+ increased 5.5 Compare Tables 5.2 and 5.3 and thus comment on the
throughout the period. [Cr4 (0H) 6 ]6+ was constant after a few question of whether the class-a and -b, and hard and soft
months. Comment on these observations in the light of the classifications are simply restatement of the same experimental
reported K •• values: observations.
[Cr0 (0H)•]"+
5.6 Use Table 5.4 to associate numbers with the broad
Kab = [Cr(OH),]+[cr•• 1 (0H)•-2]<n il+
divisions given in Table 5.3 (for instance, what is the range of
logK22 = 5.1; logK34 = 6.9; logK46 = 5.2 absolute hardness values that corresponds to hard?).
Molecular orbital theory of
transition metal complexes
6.1 Introduction
The traditional approach to the electronic structure of transition metal
complexes (which is the subject of the next chapter) is to assume that the
only effect of the ligands is to produce an electrostatic field which relieves
the degeneracy of the d orbitals of the central metal ion. The most serious
defect of this model is that it does not recognize the existence of overlap,
and hence the existence of specific bonding interactions, between the ligands
and the metal orbitals. Yet calculations which assume reasonable sizes for
the orbitals (together with a considerable body of physical evidence which
will be reviewed in Chapter 12) point to the existence of overlap. How should
this be taken account of? The simplest answer is to be found in the
L•
application of symmetry ideas to the problem, and this is the subject of this
chapter. In it the readet' will be assumed to have some familiarity with the
basics of group theory. Appendix 3 gives an introduction to the subject; the
following lines are intended to provide a brief overview of aspects needed
to make a start on the present chapter.
It is convenient to focus the discussion on octahedral complexes (Fig. 6.1 ).
These are molecules with high symmetry, possessing, for example, fourfold
------·
"1M rotation axes, threefold rotation axes, twofold rotation axes, mirror planes,
and a centre of symmetry (Fig. 6.2). Symbols such as a 19 ('aye one gee '), e9
and t 19 are used to distinguish the behaviour of different orbitals or sets of
(b) orbitals under the various operations associated with the rotation axes, mirror
planes etc. of the octahedron. Note the shift from 'symmetry elements' (such
fiC. 6.1 (a) An octahedron. (b) An octahedral
complex. The perspective adopted here is the as rotation axes, mirror planes) Fig. 6.2, to 'symmetry operations'- the act
same as that for (a), and was chosen to make of rotating, reHecting etc. In group theory one is concerned with symmetry
(a) as easy as possible to visualize. The way operations, rather than with symmetry elements. Lower case symbols are
that an octahedral complex is drawn will vary,
the perspective adopted depending on the point used to denote orbitals or, more generally, wavefunctions. So, a point to
under discussion. which we shall return, an s orbital of an atom at a centre of an octahedral
96 1 Molecular orbital theory of transition metal complexes
the present chapter. In the following chapter it will be necessary to explore this degeneracy in
much more detail and so a detailed discussion is deferred until then.
Octahedral complexes I 97
X~
abbreviated form in Table 6.1 and also in Appendix 3 (Table A3.3). Because
the 4s orbital (Fig. 6.3) of the central metal is turned into itself by all the
operations of the group--multiplied by 1-its behaviour is described by the
A 1 irreducible representation of Table 6.1. In the full octahedral group
Oh (which, unlike the group 0, contains the operation of inversion in a centre
of symmetry), the label describing the behaviour of the 4s orbital is A 1 • (note
that an equivalent way of talking about an orbital labelled a 1 • is to say that
it 'is of A 1 • symmetry' or that 'it transforms as A 1 .'). In similar fashion, the
set of three 4p orbitals (Fig. 6.4) have T 1" symmetry (T 1 in Table 6.1). Finally,
the five 3d orbitals split into two sets, of E. (Fig. 6.5) and T 29 (Fig. 6.6)
symmetries (E and T 2 in Table 6.1). This latter splitting, which is denoted
11, is of crucial importance to the understanding of the properties of transition
metal complexes. One considerable attraction of crystal field theory, the
subject of the next chapter, is that it gives a very simple physical explanation
of this splitting. In the present context it is perhaps most easily regarded as
a splitting which is group-theoretically allowed and so, presumably, may
well exist. As the chapter develops we shall find good reasons for the
splitting.
Evidently, the next step is to classify the ligand orbitals according to their
symmetry types. Two problems arise. First, for a regular octahedral complex,
all six ligand orbitals look alike-how then can they be classified differently?
Second, a characteristic of a symmetry-classified set is that all of those
symmetry operations which send an octahedron into itself send one member
Table 6.1 The 0 character table of the set into itself, another member or a mixture of members of the set.
The reader who is not familiar with this pattern is asked to take it on trust
0 E 6C4 3C2 6C2 8C3 for the moment. It will be demonstrated in the next chapter in showing the
degeneracy of the metal dx'- Y' and dz, orbitals in an octahedral crystal field
A, 1 1 1 1 1 (Section 7.3); alternatively, Appendix 4 is dedicated to the problem. As these
symmetry operations send one ligand a orbital into another, why do these
A2 1 -1 1 -1 1 orbitals not already constitute a set? The answer is, they do. However, this
set can be broken down into smaller sets and it is to these latter that
E 2 0 2 0 -1
symmetry labels are attached. 3 In these sets the individuality of the ligand
Tl 3 1 -1 -1 0 a orbitals is lost. We talk of the wavefunctions of various sets of ligand
orbitals rather than of the wavefunctions corresponding to individual ligand
T2 3 -1 -1 1 0
3 Appendix 6 shows how these subsets may be obtained.
Octahedral complexes 1 99
z z z
t t t
z
z z
I
z z
shown in Fig. 6.7, the labels being chosen in conformity with the convention
given in Fig. 3.4. As mentioned earlier, in the explicit forms given in Table
6.2 it has been assumed that the ligand a orbitals do not overlap each other,
although non-zero ligand-ligand overlap integrals would only affect the
normalization factors (the numerical coefficients at the front of each
expression) given in the table, not the general form of the combinations.
The ligand a group orbitals listed in Table 6.2 are of A 19 , E 9 and T 1u
symmetries and each set will overlap with metal orbitals of the same
symmetries, so that the 4s(a 19 ), 4p(t 1.) and 3d(e9 ) orbitals of the metal will
be involved in a bonding. Of the valence shell orbitals of the metal atom
only the 3d(t 29 ) orbitals are not involved in this bonding. The interactions
of A,., T 1• and E. symmetries are shown pictorially in Figs. 6.8-6.10, where
it can be seen that there is the expected close matching between ligand group
orbitals and the corresponding metal orbitals. If symmetry had not been
Octahedral complexes 1 101
applied to the u bonding problem then we would have had to consider the
interaction between the six ligand u orbitals and nine metal orbitals, with
no means of knowing in advance that three of the 3d are not involved in
this u bonding. In contrast, not only does the symmetry break the problem
up into subproblems-the A 19 , E9 and T,.- but for each degenerate case,
E9 and T1., only one member of each set need be considered (by symmetry,
consideration of any other member must lead to the same result). That is,
for each of the three subproblems only the interaction between a single metal
orbital and a (delocalized) ligand orbital has to be treated. For simple models
the calculations really can be done on the back of an envelope! Fig. 6.11
shows, schematically, the energy level pattern obtained as a result of these
u interactions. It includes metal and ligand electrons; the detailed problem
of how to include those originating on the metal will be a major concern of
Chapters 7 and 8. Note, particularly, the splitting d in Fig. 6.11. Many
FJC. 6.8 The metal-ligand a bonding in an
octahedral complex involving orbitals of A"'
symmetry. The metal s orbital is shown at the
centre.
z z
X
......-- --........ y --........ y
Blg(2)
4p====:
4s------(
eg(2)
eg(1)
tlu (1)
a1g(l)
Fig. 6.11 A schematic molecular orbital energy research groups have tried to carry out reasonably accurate calculations
level scheme for an octahedral complex of a aimed, in part, at obtaining molecular orbital energy level schemes (such as
first row transition metal ion (only a interactions
are included). The ligand orbitals are shown that in Fig. 6.11) accurately, although this is a difficult task. It in this
occupied; the metal d orbitals are partially filled. endeavour that many of the methods to be discussed in Section 10.3.2 were
The example shown is appropriate to a d7 ion in developed.
a strong-field (low-spin) complex. The full
meaning of these latter terms will only become Figure 6.11 indicates that the idea that the ligands function as electron
evident in Chapter 7. The numbers in donors is correct. Electrons which, before the interactions were 'switched
parenthesis follow the established convention of on', were in pure ligand orbitals are really in delocalized molecular orbitals
counting from the bottom up. So, the bottom
a'<l is a,..(1), the next a,..(2). which take them onto the metal atom-the electron density on the metal
atom is increased as a result of the covalency. The consequences of this for
the concept of formal valence states will be discussed later. The most
important feature of Fig. 6.11 is the fact that the two lowest unoccupied
orbital sets are, in order, t 29 and e.(2) (this latter being weakly u antibonding).
So far, no metal electrons have formally been included and so these
unoccupied orbitals have to accommodate them. Of course, the number of
electrons occupying these orbitals is the same as the number present in the
valence shell orbitals of the metal before the metal-ligand interaction was
switched on; the d electrons of the uncomplexed metal ion may be regarded
as being distributed between the t 2 • and e.(2) molecular orbitals. This is a
situation which will be explored in the next chapter where care will be taken
to state that .1., the label given to the energy separation between a lower t 2 •
and an upper e. set, is an experimental quantity. What the next chapter will
Octahedral complexes 1 103
not show is the way that the e9 set is antibonding. The present discussion
also leads us to recognize that metal electrons in the eg(2) orbitals are to
some extent delocalized over the ligands. In Chapter 12 experimental
evidence will be adduced in support of this conclusion.
'•,,,.•
:e:····.;:: Px
104 1 Molecular orbital theory of transition metal complexes
. . --.~ r (_-. ~-
pattern holds for all the other " combinations
shown in the following figures. (b) One of the
three ligand "group orbitals of T2 u symmetry. As
for all other" combinations in adjacent figures,
-(~F·-----i\:)
the combination shown is the first listed in the
·~
appropriate section of Table 6.3.
• l •II
\
\I
I
II
\_J
\I
(b)
1
g orbitals as T 1.). The ligand n group orbitals ofT1 • and Tzu symmetries are
~(x2z + 1C3z + x 42 + n5z) therefore carried over, unmodified, into the full molecular orbital description.
t1u ~(ltlx + """ + """ + lt4x) A representative example of each is shown in Fig. 6.13. We are left with Tz.
~(xly + n3y + "5y + ns,> and T~u sets. The T'" set will interact with the metal p orbitals (also of T 1u
symmetry), but it is simpler to think of their effect on the occupied T~u u
1 ~(n2x + nay- 1t4x - 1tsy) In this case, illustrated in Figs. 6.14 and 6.15(c), it follows that the T 1u
symmetry ligand n orbitals are also occupied. Interaction with the t '"(1)
t2g ;(nly + lt3z- " " ' - n.,.)
molecular orbitals of Fig. 6.11 will raise or lower this latter set depending
~("'"' + 1C2z- " " ' - 1C5z) on whether its energy is higher or lower than that of the ligand n(T~u) set.
1 ~(lt2z- "3z+ lt4z- lt5z) As long as we retain two occupied t'" sets the orbital occupancy is unaffected
t2u + """- n4x)
'(ltlx- lt2x 4 Appendix 4 in Symmetry and Structure, 2nd edn., by S. F. A. Kettle, Wiley (Chichester and
'("1y- "ay + "5y- "sy) New York) 1995. Note that the atom labelling pattern used in this text differs from that in the
present.
Octahedral complexes 1 105
by this interaction and, because the t 29 - eg(2) separation is our concern, for
this case we may forget the t lu orbitals and their interactions.
,-, ,-, In this case it follows that the ligand n combinations of T 1• symmetry are
also unoccupied. Interaction between the ligand nand (occupied) t~u(l) sets
I+ +I
of Fig. 6.11 will result in a repulsion between the two and the t~u(1) set will
' '
··~----+-+-------·
'1- '
,_,I '
,_,
I -I \
be stabilized somewhat but the orbital occupancy will remain unchanged.
This is illustrated in Figure 6.15(b). We could, as a separate case, consider
the interaction between the t 1.(2) molecular orbital set of Fig. 6.11 and the
ligand n(t~u) set. Again, this would only be necessary if the interaction leads
Fig. 8.14 A ligand " group orbital of Tw to one of the orbital sets becoming occupied. In fact, there is no recognized
symmetry and the molecular orbital of Fig. 6.9
with which it interacts (that in110lving the metal case in which the ligand n(t~u) set, whether occupied or not, needs to be
Pz orbital). considered.
We are left with the ligand n set of T 29 symmetry. This set may interact
with the metal t 29 orbitals (dxy• d,., d,x) which, so far, have been non-bonding
because there is no ligand a combination ofT29 symmetry. The consequences
eg(2)
~
t1u
eg(1)
t1u (1)
t1u
81g(1)
(a)
Occupied
ligand 1t t1u
Unoccupied
ligand 1t t1u (c)
(b)
t2g
Ligand
eg Metal
Metal t2g
t 2" - e.(2) separation of Fig. 6.11. This behaviour is clearly consistent with
z the observation that the cyanide anion gives rise to large values of !J.; it exerts
a very large ligand field. In Chapter 10 we will see that the carbon monoxide
ligand behaves similarly.
Finally, a cautionary note. There is theoretical evidence, despite all that
has been said in this chapter, that in a complex such as [Cr(CN) 6 ] 3 - the d
electrons are not in the highest occupied orbitals. Rather, that all the CN-
n occupied and, indeed, the CN- (J bonding orbitals, all occupied, have
higher energiess The reason for this apparently ridiculous behaviour is to be
found in a phenomenon which so often wrecks our simple pictures-electron
repulsion within the molecule. How can there be filled orbitals above
partially occupied ones? The increase in electron repulsion energy in taking
an electron from a (delocalized and therefore diffuse) n occupied or (J bonding
orbital and putting it into a (largely localized and a therefore concentrated)
metal d orbital •costs' more than the energy gained from moving the electron
into what is a more stable orbital. The holes in the d orbitals are protected!
z Fortunately, this complication turns out to be unimportant for the discussion
in this chapter. It doesn't really matter too much where the d orbitals are
relative to the ligand orbitals. As has been commented 'if one is only
interested in the energy pattern, ligand field theory remains a reliable guide'.
Since virtually all of the measurements made on transition metal complexes
are concerned with the detailed energy pattern, all is well. The reader who
is either unhappy with this situation or is so interested in it that they wish
to learn more, should turn to Section 10.7 where it is encountered again, in
the context of ferrocene, and discussed in more detail. In Section 12.7
measurements that go beyond energy patterns will be described and com-
pared with theory, whereupon the relative energies of the ligand orbitals will
become of importance.
FJg. 8.18 The relationship between a cube and borne et a!., lnorg. Chern. (1984) 23, 1677; Excited States and Reactive Intermediates, ACS
a tetrahedron. Symposium Series 307 (1986) 2; and Inorg. Chern. (1991) 30,2978.
108 1 Molecular orbital theory of transition metal complexes
z
Table 6.4 Ligand group orbitals of (J symmetry in a tetrahedral
complex; the labels used are those of Fig. 6.19
E, 8C 3 and 3C2 ). Using the orbital labels shown in Fig. 6.19, the A 1 and T 2
ligand group orbitals are given in Table 6.4 and pictured in Fig. 6.20, a figure
04 which also shows the metal orbitals of the same symmetries.
As before, only orbitals of the same symmetry species may interact with
Ftg. 6.19 The labels used in the text for the
ligand u orbitals of a tetrahedral complex. each other. The A 1 interactions are straightforward but the T 2 more
complicated than the corresponding octahedral case because both metal sets
4pxo 4pY, 4pz and 3dxy• 3dy, 3dzx have T 2 symmetry and so interact with
the ligand orbitals of this symmetry. It does not appear possible to make a
general statement about the details of the outcome of the T 2 interactions but
that given in Fig. 6.21 is about as close as one can get. Figure 6.21 gives a
schematic a bonding-only molecular orbital diagram for a tetrahedral
complex. As we shall see in the next chapter, just as in octahedral complexes,
the more stable d orbital set according to crystal-field theory, that of e
symmetry, is not involved in a bonding but the less stable, the tz, is involved.
Many tetrahedral complexes involve the oxide anion as a ligand (e.g. the
Mo~- anions commonly known as permanganate, chromate, and ferrate)
and contain a metal atom in a high formal valence state. Since oxygen
is usually regarded as forming two covalent bonds and because a high metal
charge will favour ligand-to-metal charge migration, it can be anticipated
that n bonding may well be of potential importance for such tetrahedral
complexes. Unfortunately, the consequences of n bonding are not as clear-cut
as for octahedral complexes. An important difference between octahedral
(a)
and tetrahedral complexes is that the latter do not have a centre of symmetry.
Such a centre of symmetry separates the p orbitals from the d orbitals on
the metal (the ps are ungerade and the ds are gerade) and no centrosymmetric
ligand field can mix them. A tetrahedral ligand field can mix ps and ds on
the metal and there is good evidence that such mixing occurs~it will be met
in Chapter 8.
The n bonding problem in tetrahedral molecules starts difficult and
remains so throughout. It is not a trivial task to demonstrate that the ligand
n orbital symmetry-adapted combinations are of T 1 + T2 + E symmetries.
To show this it is important to chose the orientation of the ligand n orbitals
carefully if the task is to be made (relatively) easy. Guidance is given in Fig.
(b) 6.22. Appendix 4, and the character of -1 described at the end, may well
Ftg. 6.20 (a) Interaction of the ligand a1 group be needed also. The next step, that of the generation of the ligand group
orbital of Table 6.3 with the metal s orbital. orbitals, is not trivial. 6 The explicit expressions for the ligand group orbitals
(b) Interaction of the first ligand t 2 u group
orbital listed in Table 6.4 with the 6 A fairly simple derivation is given in detail in S. F. A. Kettle, Symmetry and Structure, 2nd
s ,;;;Bl;,__ _ _..(
~+e ~
Fig. 8.21 A schematic molecular orbital energy d
level diagram for tetrahedral complexes with
~~~~~~K~~-----c~=:::;:J=~~)
only u interactions included. The ligand u
electrons are stabilized by interaction with the
corresponding metal orbitals. The (three)
electrons originally in the metal d orbitals
correspond with those distributed between the
e(l) and t 2 (2) molecular orbitals.
81 (1)
1t2v
1t3v
are given in Table 6.5 using the notation of Fig. 6.22 and illustrated in Figs.
6.23-6.25, the relevant metal orbitals being included in the e and t 2 cases.
For tetrahedral complexes 1t bonding involves all of the metal d orbitals, not
just one set (as in the octahedral case), and it is not possible to give a simple
diagram analogous to Fig. 6.16. Both the t 2 (2) and e levels of Fig. 6.21 will
110 1 Molecular orbital theory of transition metal complexes
orbitals for common geometries and compared with the symmetries of the
orbitals comprising the valence set of the transition metal ion. This table is
group theoretical in origin and is designed to enable the reader to develop
a qualitative bonding argument for almost any complex with up to eight
ligands with any significant symmetry (if a molecule is only slightly distorted
from a high symmetry it often pays to pretend that there is no distortion at
all, as a first approximation). The reader who is unhappy about the plethora
oflabels in Table 6.6 need not be too concerned-these labels can be regarded
as simply indicating what interactions can occur-labels have to be identical
for an interaction to be possible; there is no vital need to enquire into their
deeper (group theoretical) significance.
As an example of the use of Table 6.6 its application to a trigonal
bipyramidal complex of D3h symmetry (Fig. 6.27) will be outlined. First, it
has to be recognized that our discussion of octahedral (Oh) and tetrahedral
(Td) complexes will be oflittle direct use in discussing ML 5 (D3 h) complexes.
Fig. 6.25 A ligand " group orbital of T2 Had we been concerned with complexes with either six or four ligands,
symmetry together with the metal p, with which
it ove~aps. The dominant orbital overlaps are however, it might well have been a good idea to start from the appropriate
indicated. The form of this ligand group orbital high symmetry case, as mentioned above. As Table 6.6 shows, in a ML 5
is particularly simple for the reason given in the complex the metal d orbitals split up into three sets, d., has A; symmetry,
caption to Fig. 6.24. Note carefully the phases
of the ove~ap--they are not incorrect! dx'-y' and dxy, together, are of E' symmetry whilst dyz and d.x, together,
have E" symmetry. The somewhat surprising fact that dx'-y' pairs with dxy
4s - - - - - . . . - -
t<1tet
Ligand 1t
(e+ t1 + t2)
'---y--/
Dodecahedral, MLa o,. a, b, e, a, b, b,
'---y--/ '---y--/
Square antiprism, MLa o.. a, b, e, a, e,
and not d.2 is easily explained-it is a general rule that the axis of highest
symmetry is (almost7) always chosen as the z axis, so here we choose the
threefold rotation axis; for an octahedron the z axis was a C4 and for a
tetrahedron an S4 .
The next step is to include, qualitatively, u bonding in the picture. The
five ligand group u orbitals are of2A~ + A2 + E' symmetries. They can easily
be obtained by the methods of Appendix 6 (treat non-equivalent ligands
separately) and are pictured in Fig. 6.28. From this figure it is evident that
7 When working with an icosahedral molecule such as [B 12 H 12 ] 2 - or C 60 life is much easier if
a C2 (rather than a C,) axis is chosen as z because x andy can then also be orientated along
C2 s and the three coordinate axes are symmetry-related. With z chosen to lie along a C, axis
the three coordinate axes are not symmetry-related.
Complexes of other geometries 1 113
dyz dzx t1 1t
t2 al + t2 e+ t 1 + t2
~
eg 2alg + a2u +big+ eu a2g + b2g + a2u + b2u + 2eg + 2eu
~
eg alg + b2g + eu a2g + b2g + a2u + b2u + eg + eu
~
e a 1 + b2 + 2e 2a 1 + 2a 2 + 4e
~
e'" 2a1 +a;+ e' a:, +a; + 2e' + 2e"
~ the threefold axis is the z
e 2a 1 + 2e 2a 1 + 2a 2 + 4e
axis
~
e 2a 1 + e a1 + a2 + 3e
~
e 3a 1 + 2e 2a 1 + 2a 2 + 5e
~
e 2a 1 + b 1 + e a 1 + b 1 + a 2 + b2 + 3e
~
e 3a 1 + b 1 + e al + bl + a2 + b2 + 4e
b2 bl 3a 1 + b 1 + a 2 + b2 2a 1 + 4b 1 + 2a 2 + 4b 2
b2 bl 2a 1 + b1 + b2 2a 1 + 2b 1 + 2a 2 + 2b 2 b 1 and b 2 may be interchanged by some
authors
b2 bl 3a 1 + 2b 1 + a2 + b2 3a 1 + 4b 1 + 3a 2 + 4b 2
~
e 2a 1 + 2b 2 + 2e 2a 1 + 2b 1 + 2a 2 + 2b 2 + 4e
~
e3 a 1 + b 2 + e 1 + e 2 + e3 a 1 + b 1 + a 2 + b2 + 2e 1 + 2e2 + 2e3
both of the a~ ligand orbitals may interact with the same metal a~ orbital.
The relative importance of the two interactions will depend on the geometry
of the complex-which of the (two) axial and the (three) equatorial ligands
are the closer to the metal? Usually, the equatorial. The problem is made
more complicated, of course, by the presence of two metal orbitals which
are A'1 , because, in addition to d,,, the metal s orbital has this symmetry.
Although by no means always justified by detailed calculations, it is often
convenient to take sum and difference of the d,, and s orbitals. This leads
to one metal A'1 s-d mixed orbital largely, if not exclusively, interacting with
the axial ligands and one with the equatorial, thereby simplifying the
problem. The metal p, orbital has A2 symmetry and so interacts uniquely
with the ligand group orbital of this symmetry.
114 1 Molecular orbital theory of transition metal complexes
may each interact with either or both of the metal orbitals of A1 ' symmetry:
sand d}
Similarly, the ligand A2 " cr orbital interacts with the metal Pz orbital
and the ligand e' orbitals interact with metal Px and Py orbitals:
with
and
Finally, there are both d orbital and p orbital sets of E' symmetry (the d
we have discussed above, the p are Px and p,). Again, the general outcome
is not clear. Nonetheless, we can make an educated guess. Although there
is no required relationship, we can actually relate the trigonal bipyramidal
problem to that of the octahedron, discussed earlier. If, for each, we separate
out two axial ligands, then in the trigonal bipyramidal they are partnered
by three equivalent coplanar ligands, in the octahedron by four. Perhaps the
outcome in the two cases is not too dissimilar. If this is so, it would be
concluded that the a'1 (dz,) and e' (dx'-y' and dx,) orbitals are actually <J
antibonding. The metal d orbitals of E" symmetry, d,z and d"" are
non-bonding. When n bonding is included, the ligand orbitals a2, a~, 2e' and
2e" are added to the problem (Table 6.6). They will not be discussed in detail;
we merely note that the a~ orbitals are only involved in <J bonding, the e"
only in n bonding and the e' in both <J and n bonding. The appearance of
two orbitals of a symmetry species (2a;, 2e;, 2e") in the above discussion
need cause no concern. They arise because the axial and equatorial ligands
are not symmetry related and so contribute additively. The way they were
handled for the 2a; case illustrates the general approach-include them all
together, the actual number of them is not important.
HH t1u (1)
of two electrons occurs for the former and one for the latter (we choose
integers for simplicity and ignore the rather difficult problem of how to
calculate the number of electrons transferred). The resultant charges on the
two cations are therefore both + 1 (( +3- 2) and ( +2- 1)) although
we started with one Fe3+ and one Fe 2 + complex. The charges we assumed,
3+ and 2+, are free-ion charges. One expects that the magnitude of an
actual charge will always be less than the absolute magnitude of a free-ion
charge, for both ligands and cations. This, of course, is a restatement of
Pauling's electroneutrality principle. Contemporary calculations indicate
that whilst the actual charge of a Fe 3 + ion is likely to be rather greater than
that on a FeZ+ ion if the ligands are identical, the difference between them
is only of the order of one-third of an electron. The actual charges themselves
would be of the order of unity (positive). It is not surprising, therefore, that
the use of free-ion charges and valence states, will sometimes prove difficult.
In a transition metal complex containing H as a ligand, should this
be regarded asH+ or H-? The charge on the metal depends on which we
choose. In practice, the problem could either be sidestepped or worked the
other way round-a formal charge first assigned to the metal from which
the, equally formal, charge to be allocated to the H ligand would be deduced.
Alternatively, the problem could be resolved by appeal to chemistry-does
the H ligand behave more like H+ or H- in its reactions? In transition metal
chemistry the answer is quite often H- because transition metal hydrides
are more characteristically reducing agents than acids. Although the charges
indicated by the symbols FeZ+ and Fe3+ are misleading, this representation
is far from valueless. In particular, these charges lead to a correct count of
the number of electrons in the t 29 and upper e9 orbitals such as in Fig. 6.11.
It is essential to get this number right if we are to be able to correctly interpret
the physical and chemical properties of a complex. The usual compromise
adopted is to refer to iron(III) and iron(II), as has usually been done
throughout this book, rather than Fe3+ and Fe 2 +, thereby avoiding the
difficult problem of the actual charge distribution within the molecule. At
some points in the text, usually for emphasis or to facilitate electron counting,
there has been a reversion to the Fe3+ and Fe 2 + convention.
It is common practice to assign a charge to species such as [Fe(H 2 0) 6 ] 2 +
This assumes that there is no covalent interaction between the complex ion
Experimental 1117
6.6 Experimental
The content of this chapter has been theoretical and so an experimental
section seems somewhat out of place. Yet it is not, for experimental evidence
from a variety of sources is becoming available which helps pin-point the
strengths and weaknesses of the discussion. The focus is on X-ray and
neutron diffraction data. Although in X-ray crystallography it is usual to
assume that atoms are spherical (so that it is assumed that on each atom in
a crystal there is a spherical distribution of electron density and it is this
which is responsible for diffracting X-rays), the precision of current apparatus
and techniques, at their best, is such that deviations from spherical can be
measured. In Fig. 6.30 are shown these deviations for the cobalt ion and
four coplanar N02 ligands in the complex ion [Co(N02 ) 6 ] 3 -. In this
complex the cobalt(III) has a t~.e~ configuration and the consequences
of this are evident in Fig. 6.30--there is a depletion of electron density in
the 'e• region', i.e. along metal-ligand bonds, and buildup in the 't 2 g regions',
i.e. between metal-ligand bonds, just as expected. The reader may also note
a buildup of electron density near the N atom of the Co-N bonds; even the
N-0 bonding electrons are visible. This result is typical of X-ray electron
density difference measurements. They support the general picture presented
both in this chapter and in Chapter 7.
A1 1 1 1 -1 -1 -1
A2 1 1 -1 -1 -1 1
E" 2 -1 0 -2 1 0
Crystal field theory of
transition metal complexes
7.1 Introduction
Although little use is made now of the theory presented in this chapter,
it contains the basis of all of those that are used. It provides the foundation ,
particularly for the understanding of spectral and magnetic properties; all
else is elaboration and refinement. A knowledge of simple crystal field theory
is therefore essential to an understanding of the key properties of transition
metal complexes and particularly those covered in Chapters 8 and 9. This
chapter deals exclusively with transition metal complexes. In one or more
of their valence states, the ions of transition metals have their d orbitals
incompletely filled with electrons. As a result, their complexes have character-
istics not shared by complexes of the main group elements. It is the details
of the description of these incompletely filled shells which is our present
concern; this is in contrast to the discussion of the previous chapter where
the topic was scarcely addressed. Ions of the lanthanides and actinides
elements have incompletely filled f orbitals and so necessitate a separate
discussion which will be given in Chapter 11.
In 1929 Bethe published a paper in which he considered the effect of
taking an isolated cation, such as Na +, and placing it in the lattice
of an ionic crystal, such as NaCI. In particular, he was interested in what
happens to the energy levels of the free ion when it is placed in the
electrostatic field existing within the crystal, the so-called crystal field. The
energy levels of a free ion show a considerable degeneracy, particularly if
one is prepared to ignore effects which cause only small splittings. That is,
in the free ion there exist sets of wavefunctions, each member of any set being
quite independent of all other wavefunctions (i.e. orthogonal to them), yet
all members of any one set correspond to the same energy. What happens
to these ions when placed in an ionic crystal? Do the wavefunctions which
in the free ion had the same energy still all have the same energy in the
122 1 Crystal field theory of transition metal complexes
crystal? Bethe showed that in some cases the free ion degeneracy is retained
and in others it is lost, the crucial factors being the geometry of the crystalline
environment and the term ('S, 3 P, 2 D, 1F etc.) of the wavefunctions of
the free ion.
Two years later, in 1931, Garrick demonstrated that a simple ionic model
gives heats of formation for transition metal complexes which are in
remarkably good agreement with the experimental values. That is, these
complexes behave as if the bonding between the central metal ion and the
surrounding ligands is purely electrostatic, just as in a simple picture of the
bonding in NaCI. If this is so, then Bethe's work may be applied to complexes
as well as to ionic crystals and the energy levels of the central metal ion
related to those of the same ion in the gaseous state. All that is needed is a
suitable quantitative calculation to obtain the energy level splittings due to
the crystal field. This approach to the electronic structure of transition metal
complexes is known as crystal field theory and it is the subject of the present
chapter.
1 The present author has written a, hopefully, easy to read and follow, non-mathematical
text on group theory called Symmetry and Structure, 2nd edn., published by Wiley, Chichester,
1995.
Crystal field splittings 1 123
The difference between these names is rather subtle and different usages are
met. The most common is that in which the detailed mathematical specifi-
cation decreases in the order state > level > term. So, in most of our
discussion we will talk of terms, although the repeated use of this word may
give rise to some ugly language and although in some papers and texts the
word state is used almost interchangeably with it. After a phenomenon
known as spin-orbit coupling has been included we will talk of levels and
if we wish to be even more detailed and talk about an individual wave-
function, we will use the name state. Care must be taken to avoid confusing
this use of state with phrases such as 'the ground state' and 'an excited state'.
For our present purposes the important thing to note is that under the
rotational operations of an octahedron2 any S term has the same symmetry
as an s orbital, a P term the same symmetry properties as a set of three p
orbitals and so on. Key, however, is the fact that whilst a solitary electron
in a set of p orbitals gives a P term, it transpires that two electrons in the
d orbitals can give one also, as can three d electrons. We talk of 'the P term
arising from the p 1 configuration', or 'the P term arising from the d 2
configuration' and so on. A crucial factor turns out to be the number of
unpaired electrons, n, associated with each term and this is indicated by the
number (n + 1) as a superscript thus: 2 P, 3 D and so on. If the choice of
(n +I) seems odd-perhaps n seems more sensible-reflect on the fact when
n = 1, that is, when there is just one unpaired electron, this electron can
have spin up or down---there are two, (n + 1), different spin possibilities.
Although it is usual to arrive at symbols such as 3 P by feeding electrons into
orbitals, the absence of any specific reference to these orbitals in the final
symbol suggests that they are merely a convenient vehicle, and not essential.
This is the case. Symbols such asP and 3 P, like their lower case counterparts,
are a consequence of the (rotational) symmetry of a sphere. Indeed, the
symbols S, P, D, etc. are the labels of irreducible representations of the
relevant spherical group.
Although an octahedron has a much lower symmetry than a sphere it
would be reasonable to expect that many-electron wavefunctions would be
handled similarly. This is so-symbols such as 3 T 19 , 2 E 9 and 1 A 19 , like t 1 g,
e9 and a 19 orbitals, imply, respectively, triple, double and single orbital
degeneracy. In each case they are associated with a spin degeneracy which,
in each of these three examples, is identical to the spatial degeneracy.
However the two vary independently and so symbols such as 2 T 1 ", 3 £ 9 and
19 are perfectly reasonable.
6A
to the content of this chapter, for they are largely responsible for the energy
level splittings which were the subject of Bethe's paper. As we will see, group
theory tells us to expect these splittings (this is what Bethe showed). However,
group theory tells us nothing about the magnitude or even the sign of the
splittings. We need a specific model to do this and the simplest is the crystal
• field model. Actually, if we try to get numbers out of the model which can
then be compared with experiment we find that the model is not a very good
one. This is one aspect which has led to the development of the more
realistic model of Chapter 6. However, historically, virtually the entire detail
of the theory of transition metal ions was developed using the crystal field
model. The trick is never to use the model to get numbers. Rather, it is used
Fig. 7.1 An octahedral complex. In this chapter to focus our attention on energy differences and the relationships between
the way that an octahedral complex is drawn them. Experimental data are then used to obtain the numbers! It is this,
will vary, the perspective adopted depending on together with the fact that (as we shall see) the results are largely symmetry-
the point under discussion. Frequently, lines will
be drawn which represent the edges of the determined that lead to the utilization of the method. Consider the octahedral
octahedron rather than chemical bonds. complex shown in Fig. 7.1. What will be the effect of the crystal field on a
singles electron of the central metal ion (Fig. 7.2)? The ligand-metal electron
z repulsion which we have associated with the crystal field will raise the energy
of the s electron (or an S term), but as there is no orbital degeneracy, no
orbital splitting can result. Next, what will be the effect of the crystal field
on a single p electron (or a P term) of the metal ion? As is evident from
Fig. 7.3 all the p orbitals are equally affected by the crystal field and so, no
matter which of them the p electron occupies, the repulsion is the same. That
is, the p orbitals (or the components of a P term) remain triply degenerate
in an octahedral crystalline field.
The case of a single d electron (or a D term) is both more difficult and
more interesting. All five d orbitals are not spatially equivalent. Three, dxy•
~ dyz and dzx• are evidently equivalent for they are equivalently situated with
Y respect to the ligands (Fig. 7.4) and may be interchanged by simply
interchanging the labelling of the Cartesian coordinate axes. The other two
Fig. 7.2 A metal s orbital in an octahedral d orbitals, dx'-y' and dz'• are not equivalent, although they both have their
complex.
maximum amplitudes along the Cartesian coordinates axes (Fig. 7.5). Inter-
change of the labels associated with each of these axes has the effect not of
interchanging the orbitals but of generating new orbitals. So, starting with
the coordinate system of Fig. 7.5 the interchange x -> z -> y -> x gives us
z z z
t t
z
z z
J.
/ ~/
X X
dxy dyz
Fig. 7.4 The d,, d, and d~ metal orbitals in an octahedral crystal field. Because they are all equivalent (they can be interconverted by rotating
around the threefold axis approximately perpendicular to the plane of the paper) they remain triply degenerate.
z z
J.
X X
'\ <'
y
then the orbitals dx" and dy"-z" are obtained. It
may help to see this if both of the octahedra
drawn are mentally rotated by 120= anticlockwise
so that the axes return to the positions they
have in Fig. 7.5. Since nothing else has been
added or removed, the orbitals dx2 and dy2-z2
must be mixtures of the original dxz and dy2-x2· y ,;:/
Fig. 7.6 in which these same d orbitals are now labelled dx' and dy'-z" We
have not invented something new; the 'new' orbitals are simply mixtures of
the 'old'. The fact that it is possible to mix two orbitals by such a trivial
operation as relabelling the axes shows that the two orbitals arc degenerate.
If they were not, the mixed orbitals would have different energies from those
of the starting pair, and it is obviously ridiculous that the energies should
be a function of the labelling of the axis system. In Appendix 4 it is shown
in a more formal way that the new d orbitals are simply mixtures of the old
ones, dz' and dx'-y'· We conclude that the d orbitals (or aD term) split into
126 1 Crystal field theory of transition metal complexes
z z
X___-
X___-
fxyz
two sets, a set of three degenerate orbitals and a set of two degenerate
orbitals. The relative energies of these two sets will be discussed shortly.
Finally, we consider the effect of an octahedral crystal field on a single f
electron (or an F term). Fig. 7.7 pictures the seven f orbitals in an octahedral
environment. The lobes of three, fx'' fy, and f., point along axes. Three,
fz(x'-y'J' fx(y'-z'J and fy(z'-x'J have lobes located in coordinate planes (for the
fz<x'-y'J orbital shown in Fig. 7.7 these are the zx and yz planes). The last,
fxyzo has lobes pointing between all coordinate axes. Clearly in an octahedral
crystal field the f orbital sevenfold degeneracy is lost to give two sets of triply
degenerate orbitals and one singly degenerate orbital.
So far it has been shown that sets of d and f orbitals (and therefore D
and F terms) split into subsets in an octahedral crystal field but nothing has
been said about the relative energies of these subsets. For the moment, the
discussion will be restricted to orbitals because it is easy to give pictures
of them. In subsequent sections the discussion will be extended to the
corresponding terms. In preparation for this extension, it would be helpful
if the reader has some idea of their derivation so that he or she is fully aware,
for example, that an F state means seven spatial (as opposed to spin)
functions, just like a set off orbitals. One of the simplest ways of appreciating
this is through the Russell-Saunders coupling scheme, that which is adopted
to obtain the explicit functions themselves. This scheme is outlined in
Crystal field splittings 1 127
F=== dx2-y.o. ct 22
;-----fxyz
(a)
(b)
properties of ions containing unpaired f electrons. The situation is more complicated, as will
become evident in Chapter 11.
128 1 Crystal field theory of transition metal complexes
+I ++ ++
} Smallll.
I +
+
II
d.
+I
+
ds
++I
d6
N
d7
High spin
RJI. 7.10 The high spin (small Ll.) and low spin
(large Ll.) possibilities for d4 -d 7 octahedral
complexes.
Low spin
A similar series exists for the variation with metal ion in which it is seen,
as already mentioned, that !l. increases with the formal charge on the ion
and down the periodic table:
Mn 11 < Ni 11 < Co" < Fe 11 < V11 < Fe 111 < Cr 11 < V111 < Co"' < Mn 1v < Rh 111
< Pd 1v < lr111 < Pt 1v
Despite the above argument, the former series should not be regarded as a
series in which the charge on the ligand increases from left to right. Other
factors are involved, as was evident from the discussion in Chapter 6.
4 It commonly appears green, but this is caused by contammation with a small amount of
Table 7.2 Crystal field components of the ground and some excited terms of dn
(n = 1-9) configurations
or u functions. In this chapter our concern is only with the crystal field
splitting of terms derived from d" configurations and because all d orbitals
are centrosymmetric, we shall encounter only g suffixes.
Table 7.2 gives the behaviour for all of the transition metal ions (d 1-d 9
configurations); it lists ground terms and, where the information is needed
later, the behaviour of the lowest excited state. Two points should be noted
in connection with this table. First, that only S, P, D, and F terms occur;
for each the splitting is similar to that of the corresponding orbitals. Secondly,
the table has some symmetry. Apart from the first column, the bottom half
is the mirror image of the top half. These two features combine to simplify
the remainder of our discussion of weak field complexes. The ground and
excited terms of transition metal ions have been introduced in Table 7.2
without any justification apart from that in Appendix 5. Such a justification
will be needed for the f electron systems in Chapter 11. The reader who
is unhappy with the present ex cathedra presentation should read Section
11.6 where the procedure is detailed; they may be fortified by the knowledge
that d electron systems are easier than their f electron counterparts! The
crystal field splittings given in Table 7.2 follow from the discussion earlier
in this chapter.
We turn now to the problem of the relative energies of the crystal-field
components listed in Table 7.2. Consider the D terms, which give T 2 • and
E 9 components. Which component is the more stable and by how much? It
will be taken for granted that it is an experimental fact that the five d orbitals
split as shown in Fig. 7.9 the splitting being denoted by .1.. Consider the
d 1 case, Fig. 7.11. Here, the ground state will be that in which the electron
occupies the lowest, t 2 ., orbitals. Just as a d 1 configuration gives rise to a
2 D term so, too,~~. configuration (which is what we have here) gives rise to
e:
a 2 T 29 term. Similarly, the (excited) configuration gives rise to a 'E• term.
We conclude that, for the d 1 case, the 2T 29 term is the more stable because
it means that the solitary d electron is in the t 29 orbitals. The (excited)
t2g
2 E. term is generated from it by excitation of an electron from the t 2 • orbitals
(a) (b)
to the e• orbitals. This, by definition, requires an energy .1., so we conclude
that the 2 T 2 • and 2 E. terms are also separated by the energy .1..
Fig. 7.11 (a) Ground and (b) excited states
derived from the d1 configuration in an Two points should be noted. First, a detailed argument is required to
octahedral crystal field. relate the splitting of orbital energies to the splitting of term energies.
Weak field complexes 1 133
eg
Fig. 7.18 (a) Ground and (b, c) two excited
state configurations of a d3
ion in an octahedral
crystal field. Note that there is only one
distinguishable way of arranging the electrons in
t2g
the three orbitals in the ground state-and so ~-
the ground state is orbitally non-degenerate.
(a) (b) (c)
sfg
•r,g
6s 6A1g so 4F 3F
3T2g
sr,g
4hg
d5 d6 Ll d7 Ll
3A2g
Weak Strong
field field
limit limit
ll=O Ll==
By the end of the discussion of weak field complexes the effects of both
the (weak) crystal field and electron repulsion had been included. To be
consistent we must consider the effects of electron repulsion in strong field
complexes, and this we now do. Electron repulsion causes some arrangements
of electrons within an unfilled shell to be more stable than others-those
arrangements in which unpaired electrons are kept farthest apart and have
parallel spins will be the most stable (Hund's rules). That is, in a free
atom or ion, electron repulsion causes the terms arising from a configuration
to have different energies, as we have seen. Similarly, in a crystal field, the
terms arising from a configuration like t~" will, in general, have different
energies because of electron repulsion. How does one determine the terms
arising from such a configuration? What are the relative energies of these
terms? We shall answer these questions by looking at the group theory of the
problem. Results that are qualitatively correct will be obtained, with no need
to evaluate a single integral.
Table 7.4 is a table of direct products; its derivation is included in Appendix
3. This table is important, for it is used whenever one is simultaneously
interested in two similar quantities associated with an octahedral molecule;
for example, if we are interested in the symmetry properties of two electrons
as a pair rather than as individuals. Similarly, this table will be used to
discuss spectra, for which the ground and excited states of a molecule have
to be considered simultaneously. Table 7.4 does not give g and u suffixes;
they could have been included, but this would have made the table four
times larger with no increase in real content. The way that these suffixes may
Strong field complexes 1 139
oh Al A2 E T1 Tl
Al Al A2 E Tl Tl
A2 A2 Al E Tl Tl
E E E A1 +A2 + E Tl + Tl T1 + Tl
Tl Tl Tl rl + r2 A1 + E + T1 + T2 A2 +E+T1 +T2
Tl Tl Tl Tl + Tl A2 + E + T1 + T2 ~+E+T1+T2
be added will be shown shortly but first the meaning of the entries in Table
7.4 must be explained.
As our immediate aim is to complete Fig. 7.20 our discussion will be
confined to the d 2 case. Consider first the ti.e~ configuration.The first
electron may be fed into any one of three r2 " orbitals and the second into
any one of two e,. That is, there are 3 x 2 = 6 ways of feeding the two
electrons in; there are six orbitally-different wavefunctions. Table 7.4 shows
that the direct product of t 2 (extreme left-hand column) with e (top row)
written T 2 x E, is equal to T 1 + T 2 or, including g suffixes in an obviously
sensible way, T 29 x E, = T 19 + T 29 (note that T 2 , x Eu or T 2 u x E9 would
have given T 1u + T 2 uJ. The other possibility, T 2 u x E,, would also have
given T,. + T 2 , (g x g = u x u = g; u x g = g x u = u). The sum of the
degeneracies implied in T 19 and T 29 (3 + 3 = 6) is the same as the number
of orbital wavefunctions arising from the ti,e~ configuration (3 x 2 = 6). It
will not surprise the reader to learn that these six ti,e~ two-electron
wavefunctions divide into two sets of three each, one set of T 19 symmetry
and one set of T 29 symmetry. That is, the configuration tie; gives rise to T 19
and T 2 , terms. This is no accident-it is a group theoretical requirement and
would be just as valid in a different context (molecular vibrations, for
instance, if we simultaneously excite T 29 and E, molecular vibrations of an
octahedral complex the molecule could end up in either a-vibrational-T 1 "
or T 2 " state). So far, the spin of the electrons has not been mentioned. Because
in the tLe; configuration the two electrons always occupy different orbitals,
there are no constraints and all paired (singlet) and parallel (triplet) spin
arrangements are compatible with all the orbital symmetries that arise. We
conclude that the tLe; configuration gives rise to 3 T 1,, 3 T 2 ,, 1 T 1" and 1 T2 ,
terms. This means that the line at the strong field limit in Fig. 7.20 with
slope ! D. is a superposition of lines corresponding to these four terms.
However, because Fig. 7.20 is only concerned with triplet spin terms, only
the labels 3T 19 and 3 T 29 have to be added to this line in Fig. 7.20.
As a check on the answer that has been obtained it is helpful to count
wavefunctions. Previously, we only counted orbital functions; what if we
include spin° Now, there are six different ways of putting an electron into
the t 2 , orbitals (three orbitals, two spins) and four ways of putting one
into the e., a total of 24 different ways of putting in the two electrons;
that is, there are 24 different wavefunctions. This is just the number implied
by 'T,, + 3T 2 " + 'T," + 1T 2 , (9 + 9 + 3 + 3). The counts agree, as they
have to.
We now consider the e~ configuration, associated with the strong field
line of slope %D. in Fig. 7.20. The first electron can be fed into the
140 1 Crystal field theory of transition metal complexes
e, orbitals in any one of four ways (two orbitals, two spins). The second
cannot be in the same orbital with the same spin as the first and so the
number of distinct two-electron wavefunctions is 1(4 x 3) = 6. The 1 arises
from the word distinct: for instance, the 4 x 3 includes both (j l) and (L j)
yet these are not distinct; we have counted everything twice. As Table 7.4
shows, the direct product E, x E 9 = A 19 + A 29 + E,. We know that the e;
configuration must give rise to at least one spin triplet (because we can put
one electron into each e, orbital and they can be put in with parallel spin).
But, as Fig. 7.20 shows, in the weak field limit there is no 3 A 1 , or 3 E, term,
only 3 A 29 • We conclude that the e; configuration gives rise to the terms
19 + A 2 , + E 9 • This conclusion can be checked by counting the number
1A 3 1
field limit but has to correlate with the 3 T 19 term arising from the t~ 9 e:
configuration, a term which has an energy of! .1. (this has to correlate because
if it did not the non-crossing rule-the rule that the energy levels of terms
of the same symmetry and spin degeneracy do not cross-would be violated).
So, the term 3 T 19 (P) changes its dependence on .1. as .1. itself changes-and
this is caused by the presence of the second 3 T 1• term, with which it
interacts-and it is this latter that we are interested in. The moral is: be
careful when there is more than one term of a given symmetry and spin
multiplicity. Actually, it is not difficult to calculate the energy of the 3T 19(F)
term. Note that it is the splitting of the 3 F weak field term that gives
3 A 29 + 3 T 1 • + 3 T2 • terms. The word splitting implies that the energies of the
components sum to that of the 3 F term; their .1. dependencies sum to zero.
Denoting the energy of the 3 T,.(F) term by S(T,.), taking care to weight each
energy by the number of wavefunctions with this energy, we have
(3 X ~ .1.) + (9 X 8(T1g) + (9 X g.1.) = 0
That is, S(T1•) = -~ .1.; our caution was justified-in the weak field limit the
3 T 1.(F) term does not have the same energy as in the strong field limit. The
two energies differ by -! .1., equal and magnitude but opposite in sign to
the difference between the energies of the 3 T,.(P) term in the same limits.
The crystal field has caused the two 3 T 1• terms to interact; they 'push' each
other apart by equal and opposite amounts. Figure 7.20 can now be
completed. The final version is shown in Fig. 7.21 which embodies all the
content of the above discussion. One final point. As befits the above
discussion, in this figure straight lines are drawn representing 3 A 2 • and 3 T 29
terms connecting the weak and strong field limits. Given what was said
earlier about the non-linear scale of the .1. axis in this figure, such straight
lines cannot be justified. The moral is clear: Fig. 7.21, and related ones which
will be presented shortly, are of pedagogical use only. However, they serve as
an excellent introduction to related diagrams (Tanabe-Sugano diagrams)
which have well-defined energy scales.
The discussion so far almost, but not quite, enables an extension to all
d" systems either by the electron-hole parallel, by neglect of half-filled shells,
or both. The omissions can be dealt with by arguments paralleling that which
follows for the d. configuration, a configuration which poses a problem.
The most obvious way of obtaining the orbital terms arising from this
configuration is to simply take each of the orbital terms arising from the d.
configuration and combine each with a further t 29 orbital function. That is,
to form the triple direct product T 2 • x T 2 • x T 2 • or, equivalently, consider
the sum of direct products
(Alg X T2g) + (Eg X T2g) + (Tlg X T2g) + (T2g X T2g)
This would be wrong. Some of the spin singlet wavefunctions of the T~.
configuration represent two electrons occupying the same orbital. To form
direct products blindly would, for some of the three-electron wavefunctions,
be to allocate all three electrons to one orbital! The simplest way to avoid
this problem and obtain the correct answer is as follows. In the 3 T 19 term
of the tL
configuration we know that the electrons have parallel spins and
so must occupy different orbitals (because of the Pauli exclusion principle).
Adding a third t 2 • electron can never give us three electrons in one orbital.
142 1 Crystal field theory of transition metal complexes
Weak Strong
field field
limit limit
Figure 7.22 gives diagrammatic representations of the 3 T 1, orbital wave-
functions (spin is not specified). Also in Fig. 7.22 are shown all the possible
ways of orbitally allocating electrons in tL configuration, again without
specifying spin. It is easy to see that all of the tL arrangements may be
obtained from those of the orbital components of the 3 T 19 term. This suggests
that the direct product T 19 x T 2 , will give us the symmetries of all the sets
of d, three-electron wavefunctions. This is so; from Table 7.4 it is seen that
they are A 29 , E,, T 19 and T 29 • The spin multiplicities of these terms now have
to be added. It is easy to sec that only one spin quartet term can exist and that
this is orbitally singly degenerate~ there is only one way of allocating three
electrons with a spin to the three t 29 orbitals. The other terms must therefore
be doublets; that is, we have 4 A2 ,, 2 E,, 2T 1 , and 2T 29 . Again, the total
degeneracy (4 + 4 + 6 + 6 = 20) equals the number of distinguishable and
allowed ways of feeding three electrons into the t 29 orbitals:
(b)
t~geg tige: 4T
41 , 4 T2 g, 2 A41 , 2A2g, 2 2 E,, 2 2 T41 , 2 2 T2 g
t 2 ge~ t~ge~ 4T 2 2 T2 g
41 , 2 2 T1,,
e3g t~geg 'Eg
3T
4, 6 tig t~ge; 41 , A41 , 'E,, T41
1 1
In the preceding paragraphs, our concern has largely been with the
d 2 case. The results obtained for this configuration may readily be extended
to other cases; this extension is given in the next section. The present section
is concluded by enquiring into the d orbital occupancy when a weak field
d 2 ion is in its ground state (the d 7 case is similar). The energy of the 3 T 19
ground term, -~ L\, must correspond to an electron distribution of ~
electrons in the t 29 orbitals and 1 in the e, orbitals:
(g X -~ M + (t X g!l) ~ -~"'
Configuration Terms•
d1 , d 9 20
d2, da 3F, 3p, 1G, 'D, 15
d3, d7 •F, •p, 2H, 2G, 2p, 220, 2p
d4 , d6 5 0, 3 H, 3 G, 2 3 F, 3 D, 2 3 P, 1 /, 2 1 G, 1 F, 2 1 0, 2 1 5
d5 65, •a, 4F, •o, 4P, 2/, 2H, 2 2G, 2 2F, 3 2D, 2P, 25
a 2 20 means that there are two distinct 2D terms.
Intermediate field complexes 1 145
Octahedral Tetrahedral
Fig. 7.23 A modified Orgel diagram for weak
field complexes with free ion 0 terms as ground
state (d\ d4 , d6 and d9 ), I
Tetrahedral Octahedral
I I
e1 (d 1 ) } t2} eg' (d 4 )}
e3 ti(d 6 ) t2~ el (d 9 )
Stro'-n-g----,..--Z-e-ro--,..-----S_.Jtrong
field field field
limit limit
146 I Crystal field theory of transition metal complexes
Tetrahedral Octahedral
e4 ti (d 7 )} tJ (d 3 ) }
e2 (d 2 ) t2~ ef (d8 ) Octahedral Tetrahedral
L-------~--------~
Strong Zero t.- Strong
field field field
limit limit
tilt
Strong Zero Strong
field field field
limit limit
3
eg
7. 7 Non-octahedral complexes
So far, this chapter has been concerned entirely with octahedral complexes.
Whilst the majority of complexes are octahedral, almost all of them display
some slight deviation from the ideal geometry. Other complexes have quite
different geometries, as was seen in Chapter 3. For the moment the discussion
will be extended in a more limited way, so as to include only tetrahedral
and square planar complexes.
==
drawn as in Fig. 7.28 then x y z so that dxy• d,, and dzx must be
degenerate (Section 6.3 and Appendix 4). The relative ordering of these two
sets energetically may also be seen from Fig. 7.28. For example, one may
say, loosely, that each lobe of the dx, orbital is half a cube edge away from
each ligand (regarded as a point), but each lobe of the dx'-y' orbital is half
a cube diagonal away. As for the octahedral case, this argument must be
regarded as indicative only. The conclusion that the e set is less destabilized
by the ligand field than the t 2 set (Fig. 7.30)-the opposite sense of
splitting to that which occurs in an octahedral field-is confirmed by
experiment and by detailed calculations. Although the agreement between
experiment and the crystal field calculations is no better than for the
octahedral case, these calculations also show that, for a given metal and
ligand and constant metal-ligand distance, the magnitude of the splitting in
a tetrahedral complex is ~ of that in an octahedral. That is,
A,., = - ~ Aoc,
a tetrahedron. Other cases (and the octahedral also, when in the interest of clarity) carry suffixes.
150 1 Crystal field theory of transition metal complexes
Free ion
eg (dyz• dzx)
Free ion
eg
conclude that in square planar complexes dxy and dx'- ,., suffer an extra
repulsion. This is but a minor disadvantage and Fig. 7.33 is the one generally
encountered. Note that whilst b. may be regarded as the energy difference
between the centres of gravity of the upper pair and lower trio of levels, two
additional splitting energies are needed to give a complete specification of
the energy level diagram for square planar complexes. These arc the splittings
of the b 1 ,-a 1 " orbitals and of the b 29 -e 9 . There is no evident connection
between these energies because the consequence of removing the ligands on
the z axis must be expected to be different for the d=' and the dxz and dyz
152 1 Crystal field theory of transition metal complexes
Al Al
A2 81
E A1<z"> + a1<x>- >">
rl A2 +E
T2 B2 (xy) + E(zx, yz)
orbitals. Not surprisingly, theoretical efforts have been made to relate the
two splittings and thus make the problem of interpreting the d-d electronic
spectra of square planar complexes more tractable.
Note that the symmetry labels used for square planar complexes are
different from those used for octahedral complexes. The correlation between
the two sets is shown in Table 7.8. In this table are included some symmetry
labels which have not yet been used in the discussion, anticipating a need
for them in later chapters. The g and u suffixes have been omitted in both
geometries in order to keep the table compact. (The rule, of course, is g ..... g
and u --> u.) There is some freedom about the choice of the B1 and B2 labels
in square planar complexes in that they may be interchanged. What one
author called B 1 another may call B2 • In this book the choice shown in Table
7.8 is used. Table 7.8 may be used for either orbitals or terms. For example,
the 2T 2 • term of a d 1 configuration in an octahedral ligand field splits into
two, 2 B2 • and 2 E., if the ligand field is reduced to square planar. Of these,
the 2 E• is more stable because it corresponds to the single d electron
occupying the d.. and d,, orbitals (which are more stable than dx,, cf.
Fig. 7.33).
Table 7.9 Correlation between symmetry labels used for true octahedral, trigonally
distorted octahedral and digonally distorted octahedral geometries, As for Table 7.8
g and u suffixes are omitted but must be added. So, A1 (xy) and A(i') mean,
respectively, that dxy transforms as A:~g (trigonally distorted octahedron) and dz'
transforms as Ag (digonal distortion). For digonally distorted complexes the labels
Bv 8 2 and 8 3 (with or without g or u suffixes) may be used differently by different
workers, although those with g suffixes are never interchanged with those with u
suffixes
Al Al A
A2 A2 A
E E A(z 2 ) + A(x' - y")
T1 A2 + E B 1 (xy) + B2(yz) + 83(zx)
the other. Which set is used depends on the molecular geometry, just as
different d orbital sets are used for octahedral and trigonally distorted
octahedral geometries.
One final point: just as we found that the number of unpaired electrons
in an octahedral complex can vary with the crystal field, so, too, in other
geometries. However the effects tend to be more subtle and difficult to
disentangle. Thus, for five-coordinate complexes they are interwoven with
the trigonal bipyramidal and square pyramidal structural possibilities. 8
c:+7
field theory: that when a quantity is available from spectral data on free
ions-an electron repulsion energy, for instance-it should used in crystal
field calculations. Covalency means that the orbitals occupied by the d
r---'
electrons in a complex must be expected to differ from those of the free ion
(b~ -- -
and so, too, therefore should the value of electron repulsion and other
energies. In practice, this means that such quantities become parameters in
the theory, given those values which produce the best agreement with
FJg. 7.34 A digonally distorted octahedral
complex: (a) an octahedral complex and a experiment. These additional parameters will be met in the next two chapters.
distortion which is symmetric with respect to The magnitudes of the additional parameters are determined experimentally,
the c2 axis indicated; (d) the distorted molecule
with the original octahedron indicated dotted. 8 SeeR. R. Holmes, J. Amer. Chern. Soc. (1984) 106, 3745, for a detailed discussion.
154 1 Crystal field theory of transition metal complexes
(a) (c)
but the number of independent observables is usually less than the number
of parameters. It cannot. therefore, be asserted that ligand field theory, is in
general, proven. One may say only that it provides a consistent explanation
for experimental data using parameters which, almost invariably, have
physically reasonable values. Not surprisingly, there have been attempts to
devise cunning experiments which provide additional information so that we
end up with more data than parameters. For instance, instead of studying
complexes randomly orientated in solution one might look at molecules fixed
in a single crystal of known structure. The orientation of the crystal, and
with it any inherent molecular anisotropy (such as a tetragonal distortion),
could be varied relative to the direction of polarization of polarized incident
light, for example, and so give additional data. However, octahedral and
tetrahedral complexes are isotropic (x = y = z) so this method is only
applicable to lower symmetry molecules requiring even more parameters to
start with, as was seen in Section 6.9. Suitable cases for study are already
bad ones! Again, instead of working at one temperature, we could study a
property- the magnetic properties of a complex for instance-as a function
of temperature. The varying thermal population of low-lying energy levels (a
magnetic field frequently induces suitable small splittings, as we shall see in
Chapter 9) then gives information on the energy separation between these
levels. Unfortunately, at the very low temperatures that really lead to the
depopulation of levels that are only slightly split apart, and so offer the
greatest potential for additional useful information, other weak phenomena-
such as the freezing out of lattice vibrations-have to be considered and this
requires yet more parameters! We seem to be hitting a brick wall at every
turn. This situation has inspired a search for theoretical models which might
reveal interrelationships between the various parameters of ligand field
theory whilst themselves involving relatively few, if any, additional parameters.
Many such models have been investigated, including those that we will
meet in Chapter 10. Entire books have been written on the subject. 9 Suffice
to say that the search continues. Although clear progress has been made, no
single model is currently accepted as providing a generally applicable
method.
We conclude by reconsidering what we must now call the ligand field
splitting parameter of an octahedral transition metal complex, Ll, since the
factors affecting its magnitude are evidently more complicated than we at
first supposed. We have encountered three such factors:
9 See, for example, M. Gerlach and R. S. Slade Ligand-Field Parameters, Cambridge
University Press, Cambridge, 1973.
Questions 1 155
This does not exhaust the list of factors influencing the magnitude of t1, but
there are believed to be no others of comparable importance. A recitation
of most of the evidence that supports the ligand field model in preference to
the crystal field model is deferred until Section 12.1.
Further reading 1970 and, more simply, in Chapter 12 of The Chemical Bond
by the same authors, J. Wiley, Chichester, 1985.
In practice it is not possible to separate the further reading An excellent introduction is given in 'Ligand Field Theory'
relevant to this chapter from that appropriate to Chapter 6. by J. S. Griffiths and L. E. Orgel, Quarterly Rev., Chern. Soc.
The contents of the two chapters go together. In addition to (London) (1958) 11, 381. This particular article was perhaps
the references given at the end of Chapter 6, two others which more responsible than any other for the introduction of the
follow the pattern of the present chapter but with a more ideas of crystal and ligand field theories into inorganic chem-
mathematical approach are Chapter 13 of Valence Theory by istry.
J. N. Murrell, S. F. A. Kettle and J. Tedder, Wiley, London,
8.1 Introduction
In this chapter our concern will be with the electronic spectra of transition
metal complexes, particularly those of the first transition series. Although
nowadays there are fewer studies of these spectra, per se, than previously,
their study is essential if the electronic excited states of complexes are to be
understood. The energy required for the promotion of an electron from one
orbital to another or, more precisely, the excitation of a molecule from its
electronic ground state to an electronic excited state, corresponds to
absorption of light in the near-infrared, visible or ultraviolet regions of the
spectrum. For transition metal complexes the absorption bands in the first
two of these regions are relatively weak and are associated with transitions
largely localized on the metal atom. The ultraviolet bands are intense. They
are associated with the transfer of an electron from one atom to another and
so are called charge-transfer bands. These bands are responsible for the
colour changes associated with indicators for inorganic cations, such as the
thiocyanate test for Fe 111 or the indicators used in EDTA (compleximetric)
titrations, but in these cases the charge-transfer bands fall in the visible region
of the spectrum. The intense bands will be dealt with in the latter part of
this chapter; for the moment our concern will be with the weaker bands.
Whilst these are most simply explained by crystal field theory, a detailed
comparison of the data with the theoretical predictions will show that ligand
field theory provides a more appropriate explanation.
When spectral bands are weak there is a reason. Often it means that they
are bands which are forbidden but- obviously- not totally forbidden. The
weak bands with which we are concerned are of this type. Electronic
transitions which correspond to strong bands are electric dipole in type.
Classically, the electric vector associated with the incident light beam behaves
like a pair of alternating + and - charges across the molecule. These
oscillating charges induce an oscillating dipole in the molecule; when the
Electronic spectra of V 111 and Ni 11 complexes 1 157
all that is needed to interpret the spectra of octahedral V111 complexes. 15B
is the separation of the 3 F and 3 P free ion terms, so it should be possible
to obtain the value of B from atomic spectral data. This corresponds to the
pure crystal field approach, but, significantly, agreement between experiment
and theory for complexes is only generally obtained if B is treated as a
parameter, the value of which may be varied. It turns out that B has to be
given a value which is rather smaller than that obtained from atomic spectra.
This parameterization of B is part of the ligand field method and is consistent
with our ideas of the consequences of covalency in transition metal complexes.
Covalency implies
1. that the metal electrons will be partially delocalized onto the ligands;
2. that the effective positive charge on the transition metal will be smaller
than in the free ion.
Both of these effects mean that the metal electron cloud will be more diffuse
in the complex than in the free ion, and repulsion between the electrons
making up this cloud is therefore reduced. This conclusion is confirmed by
a more detailed analysis.
The energies of the crystal field spin triplet terms of the d 2 configuration,
relative to the 3 F free-ion term as zero, are given explicitly below. It is beyond
the scope of the present text to derive them so we give these expressions
without proof;2 however, the reader may readily check that they reduce to
those given in Section 6.5 for the weak field limit 3 !1--+ 0 (expand the square
root, using the binomial theorem in the form (y 2 + xy) 112 ::::: y + x/2). These
expressions also give the strong field energies; note that the 15B term was
not included in the energies given in Section 6.5, not that this is important
because in the strong field limit the effects of electron repulsion are negligible
in comparison with the ligand field. In Chapter 7 this situation was
accommodated by making the ligand field infinite but in the present context
it is simpler to set B = 0. The square root expressions are those which allow
for the mixing of the 3 T 1.(F) and 3 T 1,(P) terms by the crystal field. Indeed,
it should be evident from the form of these two energy level expressions that
they arc obtained as the roots of a quadratic equation. The energies of the
3T
19 and A 29 terms are those which were obtained in Chapter 7.
3
Term Energy
and 3 T 1.(P) <-- 3 T 1.(F) transitions, have energies given by the differences
between the two relevant expressions above:
70
Transition Energy
60
3 T2g +-- 3T,fF) ~[!'. - 158 + (2258 2 + 1888 + 82)112]
15 3T11(P) +-- 3T11(F) (2258 2 + 188 8 + 8 2)112
50
3T2g
We have experimental transition energies to fit these expressions, so we can
L..--""7- 1A1B
now obtain values for A and B. This task is not difficult. Divide the two
expressions above by B. We then have
30
E( 3 T2B +-- 3 T:,fF))
1G -1 [A
- - 15 + ( 225 + 18-
A + -A2)112] (8.1)
20 8 2 8 8 82
E( 3 T11(P) +-- 3 T1g(F)) A A2)112
( 225 + 188+ 82 (8.2)
8
The right-hand side expressions are functions of A/B. Similarly, the energies
of the terms themselves, divided by B, may be expressed as functions of
10 20 30
!'./8
A/B. Tanabe and Sugano have published diagrams which show these
relationships; they take as their energy zero the energy of the ground
FJC. 8.3 A complete Tanabe-Sugano diagram
for octahedral d2 complexes. state, so that, for the d 2 case, they effectively plot the functions
and
0.8
:=:i~
~~~ 0.6
~it
Fig. 8.4 A plot of eqn 8.3, useful in the
interpretation of the spectra of octahedral d2 ~;.:1:. 0.4
~~~
complexes.
""'"
Cui~ 0.2
ar
0
In the case of [V(H 2 0) 6 ] 3 + the value of this quotient is 17 200/25 600 = 0.67.
We can now proceed in one of two ways-and a third will immediately occur
to the computer enthusiast; it is not difficult to write a suitable program to
do the work. We could apply a trial and error process to Fig. 8.3 until we
find the correct value of !3./B (in our case tJ.jB = 28, shown dotted in Fig.
8.3). Alternatively, we could plot the right-hand side of eqn 8.3 against tJ.jB.
Such plots are useful when a large number of data have to be analysed. That
appropriate to the d 2 configuration is shown in Fig. 8.4 where, again, the
[V(H 2 0) 6 ]3+ case is indicated by dotted lines. Having determined tJ.jB from
Fig. 8.4 (again, of course, obtaining a value of 28), we return to Fig. 8.3 and
use this ratio to obtain E{'T1.(P) <-- 3 T 19(F))/B and E('T 1.(P) <-- 3 T 19 (F))/B.
Alternatively, and more accurately, we could substitute for !3./B in eqns 8.1
and 8.2. The right-hand sides of these equations are then found to have
values of 25.9 and 38.6, respectively so, using the experimental transition
energies, both give B=665cm- 1 • Since tJ./B=28 it follows that 11=
18 600 em - 1 . This analysis follows the ligand field approach, with B regarded
as a parameter.
In the crystal field approach we use the free-ion value of B, 860 em - 1 . It
follows that E('T 19 (P) <-- 3T 19 (F))/B = 17 200/860 = 20 and E('T 19 (P) <-
3T10(F))/B = 25 600/860 = 29.8. Using Fig. 8.3 these lead to values of tJ.jB
of 22.5 and 18.0 respectively; that is, tJ. values of 19400 and 15 500 em - 1 •
This internal inconsistency does not occur in ligand field theory. However,
it is only removed by introducing an additional parameter, and, consequently,
there are no additional data with which to test the theory, for we have used
two experimental quantities (transition energies) to define two parameters
tJ. and B. Only if the 3 T 29 <-- 3 T 19(F) were to be observed could we test the
theory. Using the ligand field values for tJ. and B obtained above, together
with the expression for the transition energy obtained from the term energies
given earlier, this transition is predicted to be at ca. 36 000 em- 1 , where it
is obscured by strong charge-transfer bands. However, a weak band has been
observed in this region in closely related complexes.
Electronic spectra of v"' and Ni 11 complexes 1 161
1lc
Fig. 8.5 The d-d spectrum of the -e"'
0
[Ni(H 2 0) 6 j>+ cation (d 8 ). <J)
.c
<(
60
50
40
Fig. 8.6 The part of the d8 Tanabe-Sugano
diagram believed to be relevant for the en
interpretation of the spectrum in Fig. 8.5.
"' 30
'
~--~~----~----~------''A2g
10 20 30
fl/B-
(remember, the sign of !1 has to be changed, here we are dealing with two
holes). In the crystal field model the free-ion value of B, 1082 em - 1 is
inserted, to obtain a second value of !1. This can be done by dividing
throughout by B, to obtain
14 100
- - =-
1[ t. ( t. t.
15 + 3- - 225 - 18- +- 2)1/2]
1082 2 8 8 82
and solving for !1/B, thus obtaining !1/B = 7.71. With the free-ion value of
B (the crystal field model) this gives !1 = 8340, a bit different from the
previous value. In the ligand field model we insert !1 = 8500 and solve for B
(the B 2 terms disappear) to obtain B = 900 em - 1 .
In this particular example there was a third band to consider. Really, this
should be set on a par with the other two to obtain values for !1 and B,
particularly on the ligand field model. Instead, we will use our calculated
ligand field values of !1 and B to predict its position-for this will provide
a test of the ligand field model. The energy ofthe 3 T 19 (P) +- 3 A 29 transition is
Table 8.1 Free-ion values of the electron repulsion parameter B for first row
transition metal elements with d" configurations, measured in em~ 1 . A blank
indicates that the value is not known; a dash indicates that the configuration has
either one or no electrons outside a closed shell
Metal atom charge Metal atom
Ti v Cr Mn Fe Co Ni Cu
This series has been called the nephelauxetic series (nephelauxetic = cloud
expanding). There are exceptions, but this order holds quite well. It should
be noted that what are probably the most polarizable ligands give the
lowest B values and vice versa. Another way of looking at the list is to
recognize that hard bases are to the left and that soft are to the right (see
Section 5.4). A similar list exists for metal ions but it is difficult to attach
any clear interpretation to it, although some regularities may be teased out.
Such a list, with B decreasing left to right relative to the free ion values, is
Mn 11 "' V" > Ni 11 :>: Co" > Mo111 > Cr'" > Fe111 > Rh 111 :>: lr111 > Co 111
> Mn'v > Pt'v > Ptv'
solenoid. We see that the separation of the spin and orbital properties of the
electrons is equivalent to assuming that two magnets, spin-derived and
orbital-derived, do not interact. Of course, they interact, i.e. couple, to some
extent and this is reflected in the phenomenon of spin-orbit coupling. The
formal theory of spin-orbit coupling treats the phenomenon as the coupling
of two angular momenta rather than bar magnets. We shall have to move
in this more formal direction in our discussion of the properties off electron
systems (Chapter 11), although we shall there attempt to compensate by
introducing pictorial representations of spin-orbit functions. The conse-
quences of spin-orbit coupling for d electron systems are discussed in more
detail in the next chapter. At this point it is sufficient to simply state that
spin-orbit coupling causes splitting of some of the degeneracies implicit in
the orbital energy level diagrams encountered so far. For example, a 4>f19
term is 12-fold degenerate (4 x 3). This splits into three sublevels as a
consequence of spin-orbit coupling because, loosely, some arrangements of
spin and orbital magnets are energetically more stable than others. The
effective number of spin-orbit components of crystal field terms depends
only on their spin and orbital multiplicities and these are shown in Table
8.2. The actual magnitude of the splitting caused by spin-orbit coupling may
be determined from the electronic spectra of gaseous transition metal atoms
and ions (excited, for instance, in an electric discharge). For these it depends
on the metal. Although, in theory, the dependence should vary with the
atomic number Z as Z\ the actual experimental dependence on Z is not so
simple, as Table 8.3, which gives spin-orbit splitting constants for some
typical ions, shows.
For Tim the free-ion value of the spin-orbit splitting constant, denoted
by C (zeta), is 158 em -I. This is the quantity which determines the actual
spectral splittings observed and which varies from one ion to another.
So, given band half-widths of ca. 3000 em- 1 , fine structure caused by
spin-orbit coupling is most unlikely to be seen in the spectra of Tim
complexes. For Ni 11 the value is 703 cm- 1, which means that small splittings
might be caused by this effect, although none have been unambiguously
found. Spin-orbit coupling is more likely to be of importance in the
assignment of the very weak bands associated with spin-forbidden transitions
(which, as we will see, largely owe their intensity to spin-orbit coupling).
These are frequently much narrower than those corresponding to spin-
allowed transitions and, consequently, much smaller splittings will be visible.
For Rhm and Irm, both of which form stable complexes, Chas values of ca.
1400 and 4400 em -1, respectively so, for complexes of these ions spin-orbit
splitting of d-d bands is to be expected. Indeed, spin-orbit coupling is of
potential importance for all ions of the second and third transition series.
Crystal field term 1Al;l "Al;l 2A:>g 3A2g •A:>g 2Eg 5Eg
Spin-orbit levels 1 1 1 1 1 1 1
Crystal field term 3Tl;l •rlg 5r1g 2r2g 3T2g •r2g 5r2g
Spin-orbit levels 3 3 3 2 3 3 3
166 1 Electronic spectra of transition metal complexes
0 1+ 2+ 3+ 4+ 5+ 6+
Ti (123) 104 131 158 (3716) (3973)
v (179) 154 187 220 253 (4900)
Cr (248) 219 256 296 337 378
Mn (334) 294 343 338 436 486 536
Fe (431) 388 441 499 554 612 673
Co (550) 500 561 625 695 760 830
Ni (691) 634 703 775 851 934 1012
Cu (857) 870 931 1037 1127 1224
m 7 7.5 7 6.5 6 6 5.5
Ru (1042) 968 1082 1201 1319 1441 1564
Rh (1259) 1177 1299 1426 1567 1689 1825
Os (3381) 3174 3531 3898 4259
lr (3909) 3690 4056 4430 4814
Data are adapted from Handbook of Atomic Data by S. Fraga, J. Karawowski and K. M. S. Saxena, Elsevier,
Amsterdam, 1976.
A more recent compilation is given by J. Bendix, M. Brorson and C. E. Schaffer, /norg. Chem. (1993) 32,
2838 but the definitions that they adopt, although with merit, are not those generally encountered in the
literature.
This is why our discussion of crystal and ligand theories has been exemplified
by complexes formed by elements of the first transition series. Had we
included spin-orbit coupling within the crystal field model (and in a more
complete treatment this would have been done), within the crystal field model
it would have appeared with its free-ion value. Not surprisingly, in ligand
field theory it becomes a parameter which, characteristically, is found to have
a value somewhat lower than that found for the free ion.
has been discussed above, some of the orbital degeneracy may be lost because
of spin-orbit coupling. If this mechanism were to totally relieve the orbital
degeneracy then there would be no need to consider the Jahn-Teller theorem.
The Jahn-Teller theorem (which is group-theoretical in origin) tells us
that a regular octahedral complex will often be unstable with respect to a
distortion, but it says nothing at all about the magnitude of the distortion. An
exceedingly small distortion, small enough to escape detection by the most
sensitive technique, could in principle satisfy the Jahn-Teller requirement.
Indeed, there is often argument over the crystallographic evidence cited in
FJg. 8. 7 An octahedral complex, compressed support of the phenomenon, although the importance of the Jahn-Teller
along one fourfold axis (taken to be the z axis). effect is generally accepted and it is frequently invoked in explanations. The
most convincing evidence for the operation of the Jahn-Teller effect in
transition metal complexes is found in studies on Cu11 complexes, although
its importance is rather more clearly established in other areas, notably some
parts of solid-state physics.
Physically, the Jahn-Teller effect may be regarded as operating as follows.
Suppose that an octahedral complex is momentarily distorted as a result of
a molecular vibration and has the shape shown in Fig. 8.7. This vibration
will cause d., and dx'-y' to lose their degeneracy (the distorted molecule has
the symmetry of the trans octahedral molecule in Table 6.6). Suppose too
that in the undistorted complex the e9 orbitals were occupied by a single
electron. In the distorted complex this electron could be in either d., or
dx'-y'· Consider both possibilities. The distortion shown in Fig. 8.7 results
from a metal-ligand stretching mode which is of the form shown in Fig.
8.8(a); from this figure two things should be evident. First, that the caption
to Fig. 8.7 is incomplete-not only do the axial ligands move in, but the
equatorial ligands move out, although by a smaller amount, actually one half.
Secondly, the general nodal pattern and amplitudes in Fig. 8.8(a) follow
those characteristic of a d., orbital. We know that, in an octahedron, d., is
partnered by dx'-y'· So, the vibration in Fig. 8.8(a) is partnered by that in
Fig. 8.8(b). The two, as a pair, have e9 symmetry. We could discuss either
member but the discussions would differ in detail. We shall consider only
the d.,.Jike vibration, that shown in Fig. 8.8(a.) This is the one for
which, if it led to a static distortion, simple crystal-field arguments would
lead us to expect that the solitary e9 electron would be more stable in the
dx'-y' orbital. If the arrows in this diagram are reversed, so that the
distorted complex is one in which the four in-plane ligands move in and the
axial ligands move out compared with their positions in the undistorted
molecule, then we would expect the e9 electron to be more stable in d.,.
Consider now Fig. 8.9, which shows a diatomic potential energy diagram
which we apply, separately, to each bond in our complex. We compare the
cost in energy terms of the (two short, four long) and (two long, four short)
alternatives. In doing this we are just talking about 'cost' and so ignoring
the Jahn-Teller stabilization which is the driving force for the distortion-
but, as said earlier, the argument is indicative only, it certainly is not
complete.
Fig. 8.8 Two metal-ligand bond-length change Remembering that the amplitude associated with the two ligands is twice
vibra~ons of an octahedral complex. Those that associated with the four, it can be seen that the inherent asymmetry in
shown, together, have Eg symmetry; (a) has a
displacement pattern of a d, orbital, (b) the potential energy surface indicates that the cost of (two short, four long)
resembles dx2-y2· will be greater than that of the (two long, four short). This conclusion is in
168 1 Electronic spectra of transition metal complexes
~
Fig. 8.9 A M--l (diatomic) potential energy
diagram which enables an assessment of
energy changes consequent of the alternative
distortions shown in Fig. 8.8.
accord with the distortions that have been observed in crystal structure
determinations and other techniques. Octahedral complexes of Cu11 are
almost invariably distorted, commonly in the way shown in Fig. 8.10. This
is entirely consistent with the operation of the Jahn-Teller effect because the
ground-state electron configuration of a Cu11 ion in an octahedral field is
~~.e~, 2 E 9 • It has an odd number of electrons in the e9 orbitals.
Fig. 8.10 The distortion commonly found in There is one result of the general theory which should be mentioned. It
octahedral complexes of Cu". In real life the can be shown that in a regular octahedron the Jahn-Teller effect only
picture is often more complicated because operates by way of vibrations of e9 symmetry when the electronic state is
the two long-bonded ligands differ from the
other four. "E 9 (the value of n is irrelevant because the Jahn-Teller theorem applies only
to space functions, not spin). For electronic states of either "T19 or "T29
symmetries then the Jahn-Teller effect operates through vibrations of either
e9 or t 29 symmetries; a vibration of the latter symmetry is shown in Fig. 8.11.
Of course, the orbital degeneracy in an octahedral complex may be relieved
by distortions other than those shown in Figs. 8.8 and 8.11. However,
in such cases we may conclude that whatever is responsible for the distortion
it is not the Jahn-Teller effect. In particular, all of the Jahn-Teller-active
vibrations of an octahedron carry the g suffix and this means that they
cannot give rise to a distortion which destroys the centre of symmetry of an
octahedron. Distorted octahedral complexes which lack a centre of symmetry
cannot owe their distortion to the operation of the Jahn-Teller effect.
Fig. 8.11 A vibrational mode of T211 symmetry
of an octahedral complex. These vibrations This account of the Jahn-Teller effect indicates why it is of little importance
change bond angles, not bond lengths. when the t 29 orbitals are unequally occupied. Occupation of these orbitals
Jahn-Teller effect 1 169
2 long 2 short
4 short 4 long
more stable less stable
Fig. 8.12 The effect of the distortion shown in
Fig. 8.8(a) on an excited state in which the involves less destabilization than does occupation of the e9 orbitals-they
Jahn-Teller effect is operative. It follows that
the excited state must be orbitally degenerate; are, to a first approximation, non-bonding.
this figure demonstrates the loss of that The Jahn-Teller effect is of importance in the spectra of octahedral
degeneracy consequent on the distortion. transition metal complexes because the transitions observed usually involve
the excitation of a single electron from a t 29 to an e9 orbital. It follows that
there must be an odd number of e9 electrons either before or after the
transition. That is, if the ground state is not subject to a Jahn-Teller
distortion, the excited state is, and vice versa. To explore the effect of a
Jahn-Teller distortion on the electronic spectrum of a complex we shall
consider a simplified model. Suppose that the Jahn-Teller effect is operative
in the excited state and that molecular distortions have transiently distorted
the molecule in its electronic ground state so that it has the shape shown in
Fig. 8.1 0. Suppose too that the stable molecular geometry in the excited state
is also that shown in Fig. 8.1 0-that is, it is that briefly assumed by the
ground state. Then there also exists a less stable excited state for which the
molecular geometry is as shown in Fig. 8.12. The two excited states differ in
that in the one the e9 electron is in dx'-y' and in the other it is in dz'· If
whilst the ground state has the molecular geometry of Fig. 8.10 it absorbs
light and assumes an electronically excited state then we have to consider
the two possibilities shown in Fig. 8.12. First, the excited state could be the
more stable one, so that the molecule finds itself in its vibrational ground
state. Secondly, the molecule could assume the less stable excited state but
would then find itself well away from its equilibrium geometry-it is
Q) vibrationally excited, possibly by several quanta. These quanta may each
"t1l<= have an energy of ca. 300 em- 1 (assuming no great change in frequency
-e
0 between ground and excited state). Remembering the intrinsic difference in
<J)
.0
<( energy between the two excited levels in their vibrational ground states, we
conclude that such excited state Jahn-Teller distortions may well be
spectrally apparent. For example, this is believed to be the explanation of
the asymmetry in the 2 E9 <-- 2 T29 d-d d transition of the [Ti(H 2 0) 6 ]3+ ion
shown in Fig. 8.13.
Fig. 8.13 The asymmetry in the d-d spectrum
of the [Ti(H 2 0) 6 ] 3 + ion (d 1 ), believed to result
If the Jahn-Teller effect operates in the electronic ground state of a
from the operation of the Jahn-Teller effect in complex it is unlikely to be apparent in the electronic spectrum. If the two
the excited state. ground states are split sufficiently far apart for transitions to the excited state
170 1 Electronic spectra of transition metal complexes
to be resolvable, the upper of the two ground states will not be sufficiently
thermally populated for transitions from it to be seen. On the other hand,
in this situation there will be an additional transition in the infrared, although
probably difficult to locate amongst the forest of vibrational bands in this
spectral region. This additional band corresponds to a simultaneous electronic
and vibrational excitation from the lower of the ground states to the upper.
A Jahn-Teller distortion may be either static or dynamic and so far only
the former has been considered. There are, in fact, theoretical reasons for
expecting static Jahn-Teller distortions to be atypical, which is perhaps why
it is so difficult to point to firm evidence for the effect in ground states. The
dynamic case occurs when the potential barrier separating, for example, the
three equivalent distortions of the type shown in Fig. 8.1 0, an elongation
along x, y or z axes, is of the order of kT. The distortion then rapidly
alternates between the possibilities. In this way a small Jahn-Teller effect
may be unobserved--the time average is an undistorted structure (as will
become apparent towards the end of Chapter 9, small distortions may be
detected by magnetic measurements). Because the time taken to jump from
one configuration to another is of the same order as the time taken for a
simple vibration, a measurement taking much longer than this to perform
will give only an average. A measurement of 10- 10 s duration takes too long!
The dynamic Jahn-Teller effect does not affect the visible appearance
of the electronic spectra of transition metal complexes but this is because of
the energetics of the effect, not because of the timescale.
of a multitude of sharp ones. The relative breadths of the spectral lines will
be determined by the relative slopes of the lines which, in a Tanabe-Sugano
diagram, represent the excited states. For example, Fig. 8.3leads us to predict
that transitions to the 3T 1.(P) and 3T 29 terms of the d 2 configuration will
give bands of similar widths. As Fig. 8.1 shows, this prediction is roughly
confirmed by experiment. We further predict that the transition to the 3 A 29
excited state will give a band which is about twice as broad as the other
two. This, together with its inherent low intensity (see below), is no doubt
why it is difficult to observe. However, as the Ni 11 example given earlier in
this chapter shows, it is possible to see such transitions in favourable cases.
On the other hand, some of the lines in a Tanabe-Sugano diagram are
almost parallel to that representing the ground state. These lines are
invariably amongst those derived from the same strong field configuration
as the ground state. For example, as Fig. 8.3 shows, in the d 2 configuration
the .1. dependence of the 1 A 19 , 1 E 9 and 1T 29 terms is roughly parallel to that
of the 3 T 19(F) ground state. These four terms are those which arise from the
t~ 9 strong field configuration. That is, a transition from the ground to another
of these terms corresponds to a rearrangement of electrons within the t 29
orbitals and so does not depend on the t 29-e9 separation, .1.. This statement
is rigorously true in the strong field limit but is only approximately true in
weak fields. In the latter case, mixing of the 1 A 19 , 1 E9 and 1T 29 excited terms
e;
derived from the t~ 9 e; and configurations, into those derived from the t~ 9
configuration, makes the lowest 1 A 19 , 1 E9 and 1T 29 term each have a different
dependence on .1.. However, in intermediate and strong ligand fields it is true
that the energies of transitions to these terms are scarcely modulated by the
.1. variations caused by ligand vibrations. It follows that transitions to these
terms will appear as relatively sharp lines in the spectrum. However, they
are very weak because they are spin forbidden (a spin triplet to a spin singlet)
and for this reason are not easy to detect. The existence of these transitions is
exploited in those lasers which involve transition metal ions, for instance in
the ruby laser. Ruby is an Al 2 0 3 crystal containing a small amount of Crill
as impurity, the chromium being approximately octahedrally surrounded by
oxygen atoms. It is on this Crill, a d 3 ion, that we focus our attention. As
Fig. 8.14 shows, the energies of the transitions from the ground term to the
lowest 2 E 9 , 2 T 19 and 2 T 29 terms are essentially independent of .1. and so the
bands are sharp. Emission of light from Crill ions in the 2 E9 excited term as
they fall back to the ground state gives rise to the essentially monochromatic
red light emitted by the ruby laser.
In the above section only the effect of the totally symmetric breathing
vibration on the value of .1. was considered. Similar conclusions follow if
vibrations of other symmetries are considered in detail. As will be seen
in the next section, some of these other vibrations are responsible for the
appreciable intensity in formally forbidden d-d transitions.
8. 7 Band intensities
The fact that very weak d-d transitiOns may be observed in atomic
spectroscopy indicates the approximate nature of the rule that only s ..... p,
p <--> d, d <--> f etc. orbital transitions are allowed. The spin-allowed d-d
transitions in an octahedral metal complex are of much higher intensity than
172 1 Electronic spectra of transition metal complexes
4 T1g IPJ
- - - - - - - , - - - - - - - - . - - - - - - - - "T-----
''
70
'
''
'
60 ---------,--------
' r---
' ---
'
'''
'
'
50 '
- - - - - - -'1 - - - - - - - - .J---
Fig. 8.14 A simplified Tanabe-Sugano diagram ' '
for the d3 configuration (appropriate to Cr"), '
' '''
showing the excited terms which have a
dependence on !1/B which is almost the same
as that of the ground state. Such parallel co 40 4T2g
LiJ
I
behaviour only occurs when the ground and
excited terms correlate with the same strong
field configuration. In the present case this is
~. (see Table 7.5). 30
~~--~~~r:~~~---:· 2Gg
'
;__;_;:_;:_: : :;_
'
'2r1g
20 d.~.-..,b.=~~~-t-:;;_:;:i:;;_;_ i 2Eg
2G
4p ''' '
10 ''
----~--------
4F
0 10 20 30
!J.fB-
the centre of symmetry of the molecule will not introduce any intensity, for
the essential requirement is that the vibration mixes atomic orbitals of q and
u types (and, so, in our example, d and p orbitals, respectively). This topic
is explored in some detail in Appendix 8, an appendix that shows that there
will be perhaps ten vibrational subpeaks, each with a L1 modulation and
Jahn-Teller broadening, leading to the relatively weak, broad peaks observed.
The above discussion is associated with a breakdown in the Born-
Oppenheimer approximation. This states that electronic and vibrational
energy levels may be treated separately, so that under this approximation
an electronic transition is something quite separate from a vibrational
transition. However, in our discussion we have found that the one depended
on the other; that there is a coupling between vibrational and electronic
states. This is called vibronic coupling. The intensity introduced into a d-d
transition by vibronic coupling cannot come from nowhere. As is evident
from our discussion, it is in fact 'stolen' from what in a regular octahedron
is an allowed transition-for example, the equivalent in ·the complex of the
free ion d --> p transition, a transition which will lie in the ultraviolet. A more
detailed analysis shows that the magnitude of the stolen intensity is expected
to be an inverse function of the energy separation between the d-d band
and the allowed band. This means that the highest energy d-d band, that
with the smallest separation from the allowed band, will generally be the
most intense, other things being equal. This perhaps explains why the
29 transition in [Ni(H 2 0) 6 ] + -a two-electron promotion (see
3 T (P) <-- 3 A 2
19
Fig. 7.18)-is seen at all.
We now turn to the origin of the intensities of the spin-forbidden d-d
transitions. One possible source is a magnetic dipole mechanism. In electron
paramagnetic (spin) and nuclear magnetic resonance spectroscopies (EPR
and NMR), electromagnetic radiation is used to 'turn over' the spin of either
an electron or nucleus in a molecule (this simple view of the process is
adequate for our present purposes). The electromagnetic radiation does this
by virtue of its associated magnetic vector and the process is said to be
magnetic dipole allowed. Classically, the alternating N-S magnetic dipole
associated with the light wave causes similar magnetic dipoles in the molecule
to alternate in direction in phase with it. When the alternation coincides
with a resonant frequency, a transfer of energy from the light wave to the
molecule occurs. Although a similar process undoubtedly contributes to the
intensity of the spin-forbidden d-d bands-such transitions are magnetic
dipole allowed-detailed calculations indicate that its contribution is small,
although, as we shall see in Chapter II, this mechanism is of importance for
f electron systems. Much more important is spin-orbit coupling. Our
discussion in Section 8.4 demonstrated that the spin and orbital properties
of an electron are not completely independent of each other. Earlier in this
section we described a vibrational mechanism by which an orbital transition
becomes weakly allowed, so, in principle at least, spin-orbit coupling
may transfer some of this allowedness into what is, formally, a spin-forbidden
transition. Let us consider a specific case.
As we have seen, the ground state of an octahedral d 8 complex is 3 A 29 .
A low-lying spin-forbidden transition is to a 1E 9 term and a spin-allowed
transition is to the 3 T 19 term (see the modified Tanabe-Sugano diagram
shown in Fig. 8.16). Now, spin-orbit coupling has the effect of contaminating
174 I Electronic spectra of transition metal complexes
40
'
'
'
'
'
'
- - - - - - - T' - - - - - - - - , - - - - - - - - l'
'
'
'
'
'
'
10 20
6./8---
two of the nine wavefunctions of the 3T 1 • term with those of the 1Eg term
(of which, of course, there are also two) and vice versa. When the energy
separation between these two terms is small (and this occurs at C./ B :::o II in
Fig. 8.1 6) the contamination becomes gross pollution! The result is that for
these four coupled wavefunctions the two lines on a Tanabe-Sugano diagram
do not cross but repel each other and behave as shown in Fig. 8.16 so that
those two wavefunctions which in a weak field belonged to 3 T 1 • become 1 E.
in a strong field and vice versa. This means that some of the intensity of
what was the 3 T 1• +- 3 A 2 • transition (before we included spin-orbit coupling)
is transferred to the 1 E 9 +- 3 A 2 • transition once spin-orbit coupling is
included. All other spin-forbidden transitions are believed to gain intensity
by similar mechanisms. One explanation which has been suggested for the
double-headed peak in the [Ni(H 2 0 6 )]2+ spectrum of Fig. 8.5 is that it
originates in this mixing, the d-d intensity being shared by the two
components. Against this, it has been argued that a detailed analysis,
including all interactions, leads to a diagram sufficiently different from Fig.
8.16 for it to be unlikely that this explanation holds for this particular
example-it is not true that the complex falls at the crossover point. An
alternative explanation for the double-headed band is that the splitting
originates in the Jahn-Teller effect associated with the r~.e~ strong field
configuration with which the 3 T 19(F) excited term correlates. But in this case
why is the lower band, to 3T 29 , which also correlates with r~.e; not also
split-and, if this is the explanation why is the band-splitting not found in
most, if not all, octahedral Ni 11 complexes? At the present time there is no
agreed explanation for the splitting.
Tetrahedral complexes 1 175
point developed in some detail in Section 7.7). This means, for example, that
the spectrum of a tetrahedral d 2 complex can be interpreted using a Tanabe-
Sugano diagram for the octahedral d 8 case because, as was discussed in
Section 7.4, the splitting pattern for d 8 is the opposite to that for d 2 . It is to
facilitate this application that Tanabe-Sugano diagrams quite often do not
contain any g or u suffixes in their term labels. There is only one major point
of difference that has to be discussed and this concerns the intensities of d-d
bands of tetrahedral complexes. As is evident from Fig. 8.17, where the
spectrum of the octahedral [Co(H 2 0) 6 ]2+ ion is compared with that of the
tetrahedral [CoC1 4 f- anion, these intensities are at least an order of
magnitude greater in tetrahedral complexes than in their octahedral counter-
parts. Evidently, some new intensity-generating mechanism is available in
tetrahedral complexes.
176 1 Electronic spectra of transition metal complexes
E 2
Q)
·;:;
~0
"
c:
0 600
Fig. 8.17 The spectra of the d7 species :g
[Co(H 2 0) 6 ] 2 + (which is weak field) and c:
[CoCI 4 ] 2 ~. The latter is two orders of 'iii
UJ
magnitude more intense.
400
200
''
''
'
''
''
''
'
FIJI. 8.18 The spectrum of [Co(NH3 ) 4 (ox)]+
(second down, indicated by vertical bars) to a
first rather good approximation is a weighted
interpolation of those of [Co(NH3 )s] 3 + and
[Co(ox) 3 ] 3 -. This is an example of an
application of the rule of average environment.
30000 15000
~(211[MleJ + L1[M6J)
Figs. 6.11 and 6.26 because they are not stabilized by interactions with the
metal orbitals. They might, then, be expected to give the lower energy
charge transfer bands (as, indeed might some of the other ligand orbitals
mentioned above, for the reason indicated at the end of Section 6.2). This is
probably the case but, as the intensity of charge-transfer bands is a function
of the overlap of the orbitals between which the overlap occurs, the
corresponding intensities are small. Clearly, there is a danger that these bands
may be confused with d-d transitions. This is in contrast to most charge-
transfer bands, which are intense, and so easy to distinguish.
Class 3-Metal to ligand (oxidation) charge-transfer
transitions (MLCT)
In this type of transition a metal electron is transferred to an orbital largely
located on the ligands. It is therefore, in a sense, the opposite to type of
charge-transfer transition described in the prevous section. So, it corresponds
to an oxidation of the metal and reduction of the ligand. An example is
provided by the [M(H 2 0) 6 ] 2 + ions of the first transition series, for which
Dainton has shown that the energy ofthe first charge-transfer band is linearly
related to the redox potential of the system
M2 +(aq) ..... M3 +(aq) + e-(aq)
Much of what has been said about ligand metal charge-transfer bands applies
to this class also-they, too, can be used as a basis for actinometry; it would
be attractive to combine both types in a catalysed photolytic decomposition
of water.
It is not always easy to decide whether category 2 or 3 is involved in a
particular transition. If it is 2, then, for an allowed transition in an octahedral
complex, the excitation must be from a ground-state orbital of u symmetry
(and, therefore, t 1• or t 2 • as there are no other occupied u orbitals) and the
electron must be excited into either a t 2 • or e. orbital. If both t 2 • and e. sets
are incompletely filled then both excitations may occur and two bands (or
families of bands) separated by Ll are to be expected. If the t 2 • set is full then
only excitation to the e• set will occur. This probably explains why there is
a long wavelength charge-transfer band in [Fe(CN) 6 ] 3 - (d 5 ) which has no
counterpart in [Fe(CN) 6 ] 4 - (d 6 ). A similar pattern is observed in the anions
[IrBr6 ] 2 - (d 5 ) and [IrBr6 J'- (d 6 )-the latter pair are illustrated in Fig.
8.19. In both cases the band in the d 5 species which disappears is about Ll
away from the one which remains (the Ll being that of the higher valence
ion, of course), as the above explanation requires. If we regard charge-
transfer spectra as involving an electron moving towards or away from a
cation then the fine details of the cation's electronic structure assume a
lesser importance, although it may influence the number of bands, as we
have seen. Adopting this viewpoint, then in cases such as those just dis-
cussed in which there are similar complexes differing by a single formal
charge, we would expect ligand -> metal charge-transfer to become easier-
and so of lower energy-as the formal charge on the metal ion increases.
Conversely, metal -> ligand charge-transfer would be expected to become
more difficult and so of higher energy. These, then, are criteria by which the
type of charge transfer transition involved in a particular band can tentatively
be identified.
180 1 Electronic spectra of transition metal complexes
m~
Finally, as already mentioned, it seems clear that the solvent molecules can
be involved in the charge-transfer process. Detailed discussions of charge-
Wavelength - transfer spectra in transition metal complexes quite often label the corre-
Rg. 8.20 The lowest energy charge-transfer sponding bands quite separately, giving them the label CTTS-charge
bands of the complexes [Co(NH3 )sl(] 2 +, transfer to solvent. So, the fact that Fel 3 has recently been prepared in
X ~ Cl-, Br-, 1-. Qualitatively, the less
electronegative the X, the longer the wavelength non-aqueous media suggests that the solvent-water-is not always the
(the easier the transition). mere spectator that it was implicitly assumed to be above.
lntervalence charge-transfer bands I 181
8.12 Conclusions
The discussion in this Chapter has, at times, been somewhat complicated,
the complications centring around such topics as the extraction of quantitative
data from spectra or the details of how a forbidden band obtains intensity.
The assignment of the band pattern has, tacitly, been assumed to be
straightforward. This is by no means always so. For first row transition metal
complexes there is seldom a problem but for the others the larger values of
~ mean that there is more overlap between d-d and charge-transfer bands.
Further reading 1 183
Further, the larger values of the spin-orbit coupling constants for these
elements make intensity and band-width criteria for distinguishing between
spin-allowed and spin-forbidden transitions less reliable. Even more difficult
is the case of lower-than-octahedral symmetry. It is probably true that the
electronic spectrum of no transition metal complex has been studied in as
much detail as has that of the square-planar [PtC1 4 ] 2 - ion. Despite all of
this work, it is only relatively recently that some general agreement on the
assignment of the spectrum has been achieved and few would regard as
completely improbable a revision of some aspect of the current interpretation.
Finally, apart from a brief reference to the ruby laser, the content of this
chapter has been concerned with the absorption of light. Some complexes
emit light also, giving rise to the phenomenon of fluorescence (the excited,
emitting state has the same spin multiplicity as has the ground state) and
phosphorescence (the excited state has a different spin multiplicity from the
ground state). Such processes arc currently exciting much interest but as the
identity of the actual excited states involved is seldom agreed they will not
be discussed in this chapter. They will be referred to in Chapter 14 because
-
a related topic, that of the reactions of electronically excited states, is better
defined.
Ql
"c:
-e"'
0
U)
17400 24500 33 300
~
cm-1
Fig. 8.21 The d-d spectrum of [Cr(H 2 0) 6 ] 3 ".
9.1 Introduction
In the last chapter, in Section 8.4, it was suggested that the spin of an electron
is best thought of as meaning that the electron behaves like a tiny bar magnet.
When there are several unpaired electrons the spin degeneracies can be
thought of as resulting from the variety of ways of arranging bar magnets
side-by-side. Similarly, the orbital motion of the electron, the circulation of
charge around the nucleus, can be thought of as leading to a solenoid-like
magnet. These spin and orbital magnets will interact with an applied
magnetic field , so that when an atom or molecule is placed in a magnetic
field any spin degeneracy may be removed- the different resultant magnets
behave differently. Thus, a level which is orbitally non-degenerate but is a
spin doublet has this spin degeneracy split into two levels with slightly
different energies in a magnetic field (corresponding to the N~S and S ~N
arrangements). If there is orbital degeneracy this too may be removed by a
magnetic field. As we shall see, it is such splittings which determine the
magnetic properties of a complex. Note particularly that either or both of
spin and orbital degeneracies give rise to the magnetic effects. Sometimes
statements such as 'the number of unpaired electrons in a complex may be
determined from magnetic susceptibility measurements ' are encountered.
Whilst true within their own context, they should not be read as meaning
that in electron spin lies the sole source of magnetic effects.
The splittings produced by magnetic fields are very small, about I em- 1
for a field of 0.5 T, and, for the majority of cases, are proportional
to the magnetic field. Because the splittings are so small, any particular atom
or molecule may be in any one of the several closely spaced states resulting
from the splitting. For a macroscopic sample, however, there will be a
Boltzmann distribution between the levels; at room temperature kT is about
200 em - I so ample energy is available. Clearly if we are to be able to interpret
186 I Magnetic properties of transition metal complexes
experimental results, we must consider both the splittings and the Boltzmann
distribution over them. However, because the effect of the magnetic field is
so small we must consider any other interaction which involves energies
corresponding to more than about I em- 1 • This is because such interactions
will play a part in determining the ground state of the complex before the
magnetic field is switched on-this field will only cause the subsequent
splitting. Two such interactions-spin-orbit coupling and the presence of
low-symmetry components in what is otherwise an octahedral crystal
field-have already been discussed in outline. The latter effect can seldom
be neglected. One might think, for example, that an isolated [Co(NH 3 ) 6 ]3+
ion would be accurately octahedral. This is not so, for there is an incompati-
bility between the threefold axis of each NH 3 and the coincident fourfold
axis of the CoN 6 octahedron. Add to this the effect of the environment (and
this means 'do not consider an isolated molecule; in solution it will be
surrounded by a jumble of solvent molecules and in a crystal by anions-and
these will seldom respect its octahedral symmetry'), recall the Jahn-Teller
effect, remember that some metal-ligand vibrations will be thermally excited
down to quite low temperatures and that lattice vibrations will persist to an
even lower temperature, and it becomes evident that from the point of view
of magnetism, a regular octahedral environment is a rare, if not extinct,
species. For basically octahedral molecules, however, an octahedral model
is a good first approximation and may be refined to take account of the
above effects. For the majority of this chapter we shall confine ourselves
largely to such octahedral complexes.
Before we can proceed, some of the vocabulary of magnetism has to be
introduced. Closed shells of electrons have neither spin nor orbital degeneracy
and are represented by a single wavefunction. A magnetic field therefore
produces no splitting. It does, however, distort the electron density slightly,
in a manner akin to that predicted by Lenz's law in classical electrodynamics.
That is, effectively, a small circulating current is produced, the magnetic effect
of which opposes the applied magnetic field. Because there is no resistive
damping, the current remains until the magnetic field is removed. Molecules
with closed shells are therefore repelled by a magnetic field and are said to
exhibit diamagnetism or to be diamagnetic.
Suitably oriented magnets are attracted towards the magnetic field of a
stronger magnet and the same is true for any orbital magnet and the intrinsic
(spin) magnet associated with an unpaired electron. Molecules with unpaired
electrons are therefore attracted into a magnetic field 1 and are said to to
exhibit paramagnetism or to be paramagnetic. For any transition metal ion
which has both closed shells and unpaired electrons the diamagnetism of the
former and the paramagnetism of the latter are opposed. All that can be
measured is their resultant. Fortunately, the effect of paramagnetism is about
100 times as great as that of diamagnetism so that it takes a great deal of
the latter to swamp the former. Fortunately, too, it is found that the effects
of diamagnetism are approximately additive-each atom makes a known,
and approximately constant, contribution to the diamagnetism of a molecule
and so diamagnetism may be fairly accurately allowed for once the empirical
formula of a complex is known. This means that it is possible to deduce the
~ = f.lo( 1 + ~) = f.10 (1 + K)
where (MW) is the molecular weight and Vis the molar volume. Therefore,
that is
(9.1)
- J1.2H
m=- (9.2)
3kT
(9.3)
where C, the Curie constant, is equal to NJ..J1 2 f3R. The equation x = C/T is
known as the Curie law-susceptibility is inversely proportional to the
absolute temperature. Surprisingly, this law is obeyed rather well by many
liquids and solids, and in particular by complexes of the first row transition
elements. The origin of this general agreement is not clear, except under
rather special limiting conditions. The detailed quantum mechanical treat-
ment does not at all readily lead to a prediction of a CfT type of behaviour.
For complexes of the first row transition elements it turns out that the
agreement is because the spin-orbit coupling constants are comparable in
magnitude to kT; at low temperatures they do not follow the Curie law so
well. Rearranging the above equation we find:
(9.4)
Orbital contribution to a magnetic moment 1 189
(3RTzM)1;2
(9.5)
l'eff = NAfJ
Notice that, defined in this way, llorr is a pure number, the Bohr magneton
number, which refers to a single molecule. Although often met, it is not
strictly correct to call it the effective magnetic moment, and it certainly is
incorrect to express it in units of Bohr magnetons, although this is a usage
met all too often.
explanation of atomic structure \Vas based on the assumption that the angular momentum of
an electron circulating about the nucleus of an atom is quantized and equal to nh 12n:, where n
is an integer and h is Planck's constant. That IS, nh/2n = ma 2 w, \Vhere m is the mass of the
electron. a the radius of its orbit and w its angular velocity in radians. The area of the circular
orbit is Ra 2 and the current to which the electron circulation JS equivalent is e x (c•J/27r). From
the theory of a current flowing through a circular loop of wire, the magnetic moment assoc1ated
w1th the circulating electron is equal to the product
er:.J ea 2 ro
current x area = - x l':a 2 = - -
2" 2
From Bohr's postulate, this equals
ea 2 nh he he
- x - -2 = n - = nji. where fi = -
2 2rcma 4rrm 4nm
That is, the magnetic moment is an integer times {J. \vhere fJ is the Bohr magneton. It transpires
that f3 is also the fundamental quantity in the modern qu<tntum mechanical treatment.
190 1 Magnetic properties of transition metal complexes
Com is t~,e~ and so the ground state is 1 A 111 • Earlier in the section it was found helpful to talk
in terms of the magnetic properties of a current flowing in a ring of supcrconducting material.
This analogy is helpful in the present context. In the point group Oh the set of three rotations
Rx, K,. and Rz transform as T 11r The analogy suggests that the components of the magnetic
field also transform as T 1 ,. This is, mdeed, the case. Now, if an excited state-call its symmetry
species r -is to be mixed with the ground state, 1 A 1 q, by the magnetic field, then the direct
product T 1 q x r must contain 1 A 11, the symmetry of the ground state. This only occurs when
r = T 1 q and so it has to be an excited 1T 111 state that gives rise to the TIP of Co 1ll species.
Reference to the d 6 Tanabe-Sugano diagram shows that such a low-lymg excited state does,
indeed, exist.
192 1 Magnetic properties of transition metal complexes
coupling splits terms with both spin and orbital degeneracy into a number
of sublevels. This phenomenon is relevant to the spectroscopic properties
of transition metal complexes and so in Table 8.2 were listed the number of
components obtained as a result of spin~orbit coupling.
There is a simple method by which one may discover whether spin--orbit
coupling mixes two d orbitals. For each d orbital, note the number of its
nodal planes which also contain the z axis (i.e.those in which the z axis lies
on a nodal plane). If two d orbitals differ by either one or zero in these
numbers then the two orbitals may be mixed by spin~orbit coupling. Using
this approach it follows that spin~orbit coupling mixes d orbitals as indicated
below. Numbers in brackets indicate the number of nodal planes which, for
that orbital, contain the z axis:
d,,(O)
/"""'-
dyz(l)---<lzx(l)
1X1
dxy(2)-dx'-y,(2)
The origin of this pattern will become more evident from the detailed
discussion that will be given Sections 11.4 and 11.5.
The magnitude of the spin~orbit coupling for a particular ion is usually
given in terms of one of two different so-called spin -orbit coupling constants,
( (zeta) and ).. The former is the one-electron spin~orbit coupling constant
and is useful when comparing the relative magnitudes of spin~orbit coupling
for different ions. In practice, for many-electron ions, what is measured is a
resultant spin~orbit coupling between the resultant spin magnetic moment
and the resultant orbital moment. It is this latter spin~orbit coupling
constant which is called ).. For ground states the two constants are simply
related:
where Sis the spin multiplicity of the ion and the plus sign refers to d 1 ~d 4
ions and the minus sign to d 6 ~d 9 For these latter, it is simplest to think of
holes circulating, the holes having the opposite charge to electrons. This
point is detailed in appendix 12, which contains a specific example, although
this particular appendix is specifically addressed to a problem that will be
encountered in Chapter II. Values for), (and thence 0 are obtained for free
ions from atomic spectral data and for complex ions from magnetic
measurements of the type discussed in this chapter.
splitting parameters. Because the effects of low symmetry are so much more
important in magnetism than in, say, spectroscopy, the magnitude of the
parameters is best determined magnetically. However, the values of the
distortions may be small and so impossible to detect by any other method;
the distortion may not even be evident in a structure determination, for
instance. In this situation the magnitude of the parameter becomes something
of a 'fudge factor'-it is given the value that produces best agreement
between experiment and theory.
There is an important theorem due to Kramer, which states that when
the ground-state configuration has an odd number of unpaired electrons
there exists a degeneracy which a low symmetry ligand field cannot remove.
This degeneracy, usually known as Kramer's degeneracy, arises from the fact
that an orbital may be occupied by an electron in two ways-the spin may
be up or it may be down. In the absence of a magnetic field these two
orientations have the same energy. The application of a magnetic field causes
the two to differ in energy-by about 1 em - 1 -and it is the greater
occupation of the more stable in a macroscopic sample which causes the
sample to be attracted into a magnetic field. When there is an even number
of unpaired electrons a low-symmetry ligand field can relieve degeneracies,
but the application of a magnetic field then causes the lowest state to become
even more stable.
9. 7 Experimental results
If there were no orbital contribution to the magnetic moment of a complex
ion, one would have to worry a great deal less about the effects of spin-orbit
coupling and low-symmetry fields. Table 9.1 lists the moments that would
be expected for ions of the first transition series if there were no orbital
contribution (spin-only moments) and compares these with room temperature
experimental data. Also indicated in Table 9.1 is whether a significant orbital
contribution is to be expected (that is, whether there is a ground state T 19
or T 29 term). In Section 9.10 the spin-only equation used to predict the
moments given in Table 9.1 will be derived. On the whole, the agreement in
this table is not at all bad and if all one is interested in is whether a complex
is high or low spin then, if the oxidation state is known, a simple room
temperature magnetic susceptibility measurement4 on a tetrahedral or
octahedral complex of a first-row element may readily be interpreted using
the data in this table. A detailed analysis shows that the reasonable
agreement between spin-only and experimental moments shown in Table 9.1
is somewhat fortuitous. For d 1-d 4 ions, spin-orbit coupling has the effect
of reducing the observed moment and, roughly, cancels any orbital contri-
bution. For d 6 -d 9 ions, spin-orbit coupling increases the observed moment
and adds to the orbital contribution. This, together with the increased
magnitude of the spin--orbit coupling constants for these ions explains the
few gross disagreements between spin-only and experimental values. Because
of the very large spin-orbit coupling constants of elements of the second
and third transition series, their complexes usually have moments much
4 Appendix 8 contains both a description of how such mcasur~ments are made and the
treatment of the data obtained; how the diamagnetic corrections are made, for instance.
194 1 Magnetic properties of transition metal complexes
Table 9.1 Comparison of calculated spin-only moments and experimental data for
magnetic moments of ions of the first transition series
Octahedral complexes
Ti 3 + dl Yes 1.73 1.6-1.75
v•• dl Yes 1.73 1.7-1.8
v3+ d2 Yes 2.83 2.7-2.9
cr"• d2 Yes 2.83 ca. 2.8
v2• d3 No 3.88 3.8-3.9
cr"• d3 No 3.88 3.7-3.9
Mn 4 + d3 No 3.88 3.8-4.0
cr"• d4 hs No 4.90 4.7-4.9
cr"• d4 Is Yes 2.83 3.2-3.3
Mn 3 + d4 hs No 4.90 4.9-5.0
Mn 3 + d4 Is Yes 2.83 ca. 3.2
Mn 2 + d 5 hs No 5.92 5.6-6.1
Mn 2 + d 5 Is Yes 1.73 1.8-2.1
Fe3 + d5 hs No 5.92 5.7-6.0
Fe3 + d5 Is Yes 1.73 2.0-2.5
Fe2 + d6 hs No 4.90 5.1-5.7
Co2 + d 7 hs Yes 3.88 4.3-5.2
Co2 + d 7 Is No 1.73 1.8
Ni 3 + d7 Is No 1.73 1.8-2.0
Ni 2 + da No 2.83 2.8-3.5
Cu 2 + dg No 1.73 1.7-2.2
Tetrahedral complexesb
cr>• dl No 1.73 1.7-1.8
Mn6 + dl No 1.73 1.7-1.8
cr"• d2 No 2.83 2.8
Mn 5 + d2 No 2.83 2.6-2.8
Fe 5 + d3 hs Yes 3.88 3.6-3.7
unknown d4 hs Yes 4.90
Mn 2 + d5 hs No 5.92 5.9-6.2
Fe 2 + d6 hs No 4.90 5.3-5.5
eo2• d7 No 3.88 4.2-4.8
Ni 2 + d" Yes 2.83 3.5-4.0
a hs = high spin, ls = low spin.
b Note that low-spin tetrahedral complexes are very rare-if any exist at all-and are not included in this
table.
smaller than the spin-only values. We shall see why this is so in Section 9.9.
Lest it be thought that the sole effect of spin-orbit coupling is that of making
life more complicated, it should be mentioned that a large spin-orbit
coupling tends to reduce the sensitivity of the magnetic moment to low-
symmetry fields, leading to a more octahedral-like behaviour-provided that
the complex is approximately octahedral to start with.
An example 1 195
9.9 An example
This Section gives an outline of an algebraic calculation of the magnetic
properties of a complex ion. The treatment is repeated in more detail in
Appendix 10. In real life, the calculations would be carried out by a computer
and some of the approximations that will be made here would not be
necessary. However, the language of the subject is so wedded to the algebraic
treatment that it is essential to study it, at least superficially.
The problem that will be considered is very simple, at least in principle.
It is that of an octahedral ti, complex. So, the theory is that of a strong field
complex of Ti 111 -strong field because the e9 orbitals are considered so high
in energy that they can be ignored. It would be good at the end of our
development to be able to say that the final equation gives an excellent
account of the magnetic properties of Ti 111 complexes. Unfortunately, it is
not possible to be so enthusiastic-indeed, it is possible that the theory is
fundamentally flawed, although it will only be possible to appreciate why
this is so after the theory has itself been developed.
The d, configuration gives rise to a 2 T 29 ground term-- three orbital
functions, each of which may be combined with either of two spin functions,
giving a total of six functions to be considered. Eventually, we will arrive at
a picture in which these six levels, although clustered together, are split apart.
We will be concerned with the Boltzmann distribution over these separated
levels and, in particular, the sensitivity of the distribution to the application
of a magnetic field. In this situation, it would clearly be ridiculous to consider
just the effects of the magnetic field and to ignore the larger effects of any
distortion and of spin-orbit coupling. For simplicity, a distortion from
octahedral which retains some orbital degeneracy will be considered. This
196 1 Magnetic properties of transition metal complexes
2:
all molecules
(magnetic moment of a molecule) x
(number of molecules with this moment)
XM = ------'-------------
(total number of molecules)
(9.6)
Rather than work with the rather unwieldy equations resulting from the
expressions given in Appendix 10 we shall simplify. We ignore the tetragonal
An example 1 197
---<_______
===
Crystal
field
Tetragonal
splitting
Spin-orbit<
coupling First-order S d d
Z econ -or er
e~m~n Zeeman
e ec effect
Magnetic field
distortion. With this step we obtain the Kotani model, in which just the
effects of spin-orbit coupling and the magnetic field on a t}. configuration
are considered. We thus obtain an important expression, derived in reasonable
detail in Appendix 10:
3\ (-3\)
--8exp - - +8
2
(3x- 8) exp( -=F) +8
f.leff =
Spin-only y4+
!lett= Jn!n + 2)
the spin-only formula (which will be derived in the next section), in which
n is the number of unpaired electrons, become appropriate. For complexes
of the second and third row transition series the spin-only formula is not
applicable and magnetic measurements over a temperature range are
absolutely essential, even if it is only the number of unpaired electrons which
is to be determined. For measurements on such compounds at room
temperature, the plateau has not been reached.
For electronic configurations other than the one discussed above, anal-
ogous calculations lead to relationships which are roughly similar to that
shown in Fig. 9.3. /l.rr does not generally drop to zero at 0 K and, down to
i.jkT;:::; 1.5 it may increase slightly with decreasing temperature, decreasing
as the temperature is lowered further. If the model is not simplified and k
(the orbital reduction factor) and t (the factor describing the distortion to a
tetragonal field), or related functions, are included in the final energy-level
expressions, the temperature dependence of /l,H is less than in the corre-
sponding case in which they are omitted. As an illustration of this, in Fig.
9.4 is shown a comparison between the predicted and experimental results
for [VC1 6 ] 2 -, the cation being the pyridinium ion, C 5 H 5 NH+. Simple
Kotani theory is roughly followed if the spin-orbit coupling constant is
reduced to 190 em - 1 from the free-ion value of 250 em - 1 (Fig. 9.4(a)). Much
better agreement is obtained if distortion and covalency are allowed for using
the full energy level pattern of Fig. 9.2. This better agreement is shown in
Fig. 9.4(b). In that figure, the theoretical curve is that calculated fork= 0.75,
( = 150 em - 1 and t = 150 em - 1 in the expression given in Appendix 10.
The negative sign on t implies that the orbital singlet lies lowest, not the
orbital doublet-the distortion is the opposite to that assumed in Appendix
I 0 and, indeed, earlier in this chapter. The job of fitting a theoretical curve
An example 1 199
1.8
l
•
1.2
100 200
(a) Temperature (K) -
"'
"-
1.2
100 200 300
(b) Temperature (K) -
to the experimental results is best done by computer; the need for high
experimental accuracy is evident.
This, then, is the algebraic model. What, if anything, is wrong with it?
Look again at Fig. 9.4(b ), where the best fit between the modified Kotani
theory and the experimental data is shown. The lower temperature points
seem to show a systematic divergence between experiment and theory. This
is not surprising, for as either a qualitative consideration of the effect of
decreasing temperature on the population of the levels in Fig. 9.2 or,
equivalently, a study of Fig. 9.4(a) (the Kotani plot) shows, the most severe
test of the theory is to be expected at the lowest temperatures, for here the
average magnetic moment has its greatest temperature dependence. The
low-temperature cut-off point in Fig. 9.4 is determined by the boiling point
of liquid nitrogen. What if the temperature is taken down to that of liquid
helium (3 K), an extension which is now normal. Inevitably, complications
set in! These complications can be many-faceted. Small distortions away from
true trigonal or tetragonal may become evident, as may long-range magnetic
coupling between what would otherwise be regarded as isolated magnetic
centres (some examples of this were given at the end of Section 9.1 ).
Fortunately, at such low temperatures it is possible to measure the changes
in thermal population of the magnetic-field split levels by another method,
that of specific heat measurements. Such additional data are not without
additional complications. In particular, the varying thermal population of
the multitude of lattice vibrations has to be taken into account. Fortunately,
at low temperatures this part of the specific heat of most solids has a T 3
200 1 Magnetic properties of transition metal complexes
dependence and may usually be allowed for with adequate accuracy. As this
example and as the data available in the literature make clear, high
temperature magnetic susceptibility data are of limited value-and some
would say that 100 K is high-and that much more is revealed at low
temperatures. An additional bonus is that at low temperatures it may be
possible to use electron paramagnetic (spin) resonance spectroscopy (EPR,
discussed in Section 12.6) to study transitions between levels such as those
of Fig. 9.2 and so to have independent measures of them. A limited number
of ions give room temperature EPR spectra which are well resolved but some
which do not are found to give quite sharp lines at low temperatures.
It is here that our circle closes. Such measurements have been made on
some Ti 111 salts and have shown the presence of a major energy contribution
which is not present in Fig. 9.2 and is also absent from the associated text
and Appendix 10. This contribution arises from a phenomenon that was
discussed in Section 8.5, the dynamic Jahn-Teller effect. Throughout our
development of the ti_ case, we assumed that the molecule under study was
rigid. The dynamic Jahn-Teller effect introduces (some, but not all) metal-
ligand vibrations into the picture. How important is all this? Well, it has
been emphasized earlier that the magnetic properties of a complex are very
sensitive to small distortions and therefore to vibrations. The lattice vibrations
mentioned above and in Section 9.1 do not distort molecules-in them the
molecules move as rigid bodies-and metal-ligand vibrations, also mentioned
in Section 9.1, will normally be frozen out at very low temperatures. The
dynamic Jahn-Teller effect is much less readily frozen out and so, indeed,
it is a potential complication. When it is included in the calculation of the
energy levels in the ti 9 case, patterns such as that in Fig. 9.5 are obtained.
The difference from Fig. 9.2, although small, is significant. Just as a
reasonably complete understanding of the electronic spectra of transition
metal complexes requires the inclusion of vibronic (vibrational-electronic)
effects so it seems likely that an understanding of magnetism will too. This
aspect of the subject is still in its infancy, although it is clear that there are
other aspects of the magnetism of coordination compounds in which vibronic
effects are implicated (the dynamic Jahn-Teller effect is just one form of
vibronic coupling).
The developments outlined in the above paragraph may well help to refute
criticisms of that theory of magnetism which has been developed in the last
few pages. When data from a large number of complexes were reviewed it
was concluded that
1. the theory produces too many ligand field parameters which, in the event,
often seem to have no obvious chemical relevance;
2. the models and parameters used for molecules with different geometries
seem to bear little relationship with each other; they do not vary with
geometry in a way that is obviously sensible;
3. molecules with very low symmetry could not be handled.
Since these criticisms were made the situation has improved, not only
because the need to include the dynamic Jahn-Teller effect has been
recognized, but also by theoretical developments. Rather than start with an
octahedron and distort it, the present tendency is to, theoretically, build up
Spin-only equation 1 201
,-
''
'''
'
''
'''
'
''
'''
'
' ''
:
==i
==\
''
''
''
''
''
''
'
\---
- _,/ ::
----
___
'
==',\
/
I
Crystal , /
field ', , ,/
\ ',,,
D , --------
~~~~c ',Tetragonal:,~~-'--<:,
effect
~~-======== ------
splitting
Spin-orbit ----
coupling First-order
Second-order
Zeeman
Zeeman
effect
effect
Magnetic field
Fig. 9.5 A schematic energy level diagram corresponding to Fig. 9.2 but with inclusion of a dynamic
Jahn-Teller effect.
That is, the allowed components run from S through to-Sand are (2S +I)
in number. Their energies in a magnetic field will be proportional to the
magnetic field strength and to the magnitude of the z component of S, the
value of S,. So, their energies will be of the form -2S,f3H, the negative sign
indicating that the higher S, values are stabilized and the 2 being the Lande
g factor for the electron-the splittings in a magnetic field are twice as great
as one would expect from the magnitude of the spin angular momentum.
So, their energies are
-2Sf3H, -2(5- 1)[3H, -2(S- 2)[3H, ... , 2Sf3H
3kTNA
-s
2:2Sz X f3 X exp (25
_z_
[JH)
2 s,"s kT
l'ett =
NAf3 2
s
2: (25
exp _ z _
[JH)
s,"s kT
L Sz ( 1
6kT -s 25zf3H)
+_ _
2 s,"s kT
s ( 2S f3H)
l'ett =
HfJ L 1+-z-
S,"S kT
so,
6kT 2f3H S
2
/lett= - X - X - X (S + 1) = 45(5 + 1)
Hf3 kT 3
Magnetically non-dilute compounds 1 203
That is,
shown in Fig. 9.8. X-rays are blind to magnetism5 and so the magnetic centres
are represented by circles in the first diagram (ligands are omitted). We now
admit the existence of antiferromagnetic coupling between some pairs of
magnetic centres. In the other three sets of cells in this figure are given three
possible antiferromagnetic arrangements of spin orientations within the
block of four cells. It can be seen that these are all different; the bottom pair
of arrangements contain only two magnetic unit cells. Compared with the
32 crystallographic point groups and 230 space groups of classical crystal-
lography there are 90 crystallographic magnetic point groups and 1651
magnetic space groups. By using polarized neutrons (neutrons have an
intrinsic spin and so behave like tiny bar magnets; in a beam of polarized
FIJI. 9.7 Schematic temperature-susceptibility neutrons these magnets are all essentially parallel), neutron diffraction data
plots for various types of magnetic behaviour:
(a) antiferromagnetism (the arrow indicates the are capable of allocating an antiferromagnetic material to its correct
maximum which has sometimes been called the magnetic space group. If this is not known, it is not possible to give a
Curie temperature, Tc, or the NSel temperature, complete theoretical discussion of the magnetic properties of an antiferro-
TN); (b) paramagnetism; (c) ferrimagnetism;
(d) ferromagnetism. Note that this diagram does magnetic material.
not extend to very low temperatures (see Antiferromagnetic materials show a maximum in a plot of XM against
Question 9.5, for instance) or high magnetic temperature, at the so-called Nee! point-this is shown in Fig. 9.7. Below
fields (when saturation effects occur).
the Nee! point the susceptibility is to some extent field-dependent. These
properties provide an indication of the existence, or absence, of antiferro-
magnetism. Other tests include a comparison of the results of solution and
solid-state measurements, where this is possible, and, where it is not, by
the technique of the dilution of the magnetic ions in the lattice by their
partial substitution with an isomorphous non-magnetic ion. At temperatures
sufficiently above the Nee! point, antiferromagnetic materials follow a
Curie-Weiss law. Even if the Nee! point is at too low a temperature to
be measured conveniently, the observation that a material follows the
Curie-Weiss law usually implies a residual antiferromagnetic coupling.
The third class is that of ferrimagnetic materials. Just as for materials
0 0
t t
0 0 t t
0 0
t t
fill. 9.8 Crystallographic and magnetic unit
0 0 t t
cells. X-ray diffraction would give for all of these
systems the unit cell pattern of the top
left-hand diagram. Polarized neutron diffraction
would show doubled, magnetic, unit cells for
two of them.
5 This is a marginal overstatement. If the very intense X-ray beam from a synchrotron source
is used, weak X-ray magnetic scattering-which is relativistic in origin-can be observed. The
effect is rendered more observable by tuning the X-ray energy to coincide with an energy
difference in the material under study, when a resonance enhancement occurs.
206 1 Magnetic properties of transition metal complexes
(c)
(b)
that ferromagnetic coupling occurs between two metal atoms, leading them
to have parallel spins. An example of such a ligand is shown in Fig. 9.10.
Although with cunning choice ofligand and metal ion it is possible to obtain
ferromagnetic complexes such as that shown in Fig. 9.10, in most cases the
coupling between two linked metal atoms, each with unpaired electrons,
will contain at least one antiferromagnetic coupling pathway-a pathway
involving non-orthogonal orbitals-and, perhaps, a ferromagnetic coupling
through another orbital sequence. In such cases, the antiferromagnetic
coupling almost invariably dominates. Lastly, an important word, none the
less important for having been left until the end. The discussion of the last
few paragraphs hinged on the effects of exchange between electrons on
different atoms. Irrespective of whether the outcome is ferromagnetic or
antiferromagnetic coupling, the general mechanism involving ligand orbitals
to mediate coupling between metal electrons is referred to as superexchange.
Finally, although strictly out of place in this chapter, we mention
high-temperature superconductors. They are introduced at this point because,
as in superexchange, superconductivity involves the properties of electrons
on different atoms being correlated with each other. At the time of writing
this book the detailed mechanism by which high-temperature superconduc-
tivity occurs is not known. The interested reader will find the problem
explored in Appendix 11; they should be warned, however, that this appendix
is likely to date rather rapidly!
• 'A Local View of Magnetochemistry' M. Gerloch, Inorg. temperatures' J. E. Rives, Transition Met. Chern. (1972) 7, I.
Prog. Chern. (1979) 26, 1. A more general account is 'Magnetic Symmetry' by W. Ope-
chowski and R. Guccione, Chapter 3 in Magnetism, Volume IIA,
Problems in the effects of magnetic coupling are reviewed in G. T Rado and H. Suhl (eds.),Academic Press, New York, 1965.
'Magnetochemistry: a research proposal' R. L. Carlin, Coord. More recent, and concentrating on systems of two transition
Chern. Rev. (1987) 79, 215. metal ions, is 'Magnetism of the Heteropolymetallic Systems'
An excellent current book is Molecular Magnetism 0. Kahn, 0. Kahn, Struct. Bonding (1987) 68, 89.
VCH, Weinheim, 1993; a rather different account is 'Organic Spin equilibria are covered in
and Organometallic Molecular Magnetic Materials-Designer
Magnets', a very long review by J. S. Miller and A. J. Epstein in • 'Dynamics of Spin Equilibria in Metal Complexes' J. K.
Angew. Chern., Int. Ed. (1994) 33, 385. Beattie, Adv. Inorg. Chern. (1988) 32, I.
A detailed account of magnetic phases and phase transitions • 'Static and Dynamic Effects in Spin Equilibrium Systems'
at low temperatures is 'Magnetic phase transitions at low Coord. Chern. Rev. (1988) 86, 245.
Questions 9.5. In Fig. 9.9 and the associated discussion within the text,
the ligand orbital involved in superexchange was a <I orbital.
9.1. Show that the metal d orbital and the ligand orbitals
<I Ligand n orbitals may also be involved. Figure 9.12 is the ligand
that contain unpaired spin density in Fig. 6.31 are orthogonal n orbital equivalent of Fig. 9.9(a). Give an account of Fig. 9.12
to each other. (Hint: consider their symmetries.) which parallels that given in the text for Fig. 9.9(a).
1 In the gas phase ferrocene has an eclpsed (D,.) geometry; in the crystal it is staggered (Ds;).
The energy difference between the two is small (ca. 4 kJ mol- 1 , 0.8 kcal mol- 1 ). Although the
symmetry labels change between the two, the various interactions involved in the bonding are
little different. In this chapter the staggered configuration will be the one discussed because most
of the theoretical and experimental work invoked assume this geometry.
Bonding in transition metal organometallic complexes I 213
(CP)2
FIJt. 10.1 Schematic molecular orbital energy An example in which the IS-electron rule and the symmetry-determination
level diagram for ferrocene. A more accurate
diagram is given in Fig. 10.9. (Adapted and
of interaction patterns come together rather strongly is in the molecule
reproduced with permission from Shriver, hexacarbonylchromium, Cr(COk This is an octahedral complex, to which,
Atkins and Langford, Inorganic Chemistry). in principle at least, all of the relevant arguments of Chapters 6 and 7 may
be applied. That the IS-electron rule is satisfied is easily seen; a Cr 0 species
has six valence-shell electrons-which will fill the t 2 " set of d orbitals--and
each CO ligand contributes two (J electrons from each carbon (formally,
these electrons are (J-donated to the metal), giving a total of 18. It is perhaps
not surprising that Cr(C0) 6 is colourless; CO, like CN-, is a strong field
ligand (we met this in Section 6.2.2) and so the t 2 , ---+ e, transitions fall in
214 1 Beyond ligand field theory
cb I
M
{a)
~~ dx2- y2 dxy
the near-ultraviolet region of the spectrum. This, then, is a molecule for which
Figs. 6.11 and 6.16(c) must be brought together, and this is done in Fig. 10.5.
It would be worthwhile for the reader to stop at this point and examine in
some detail the connection between these three figures. One hidden problem
relates to the simple picture of COn bonding given in Section 2.2 and shown
in Fig. 2.1. The simple picture has, effectively, two electron pairs back-
donated from metal d orbitals into the n antibonding orbitals of the CO.
The group theory associated with Fig. 6.16, in which it was shown that the
12 ligand n orbitals transform as t 19 + t 1• + t 29 + t 2 ., is to be compared
with the symmetries of the available metal orbitals; of these four sets, only
two, t 1• + t 29 , find matching partners in the metal set. So, the simple picture
Metal-fullerene complexes I 215
............................................................................................................................................................................................
n= 3 4 5 6
~+ ~~+ +- + + __ +
+ + + + + + 0 nodes
- - - -
- - - - - -
(2)
fl- ~+
-
+
- (2)
-
+
+
-
+
+
- (;)
-
1 node
~~$-b 2 nodes
(b)
FIJI. 10.2 (continued) (b) Different planar overestimates the extent of n back-bonding. In Section 10.7 the qualitative
cyclic CnRn ring systems differ in their p, orbital picture of bonding in Cr(C0) 6 will be compared with the results of detailed
nodality patterns: n = 2, zero and one; n = 3,
also zero and one. However, there are two calculations.
different orbitals with one nodal plane
containing the z axis (indicated by the (2) in the
figure). The nodal plane shown is that parallel 10.2 Metal-fullerene complexes
to the plane of the paper; the other orbital has
its nodal plane perpendicular to the plane of In recent years some considerable excitement has been generated by the
the paper. The case n = 4 has zero, one (two
of) and two nodal planes; the case n = 5 has discovery of a series of carbon cage compounds, generated in low yields
zero, one (two of) and two (two of). The case when an electric arc is struck between carbon rods in a low pressure of a
n = 6 has zero, one (two of), two (two of) and gas such as argon. Although by no means the only species formed, there has
three. The case n = 7 (not shown) has zero,
one (two of), two (two of), three (two of) etc. been particular study of C 60 and C 70 , the major products of the preparation.
Of these, C 60 has an icosahedral structure and, because of the general
resemblance to the geodesic domes designed by Buckminster Fuller, has
come to be known as buckminsterfullerene; it is shown in Fig. 10.6. The
whole family has come to be called fullerenes. In contrast to C 70 , which has
a D 5 h structure and five structurally different types of carbon atom, all of
the carbon atoms of C 60 are structurally equivalent. Our discussion will
therefore concern this molecule alone. The molecule is just over 7 A in
diameter, the central hole has a diameter of half of this. It is therefore large
enough to accommodate an atom and such metal complexes have been
prepared. They have been called endohedral complexes (as opposed to
216 1 Beyond ligand field theory
~ ~
~
I M--M
~
M
~
M
M
I Q M
Q M--M
9 M
Fig. 10.3 The benzene molecule can function as a six-electron, a four-electron and a two-electron
donor. It does not necessarily remain planar and may be involved in donation to more than one metal.
The model of donation is usually clear from a crystal structure determination and, in all probability,
also by application of the 18-electron rule. Examples of relevant molecules are, left to right:
six-electron donation-[Mo(~ 6 -C 6 H 6 )(C0) 3 ], [V2 (~ 3 .~ 3 -C 0 H 0 )(~ 5 -C 5 H 5 },H 2 ],
[Os 3 (~ 2 , ~ 2 , ~ 2 -C 0 H 6 )(C0) 9 ];
four-electron donatlon-[Os(~ 4 -C0 H0 ) ~°C 0 H 0 )], [ (Re(~ 5 -C5 Me 5 ) (C0) 2 ) 2 (~ 2 , ~ 2 -C 0 H 0 )],
[Pd 2 (~ 2 , ~ 2 -C 0 H 0 ) 2 (AIC1 4 ) 2 ];
two-electron donatlon--[(~ 2 -C 0 H 0 )Re(C0) 2 (~ 5 -C 5 Me 5 )].
CH2=CH-CH=CH-CH=CH2
3 3
2 2
1 1
'
'
~0
(a)
~0
(b)
FIC. 10.4 P. molecular orbitals of some linear number. However, they are different when viewed from the perspective of
conjugated molecules in the HOckel
approximation (this is a crude, but adequate,
the rings of carbon atoms that make up the C60 cage, each ring being
model). Nodal planes are shown dotted and approximately planar. From this point of view, it is a reasonable approxi-
the total indicated. The more nodes, the higher mation to regard them as the counterparts of the n orbitals of the organic
the energy. (a) Butadiene (in the Huckel
approximation the cis and trans species are the
ring molecules discussed earlier in this chapter. They can be treated similarly
same). (b) Acrolein-this molecule may be and the MOs to which this gives rise ordered in energy sequence according
regarded as butadiene in which a terminal CH2 to the number of inherent nodal planes that they possess. This has been done
group has been replaced by 0. The detailed
consequences for the p. molecular orbitals are
in Fig. 10.7.
considerable but the general pattern is Because of the high symmetry of the icosahedron, higher even than that
unchanged. (c) Hexatriene-comparison of this of the octahedron, orbital degeneracies of four and five occur in icosahedral
pattern with (a) indicates the nature of the
general pattern.
molecules and this will be seen in Fig. 10.7, along with their symmetry labels
in the icosahedral group (although these labels add nothing to a discussion
except at a rather detailed level). Paralleling the organic rings discussed
earlier and in Appendix 13, the nodal patterns of the C 60 cage orbitals have
to match those of the orbitals of the endohedral atom. Fortunately, again
because of the near-spherical symmetry of the icosahedron, this is rather
simple (and can be shown to be so by looking at the detailed significance
of the symmetry labels in Fig. 10.7). The C 60 no-node combination interacts
218 1 Beyond ligand field theory
4p===~
4s----~
eg(2)
3d
eg eg(1)
81g
Metal cr-only cr+n Ligand cr,
orbitals molecular molecular nand
orbitals orbitals somerr'
group
orbitals
Fig. 10.5 A qualitative a + n bonding exclusively with the s orbital of the endohedral atom, the three C 60 one-node
molecular orbital energy level diagram for
Cr(C0) 6 based on Figs. 6.11 and 6.16(c)
combinations interact with the endohedral atom's p orbitals, the five C 60
together with the associated discussion. This two-node combinations interact solely with the endohedral atom's d orbitals,
schematic diagram may be compared with Fig. the seven C 60 three-node combinations (which split into two little-separated
10.10 which is a more accurate molecular
energy level diagram for Cr(C0) 6 •
sets) interact with the endohedral atom's f orbitals, and so on. Now, just as
in the neutral hydrocarbon rings discussed earlier, the total number of
electrons to be accommodated in the C 60 orbitals is equal to the number of
atoms, 60. So, the bottom 30 orbitals are full, and this is indicated in Fig.
10.7. This means that the no-node, the one-node, the two-node and the
three-node orbitals (and 14 others, too) are all full. When they interact with
an endohedral metal atom the only interactions that can contribute a
stabilization are those with empty endohedral atom orbitals. When an
endohedral atom has occupied d orbitals in its valence shell then those d
electrons are forced to occupy antibonding orbitals if interactions involving
the d orbitals contribute significantly to the molecular stability. Similar
arguments apply to the s and p orbitals, of course.
Metal-fullerene complexes 1 219
The experimental observation that the only atoms which form stable
endohedral complexes are those with empty orbitals is at once explained.
The content of Chapter II suggests that the f electrons of the lanthanides
are expected to be too well tucked away inside the atom to be much involved
in bonding and so, for this reason, the above discussion does not apply very
strongly to the C 60 three-node orbitals; Fig. 10.8 illustrates this argument.
Although this seems reasonable, some recent rather detailed calculations on
Ce@C 28 suggest that at least some f orbital covalency may exist, although
not enough to alter the above general conclusions. Hints that f orbital
covalency is not unreasonable will be found in Chapter 11, at the end of
Section 11.2. Finally, it should be noted that there is a similarity between the
fullerenes and the metal clusters which form the subject of Chapter 15. The
methods which have been used to describe the bonding in these clusters are
also applicable to the fullerenes although in the case of the spherical tensor
FJg. 10.6 The cage molecule C60 , method (Section 15.5-the method is to some extent anticipated by the above
buckminsterfullerene. discusssion) the complication of d orbital involvement ,on the cage atoms is
absent.
t2g
gu -- -- 8
7
gg ---- 6
hu ----- 7
-----
t2u 5
hg 6
t1g 6
5
gg+ hg
********* 4
gu
**** } 3
t2u
***
*****
hg 2
t1u
*** 1
Bg
* 0
Nodal
planes
220 1 Beyond ligand field theory
---- 1
-----
'
-----.. ' ',,_
*****
I
·.,....---- d
* * * * * * * * */
J2
'
''
'
Fig. 10.8 Schematic diagram of the interaction
between the orbitals shown in Fig. 10.7 and the
**** ''
atomic orbitals of an endohedral atom. ***
* * * * *'~-- 2
1
0
Nodes
although one of these will have a hole because of the selected electron, which
is ignored for the moment. The repulsive field generated by all of the included
electrons is averaged and taken as that experienced by the selected electron,
which modifies its orbital accordingly. The selected electron is then placed in
this modified orbital. This procedure is repeated for all of the electrons, each
being selected in turn, over and over again, until the input and output
arrangements are essentially the same and a self-consistent field is obtained.
A weakness of this model is that it effectively smears the density of the
selected electron over the whole molecule; in reality, the other electrons will
tend to stay away from where the selected electron is, not from a smeared-out
distribution-the electron positions are correlated with one another. In
calculations this problem can be dealt with, in some measure, by adding to
the ground state wavefunction additional functions which change it so as to
compensate for the error. The functions most evidently available are those
involving what would formally be regarded as excited states and so different
allocations of electrons to orbitals-that is, different electron configurations.
In this way the interactions between these configurations is, has to be,
incorporated. So, the process of improving the ground state wavefunction
involves so-called configuration interaction. Other things being equal, the
lower-lying the excited state, the more extensively it is likely to be involved
in the ground state correction.
For transition metals there is a large number of low-lying excited states
to be considered; many of these-and the ground state also-may not be spin
singlets, a situation of particular difficulty (the theory is different). Further,
the d electrons are confined to a small volume and it can be difficult to
maintain a balance between these and the more diffuse ligand densities.
Unless a proper balance is maintained, the more diffuse ligand orbitals may
distort so as to compensate for the deficiencies in the less adaptable set. Add
to this the fact that the correlation energies are likely to be comparable in
magnitude to metal-ligand binding energies. Further, relativistic corrections,
needed where spin-orbit coupling or heavy elements are involved, can alter
bond lengths and total energies and you have a very difficult situation.
Indeed, several examples are known in which apparent excellent agreement
with experiment has disappeared when the calculation was improved,
showing that the initial good agreement was, in fact, illusory. Lest this appear
too pessimistic, it is to be emphasized that good calculations are beginning to
appear, but they are far from routine. Ferrocene2 provides an excellent
example. After a calculation including configuration interaction with over a
million excited state configurations and a correction for relativistic effects,
the Fe-C bond length was still 0.07 A longer than found experimentally, a
discrepancy which should be compared both in magnitude and direction
with that quoted above as resulting from far less sophisticated calculations
on organic molecules. This error of0.07 A seems to be common-it has been
found in the c-cr bond length in di-1] 6 -benzenechromium and also in
1] 6 -benzenetricarbonylchromium. The error probably reflects a common
deficiency in the functions used to describe the orbitals of the metal atom,
perhaps a need to include f orbitals. It is surely significant that it is only for
2 Sec H. P. Liithi. P. E. M. Siegbahn, 1. Almlof, K. Faegir and A. Heiberg, Chem. Phys. Lerr.
(1984) l 11, 1.
222 1 Beyond ligand field theory
used both for molecules and solids. Its status is evident in the fact that when
detailed ab initio or similar calculations are published it is normal to compare
the results obtained with those given by the extended Hiickel method. Later
in this chapter there is an example which illustrates how, despite all its
shortcomings, the method can provide insights which could scarcely have
been obtained in its absence.
a2u
a,.,
a The G C5 H5 were not included m the calculation.
228 1 Beyond ligand field theory
ferrocene. That is, it comes from the orbital which is at the bottom
of the ab initio list in Table 10.2. The reorganization energy compensates for
the fact that the a 19(3d) orbital in ferrocene is relatively low in energy. The
reason that e29 is the second orbital in the non-ab initio list, but not in the
ab initio list is similar. There is a basic and important reason for all of this.
Both of the more readily ionized electrons, a 19 and e29 , are largely metal in
character. In the neutral molecule they are confined to a small volume of
space and it is not surprising that the molecular electron density distribution
should change significantly when one of these electrons is removed. It is the
consequential change in energies that result in the apparently anomalous
position of the a 19 and e29 orbitals in the neutral molecule energy level
sequence. In contrast, orbitals which are higher lying in the ferrocene ab
initio energy level sequence are largely ligand in nature. This means that
their electron density is spread over at least 10 atoms and so is relatively
diffuse. As a result, the change in the molecular electron density distribution
following the removal of such an electron is relatively small and so the
rearrangement energy is small also. This is not just some subtle difference
between ab initio and other calculations. If, as well may be the case, we are
interested in the bonding in the ferrocene molecule, then the energy level
pattern given by the ab initio method is that which is appropriate-it more
accurately shows the stabilizations resulting from the bonding interactions.
If, on the other hand, we are more interested in the chemical reactions of
ferrocene-that is, in situations in which electron density is displaced, then
the orbital energy level diagrams of the other methods become more relevant.
Because of its relative ease of ionization, a 19 and e29 behave as if they are
the HOMOs, even if, strictly, they are not. This, of course, is a situation
which is very similar to that met at the end of Section 6.2, where for classical
coordination complexes it was found that the (incompletely filled) d orbitals
are not, in fact, the highest lying occupied orbitals.
At the present time it does not seem to be unambiguously determined
whether the lowest ionization potential of ferrocene corresponds to the loss
of an electron from the a 19 orbital or whether it comes from the e 29 • As
Table 10.2 shows, most theoretical work favours a 19-and there are electronic
spectroscopic arguments in favour of this assignment-but the ab initio
calculations point to the alternative, e 29 assignment and find support in both
EPR and photoelectron spectroscopic data (the latter will be given and
discussed in Section 12.7).1t is likely that the ab initio assignment is correct.
The results of an ab initio calculation on ferrocene are also shown in Fig.
10.9.1t is perhaps more useful than the same data in Table 10.2 in addressing
the question of whether these calculations support the simple picture of the
bonding in ferrocene given in Section 10.1. The answer, comfortingly, is yes.
This conclusion is perhaps most easily seen by looking at the empty,
antibonding orbitals in Fig. 10.9. As we have seen, occupied n orbitals in a
hypothetical (C 5 H 5 )z unit which could act as electron donors to the iron
atom are of A 19 + A 2 u + E 1• + E 1• symmetries. If these interactions do
indeed contribute to the metal-ligand bonding then we would expect this
to be signalled by destabilized, antibonding, orbitals of these same symmetries.
As Fig. 10.9 shows, such orbitals do, indeed, exist. Looking at antibonding
orbitals in this way is a simple, approximate, way of avoiding the problems
posed by the plethora of c--c and C-H bonding orbitals in the bonding set.
Three examples: ferrocene, hexacarbonylchromium and ethenetetracarbonyliron I 229
Metal orbitals
Alg p, ring molecular orbitals
FJC. 10.9 Schematic ab initio molecular orbital Rather similar arguments hold for the potentially 1t acceptor orbitals on the
energy level diagram for ferrocene (data
from T. E. Taylor and M. B. Hall, Chem. Phys.
hypothetical (C 5 H 5 ), unit, E 2 • + E 2 •• There is no metal orbital of E 2 •
Let!. (1985) 114, 338). This figure differs from symmetry but there is one of E 2 •• Again, there is an empty E 2 • orbital,
Table 10.2 in that the table, but not this figure, corresponding to an occupied one, just as required by the simple picture of
includes C5 H5 ring orbitals. This diagram does
not demonstrate the fact that because of
Section 10.1.
configuration interaction the average occupancy
of the doubly occupied bonding orbitals is
sligh~y below 2 (ca. 1.98) and that of the
o
empty antibonding slighUy above <ca. 0.02). 10.7.2 Hexacarbonylchromium
Section 10.1 contains a schematic molecular orbital energy level diagram
for Cr(C0) 6 , Fig. 10.5, which was derived from a model in which it was
regarded as a complex in which extensive 1t back-bonding occurs from metal
t 2 • to 1t*t 2 •• There is abundant evidence that this back-bonding is a real
phenomenon. Theoretically, all detailed calculations concur on its reality;
experimentally, evidence for its presence is revealed by a detailed analysis of
the electron distribution about the Cr--c--o axis, obtained from accurate,
low-temperature X-ray diffraction measurements. Perhaps the simplest
evidence is provided by the changes in the average v(CO) stretching
230 1 Beyond ligand field theory
t1u t2u
eg eg
a~g t1u
a~g t:zg
t:zg
Three examples: ferrocene, hexacarbonylchromium and ethenetetracarbonyliron 1 231
'
s, 81g
t2u
tlg
t2u
'"' 12
co It'
t1u t1u
d, eg
t2g
t?g
eg
Metal orbitals
6
t1u co (J
81g
tlu
tl.g tlg
12
t2u t2u
C01t
t?g t1u
co
t1u t?g orbitals
eg (CO)s
orbitals
81g
Cr(CO)s
molecular
orbitals
FJC. 10.10 Schematic X,. molecular orbital As was discussed in Chapters 6 and 7, the occupied ligand orbitals of an
energy level diagram for Cr(C0) 6 (data from
R. Arratia-Perez and C. Y. Yang, J. Chern. Phys.
octahedral complex can be classified as either a or lt. The latter span the
(1985) 83, 4005). Electrons which would be irreducible representations T 1• + T'" + T 2 • + T 2 •• Of these, the T,. and T 2•
counted in an application of the 18-electron are non-bonding because there are no orbitals of these symmetries on the
rule are shown as lines, those that would be
excluded are shown as dots.
metal. It is therefore not surprising to see in Table 10.3 that they are found
to have almost identical energies-any differences are due to ligand-ligand
interactions. Three of the entries in Table 10.3 show a cluster of orbitals of
T 1• + T 1• + T 2 • + T 2 • symmetries which it is tempting to equate with the
entire n set. This identification would not be entirely valid for the T 1 • orbitals.
There are two sets of orbitals of this symmetry and both are mixtures of
ligand a and ligand n. The ligand a contribution is the greater for the higher
lying, not for the lower. For the lower, the n contribution dominates. Both
of the approximate methods place an orbital of T 29 symmetry as the lowest
listed; the actual wavefunctions show a significant metal t 2 • component. The
ab initio and XIX calculations do not give this result, and we must conclude
that it is an artefact introduced by the approximations of the less rigorous
methods.
232 1 Beyond ligand field theory
10.7.3 Ethenetetracarbonyliron
This compound, Fe(C0) 4 C 2 H 4 , is prepared by the reaction between a high
c pressure of C 2 H 4 and Fe(C0) 5 and has the structure shown in Fig. 10.12.
0 It is a yellow oil at room temperature and is relatively unstable, decomposing
Fig. 10.12 The molecule Fe(C2 H4 )(C0) 4 • The into C 2 H 4 and Fe 3 (C0) 12 , the latter presumably being formed from three
arrangement of atoms around the Fe is
approximately octahedral (the edges of the Fe(C0) 4 units. This section will be concerned with the answers to two
octahedron are shown dotted). questions, to which we will find that the insights provided by extended
Three examples: ferrocene, hexacarbonylchromlum and ethenetetracarbonyliron 1 233
c
(a) 0
c
(b) 0
234 1 Beyond ligand field theory
o[XJ:~~o
bonding. Put another way, electron density in the A 1 fragment orbital means
population of an Fe-(C0) 2 anti bonding orbital. In retrospect, it can be seen
that the existence of Fe(C0) 5 , a trigonal bipyramidal molecule, provides an
Fe 81 argument in favour of the 'A 1 high energy, B 1 low energy' pattern.
This molecule differs from Fe(C0) 4 C 2 H 4 by a CO group taking the place
+ -
of the C 2 H 4 . Following the usual model of CO bonding to a metal, a
donation from the carbon of the CO group to the metal requires an empty
Ftg. 10.14 The A1 and 8 1 combinations of A 1 orbital on the iron. Similarly, back bonding into an empty n antibonding
Rg. 10.13(b) as given by the extended HUcke! orbital of the CO requires a filled 8 1 orbital on the iron. This A 1 high, 8 1
method, viewed along the linear OC-Fe-CO axes low, pattern is just that revealed by the extended Hiickel calculations.
of Fig. 10.13(b). The nodal plane indicating
Fe-C antibonding in the A1 orbital is dotted. Before concluding our discussion of the Fe(C0) 4 unit it is worthy of
The Fe-<: overlaps responsible for the bonding comment that it is often thought of as a member of an isolobal series
interactions in the 8 1 orbital are shown as such as that given in Fig. 10.16. Members of an isolobal series are always
double-headed arrows.
shown as related to each other by the symbol --..->. Membership of such a
series is based on experimental evidence as much as theoretical, but members
of an isolobal series usually have similar orbital patterns, both in terms of
orientation and energy. Experimentally, they form similar compounds-so,
each member of the middle series shown in Fig. 10.16 forms an H 2 com-
pound. However, of the members of this series only Fe(C0) 4 has the A 1 high,
B1 low energy level pattern. Theoretically, this species is the only member of
1t antibonding
I
bi
~\ I
----~\
b1 \ \ 1I
FJg. 10.15 Schematic molecular orbital
diagram showing the interaction between the n
orbitals of C2 H4 and the Fe(C0) 4 fragment
Orbitals of
Fe(C0)4
\' ,'/.I/
.· . 1t bonding
~~'-----b1...
orbitals.
"''""' /
Final comments 1 235
---u-- -u-
r
-CH, -Mn(C0)5 -Co(C0)4
-
0Fe(C0) 2 Cp
-~C>
----o- -H
----o- -Br
'-....C 0 (CO)C ~
>=&
~0 ~ ~S
~ p u ~ u ~
the series for which the two electrons accommodated in the pair of orbitals
shown would normally be regarded as paired. This might be thought of as
a weakness in the isolo bal concept, but, in fact, the value of the concept lies
in its flexibility. It has come to be widely used in inorganic chemistry.
symmetry, but not the latter, the bismuth 6s orbital can participate in
the LUMO; for a discussion on the relevance of this see the footnote at the
beginning of Section 11.3 ). One must anticipate similar, if often less dramatic,
phenomena for many compounds of the heavier elements. An example of
one of the more dramatic phenomena is the prediction that relativistic effects
may well lead to the molecule HgF4 having some stability (to date, it has
not been prepared). Finally, there is the development of spin-coupled valence
bond theory. This book, like most of its generation, has concentrated on
molecular orbital theory. Valence bond theory was mentioned at the
beginning of Section 6.2.1 but then forgotten. Subsequently, problems were
found with the molecular orbital theory model arising from electron
repulsion. Such problems are much less severe for valence bond theory, but
it has problems of its own. These are in large measure overcome in
spin-coupled valence bond theory and so this method offers the prospect of a
quite new approach to the electronic structure of many of the molecules of
this chapter. This development is for the future because, at the moment, the
method is only being applied to diatomic transition metal species.
Spectroscopy, A. B. P. Lever, Elsevier, Amsterdam, 1984. A S. J. Essex, M. Gerloch and K. M. Jupp Mol. Phys. (1993) 79,
more recent overview is provided by 'The Angular Overlap 1147, an article which is useful for the non-mathematically
Model as a Unified Bonding Model for Main Group and minded because towards the end it contains a review of the
Transition Metal Compounds' by D. E. Richardson, J. Chern. physical significance of the parameters contained within the
Educ. (1993) 70, 372. More developed forms of the model are developed angular overlap model.
the subject of a review by M. Gerloch and R. G. Wooley in Finally, the spin-coupled valence bond theory is described
Frog. Inorg. Chern. (1984) 31, 371. Most recent developments in 'Applications of Spin-Coupled Valence Bond Theory', D. L.
are covered by C. A. Brown, M. J. Duer, M. Gerloch and R. Cooper, J. Gerratt and M. Raimondi, Chern. Rev. (1991) 9/,
F. McMeeking in Mol. Phys. (1988) 64, 825 and M. J. Duer, 929.
10.2 It has been suggested that there is a similarity between 10.6 In the text it was shown that the simple picture of n
the bonding of C 2 H4 to CH 2 in cyclopropane, C 3 H 6 , and the back-bonding in metal-Cr bonding given by pictures such as
Chatt-Duncanson model of the bonding of the C 2 H 4 to Pt in Fig. 2.1 overestimates the extent of this bonding in Cr(C0) 6 .
Zeise's salt. Critically assess this suggestion. Show that it similarly overestimates it in the tetrahedral
molecule Ni(C0) 4 . For this, the character table of the T, point
10.3 On a relatively small modern computer it is now possible group will be needed.
to carry out approximate ab initio calculations on quite large
organic molecules. However, with none of the programs avail- E 8C3 3C2 6S 4
Td 6ad
able for this, is it possible to include a transition metal atom?
Outline the reasons for this and explain how approximate A1 1 1 1 1 1
methods attempt to circumvent the problem.
A2 1 1 1 -1 -1
10.4 Crystal and ligand field theories predict that the d 0 ion
E 2 -1 2 0 0
[TiH 6 ] 2 - will be octahedral. Extended Hiickel, however, pre-
dicts a C2 " bicapped tetrahedron, structure. Initially, ab initio T1 3 0 -1 1 -1
methods predicted an octahedral structure but with the inclu-
sion of configuration interaction a trigonal prismatic structure T2 3 0 -1 -1 1
is indicated (see Inorg. Chern. (1989) 28, 2893). Using this
discordance of results as a basis. suggest those situations in
which each method may be expected to make reasonably
reliable geometry predictions (if ever!).
f electron systems: the
lanthanides and actinides
11.1 Introduction
In recent years there has been an increasing study of compounds of the
lanthanides and, to a lesser extent, of the actinides. These two groups
of elements have varying numbers of electrons in their f orbitals (4f for the
lanthanides and 5f for the actinides), thus inviting a comparison with the
transition metal elements, discussed in Chapters 6- 9, with their varying
number of d electrons. However, as we shall see, such a comparison is not
particularly helpful, a situation which has contributed to an attitude
commonly encountered-that f electron systems are difficult to understand,
that the theory is difficult. It is hoped that it will be possible to demonstrate
in this chapter that this is not the case. Indeed, it is hoped to convince the
reader that a study off electron systems is not only of value in its own right
but that such a study helps in the understanding of d electron systems. It
does this by its concern with phenomena which also exist in d electron
systems but which are currently largely ignored when discussing them.
What are the difficulties with f electron systems? First, they involve f
orbitals and these are unfamiliar. Fortunately, they have already been met
in this book, in Section 7.3, where it was found useful to assess their relative
energies, the f orbital splittings, in an octahedral crystal field. Secondly,
spin- orbit coupling is important. Again, this phenomenon has already been
met (in Sections 8.4 and 9.5) but because we shall have need of a deeper
understanding, pictures of spin-orbit functions will be introduced. Thirdly,
crystal and ligand field effects are small, something which poses problems
when one is more familiar with molecules in which they are large. Finally,
coordination number and geometry are much more varied (and more
uncertain) than for d electron systems. Fortunately, the last two problems
tend to cancel each other out. When crystal field effects are small, knowledge
of the detailed ligand arrangement becomes less important.
Introduction 1 239
4f
',
~
r-
_:
' .
the 4f (see Fig. 11.1). Moving across the lanthanide series, the addition of a
4f electron to balance each increase in positive charge on the nucleus is not
sufficient to prevent electrons in the 5d and 6s (and, for that matter, the 5s
and 5p) orbitals feeling the increased nuclear charge.2 The effect on the 5p,
5d and all other outer electrons is a progressive orbital contraction and, with
it, a decrease in the ionic radius. Although the data are incomplete, it is clear
that the actinides exhibit an actinide contraction which closely parallels the
lanthanide contraction.
A study of the electric spark discharge spectra of the elements (and, so,
of the electronic energy levels) shows that the ground states of most atoms
of the lanthanide elements have an outer-shell electronic configuration of
the form ... 4f"6s 2 , although some have ... 4f"- 1 5d 1 6s 2 In the trivalent ions,
also seen in the arc spectra, the two s electrons and one other are lost; the
other one being an outer d electron if one were present in the atom. The
electronic ground states of the trivalent lanthanides therefore have filled 5s
and 5p shells, empty 5d and 6s and a number of 4f electrons which varies
from none in lanthanum(III) through to 14 in lutecium(III). These configura-
tions are detailed in Table 11.1. Following the lanthanides, the trivalent
actinide ions contain a variable number of 5f electrons, no 6d and no 7s.
The actual electronic configurations of the ground states of the atoms
and trivalent ions are given in Table 11.2.
Table 11.1 The electronic configurations of the neutral atoms and trivalent lanthanide
ions. Core electrons have been omitted
Table 11.2 The electron configuration of the neutral atoms and trivalent actinide
ions. Core electrons have been omitted
complete labels. The reader may have noted the use of the phrase cubic set
in the preceding sentence. This is because the f orbitals of Fig. 11.2 are only
appropriate for cubic molecules (in practice, this means molecules with Oh
or Td symmetries or related point groups-molecules with symmetries such
that the x, y and z axes are symmetry-related). For non-cubic geometries
one uses a different set of f orbitals. Actually, the situation is not all that
unfamiliar. Consider a d., orbital in an octahedron. It is so normal to choose
the z axis to coincide with a fourfold axis that we seldom consider any
242 1 f electron systems: the lanthanides and actinides
~
X ~y
f" = f, (2z' _ 3 x, _ 3y,1; similar are fx3 = fx (2x2 _ 3Y' _ 3,,1and fy 3 = fy (2Y' _ 3z2 _ 3,121
~y
X~ ~y
along a fourfold axis in the undistorted octahedron, although they look the
same. In fact, the dz, (threefold axis) is a mixture of the t 2 • set, dxy>
d,z and dzx (fourfold axis). For dz, the existence of two choices of axes does
not mean a change in shape; for the f orbitals a change in axis set does mean
a change in shape.
There is another way of looking at this. In Oh symmetry the f orbitals
transform as A 2 • + T 1• + T2 •. Reduce the symmetry by a distortion (com-
pression or elongation, for example) along a fourfold axis so that the
symmetry is now D4 h. The f orbitals now transform as (B 1.) + (A 2 • +E.)+
(B 2 • +E.), where the brackets correspond, in order, to the symmetries in Oh
given above. It can be seen that two different sets of f orbitals have E.
symmetry. Because they have the same symmetry they can mix and it proves
convenient to let them do this and to work with combinations of them. As
a result, some differently shaped f orbitals arise.4 The members of this
so-called general set of f orbitals are shown in Fig. 11.3, where both their
abbreviated and complete labels are given.
The lanthanide and actinide trivalent ions commonly occur with high
coordination numbers, typically 7, 8, or 9 but up to 12, in low-symmetry
geometries. Simple theoretical models, such as the extended Hiickel and
angular overlap models (Sections 10.5 and 10.6) have therefore been applied
to them since symmetry alone does not give much insight. In such calculations
the general set of f orbitals would be used. It is perhaps appropriate to
comment that if the bonding in lanthanide and actinide complexes were
entirely ionic then it would be nonsense to apply such models to them-they
depend on the existence of a (covalent) overlap between the metal and ligand
orbitals. The fact that the f electrons in these compounds have properties
which are very close to those of the isolated M 3 + ions should not lead one
to conclude that the same is true of all other electrons.
most directly concern ls electrons because s orbitals do not contain any node at the nucleus.
The Is electrons are distinguished because they may be very close to the nucleus and only avoid
capture because of their high speed, which for the heavier elements approaches the speed of
light. This, relativistically, increases the mass of the Is electrons which in turn means a smaller
orbital (in the Schrodinger model of H-like atoms the most probable distance of a Is electron
away from the nucleus is inversely proportional to the electron mass; the essentials carry over
into more complicated atoms). So, the ls electrons screen the nucleus a bit more effectively
than expected on a non-relativistic model. Seeing a smaller positive charge on the nucleus, f
and d electrons occupy orbitals which are both larger and have lower ionization potentials
than we might have expected, thus increasing their availability for bonding. In the context of
the present chapter this is particularly important in the context of the f orbitals. Although the
primary relativistic effect concerns the ls electrons, all s orbitals have to remain orthogonal to
each other. So, if the ls contracts, so too must all others orbitals, to maintain orthogonality.
There therefore is an effect on outer s electrons also, an effect which seems particularly
pronounced for the elements Pt, Au and Hg.
244 1 f electron systems: the lanthanides and actinides
z
fxz2 = fx (4z2- x2- y2)•
FIJI. 11.3 The general set off orbitals together fyz, = fy 14,, -x' _Y'l is similar
with their shortened and detailed cartesian
angular forms.
y
z
X
fxyz
actinides. This means that they are rather insensitive to their molecular
environment. So, the crystal field splittings of f orbitals are about one-
hundredth of those of d orbitals. That is, in the language of crystal field
theory, all lanthanide and actinide complexes are weak field, high spin.
Similarly, the f-f electronic spectra of the complexes are very similar to those
of the free ions as seen in arc spectra. The fact that the spectra of, nearly,
isolated ions can be seen in crystalline materials has led to extensive studies
aimed at their understanding. We will return to these spectra later in this
chapter; it is the purpose of the present section to begin to assemble the
background which will make such a discussion possible.
Electronic structure of the lanthanide and actinide ions I 245
6 Note that we use the terms static and rotating charge densities (or orbitals) in preference
to the more conventional names real and complex because they are felt to be easier to understand
by most readers.
246 1 f electron systems: the lanthanides and actinides
~:} 1 node
fxz2 is shown; fyz2 is obtained by
an anticlockwise rotation of go•.
X
fz3. 0 node
Table 11.3 Trivalent lanthanide ion ground state characteristics. For all of these ions
the f electrons are the ones that determine the (many electron) term of the ground
state of the ion
La"' fO none 0 ls
eelu fl 3 3 2F
Pr111 f2 3,2 5 3H
Nd 111 f3 3,2,1 6 4/
Pmul f4 3,2, 1,0 6 5/
Sm111 f5 3, 2, 1, 0, -1 5 sH
Eu 111 t• 3, 2, 1, 0, -1, -2 3 7F
Gd 111 f1 3, 2, 1, 0, -1, -2, -3 0 •s
Thill t• 3. 3 7F
Dy"' f9 3,2 5 sH
Ho111 flO 3,2,1 6 5/
E~" f11 3,2, 1,0 6 4/
Tmul f12 3, 2, 1, 0, -1 5 3H
Yblll f13 3, 2, 1, 0, -1, -2 3 2F
Lu 111 f14 3, 2, 1, 0, -1, -2, -3 0 ls
a From Tb111 onwards, for simplicity, the half.filled shell that is also present is not detailed.
are the effects of spin-orbit coupling and it is these that we now consider.
As will be seen, the effects of spin-orbit coupling are most simply covered
using rotating orbitals, although later an attempt will be made to give their
static equivalents.
to a 2 F term. Bringing these angular momenta together, there are only two
possible resultants, 3 ± f, i.e. i and 1. So, spin-orbit coupling splits the 2 F
ground state term of cerium(lll), an f 1 ion, into two levels, 2 F112 'doublet
F seven halves' and 2 F512 'doublet F five halves'. The subscripts are the
two possible values of j, where j denotes a sum of spin and orbital angular
momenta, j = I + s. It is perhaps helpful to think of this splitting resulting
from the interaction of two magnets (which, for simplicity can be thought
of as bar magnets), side by side. The lowest energy arrangement is that in
which the north pole of one magnet is next to the south pole of the other.
So, here, the 2 F 512 is the more stable level. It is a general pattern for atoms
that an angular momentum ofm implies a (2m+ I) degeneracy. For instance,
spin functions with s = f are doubly degenerate; s functions with angular
momentum of 0 are singly degenerate; p functions with an angular momentum
of I are threefold degenerate and so on. So, here the j = i functions are
eightfold degenerate and the j = t are sixfold, a total of I4 functions. This
is just the number implied by the parent symbol, 2 F; a 2 x 7 fold degeneracy.
The reader may reasonably complain that we have treated the two angular
momenta in different fashion in the above development. We have taken the
maximum orbital angular momentum and added to it, in turn, each component
of the spin angular momentum. It is by no means self-evident that this
procedure is valid; unfortunately it would break the continuity of the
discussion too much to present a justification for it here but one is given in
Appendix I2. Using this procedure we arrive at the spin-orbit split levels
given in Table Il.4.
The final question to be answered for each configuration is which of the
spin-orbit levels becomes the ground state? An answer has already been
given for the f 1 case, where it was concluded that the smallest value of j
becomes the ground state (an analogy was drawn with a pair of magnets
being most stable when the north pole of one is next to the south of the
other). This result is general for the first half of the lanthanide series, the
lowest value of j is the most stable. Just as was found for d electron
systems, so too for f electron systems; for more than half-filled shells it is
simplest to work in terms of holes rather than electrons. Now, if in an orbital
there is a single electron with spin up, then if we describe this situation using
the hole formalism then we have to talk of a hole with spin down (this is
the spin of the electron which is absent). It follows that the spin-orbit state
which is most stable when talking about electrons will be the least stable
Table 11.4 Spin-orbit levels arising from f electron ground state Russell-Saunders
terms
Table 11.5 Ground and low-lying electronic levels of the lanthanides and actinides
Note the symmetry in this table (compare the first level listed for an f" ion with the last listed for the 114 -").
This symmetry is detailed in the text.
when talking about holes. So, for the second half of the lanthanide series
it is the spin-orbit state with the highest j value which becomes the ground
state. This is not the easiest of arguments to follow and so an example is
given in some detail in Appendix 14. For the f1, the half-filled shell case,
where the orbital angular momentum contributions sum to zero (although
the spins most certainly do not) there is no spin-orbit splitting.
We are now in a position not only to detail the ground states of all the
lanthanide and actinide trivalent ions but also to give some of the low-lying
excited states when these result from spin-orbit splitting. Excitation to these
low-lying levels corresponds to energies in the infrared or near infrared
regions of the spectrum for the lanthanides, the splittings resulting from
spin-orbit coupling being of the order of 1000 em -I. These states are detailed
in Table 11.5. This Table has an underlying symmetry, perhaps most readily
revealed if the 8 S, f 7 , entry is moved sightly to the right and then regarded
as an approximate centre of symmetry.
dxy
Py
Table11.6
Property Function
-- --
g,(x'-r> dxy Spin!
--
Py
-- -- --
1
4 2 ~
momentum
-c2 -c2
Number of angular 4 2 ?
-- -- --
nodes (Rg. 11.5)
x2 x2
Smallest angle between 45' 90' 180' ?
nodes (Fig. 11.5)
The number, of course, is one-half and the angle three-hundred and sixty
degrees. This seems nonsense. What meaning can be given to half a node
and what does it mean to have three-hundred and sixty degrees between
complete nodes? Actually, both are entirely sensible. The essential step is to
change the identification of the identity operation (the E at the head of
character tables, the 'leave alone' operation). This operation can be equated
with a rotation of 360° (although this equation is not often explicitly stated).
Instead, we now equate it with a rotation of 720°. It is easy enough to make
this change for the functions shown in Fig. 11.5. This is done in the top row
of Fig. 11.6 which shows the angles oo through to 360°, compressed into the
(actual) region oo to 180°. The diagrams are completed in the middle row,
where the angular pattern from 360° through to 720° simply duplicates that
for oo to 360°. In this row the static counterpart of the angular momentum
! case has been included, which can be seen to comply with the requirements
imposed on it by the completed Table 11.6. It is evident that these diagrams
are becoming rather congested and so at the bottom of Fig. 11.6 are given
simplified pictures, drawn in a pattern that will be followed for the remainder
of this section.
Although a picture of an angular momentum = ! function (more accurately,
a 1!1 function) has been obtained, it still seems somewhat artificial. It can be
given more physical reality using a Mobius strip. The top of Fig. 11.7 shows
a long strip of paper, creased so that it can easily be bent (away from the
viewer) along its central axis (indicated by the arrows). Lobes are drawn
with phases as shown, the lobes being terminated by nodes at the edge of
the paper. Suppose that the back of the strip of paper is coated with a contact
adhesive, so that as soon as the strip is folded back it sticks to itself. Join
the ends of the strip together but in doing so twist the strip so that the back
comes to the front. The result is shown at the bottom of Fig. 11.7. Starting
at any point on this Mobius strip it is necessary to go round twice,
akin to rotating by 720°, before regaining the starting point. The nodal
pattern encountered in traversing the Mobius strip is just that shown
in Fig. 11.6 for the angular momentum =!function. Having pictured a static
angular momentum ! function it is not difficult to extend the approach to
other half integer functions and this is done in Fig. 11.8, which also shows
the corresponding unfolded, unstuck, Mobius strips. A few points remain
to complete this section. Just as in all other cases except the angular
momentum= 0 case, all functions appear in pairs (only one member of each
i')
U1
i')
..
~
~
540
a
:::1
540 540 ..
~CD
3
!'!
720 :Et
:CD
360 I === :::w:::: .......:: • X 360 I :::== 7t\: ::::::::::::::: • X 360 X
0 :fir
·a
::r
..
:::1
a:
CD
180 180 180 :
. :::1
...
Q.
..
~
:;·
CD
.a:
720
360 I 3> <: I • X 360 ::;:;. )I(< • X 360 >X< I X 360 X
=-= • 0
+ 720
I T::-:=>Y<C .· I~-;- X -+----:;;,I(E----+-- -x -+-------3~---+--x X
0
+ -
Fig. 11.6 The identity operation~ 720' representation of the orbitals of Fig. 11.5, together with the spin~~ function.
Spin-orbit coupling in pictures 1 253
............................................................................................................................................................................................
720 720
360 0 360 0
-
The two static functions corresponding
® ®*.
to angular momentum ~ ± ~
~. 11.9 The pairs of static (real) functions
corresponding to angular momentum~ ±~
and±~.
720 720
'"' 0 ""' - 0
pair is shown in Fig. 11.5). So, too, for the angular momentum = f, ~, ... ,
and so on cases. Figure 11.9 shows both members of the f and ~ pairs.
Finally, this section has been entirely concerned with half-integer values of
the angular momentum. Spin-orbit functions may also have integer values
of angular momentum; for these the pictures look the same as for the familiar
orbitals with the same angular momentum, although their meaning, of
course, is somewhat different.
's 'So
3p lpo 1p1 tp2
'o 'o,
3F 3p2 3F3 3F4
'G 'G•
,,
3H 3H4 3Hs 3H6
'16
to give a 1 I term. This 1 I term and the 3 H are the start of a neat pattern
based, ultimately, on the pattern that the angular momenta of the two
electrons can either add or subtract, 8 but never go negative. The greatest
orbital angular momentum they can have as a pair is 6 and the smallest 0,
seven values in all. Write down, in order, all the term symbols corresponding
to all values of the orbital angular momenta from 0 to 6. We obtain
S+P+O+F+G+H+I
Now add the only two possible spin labels, I and 3, alternately, in such a
way as to include 3 H and 1I. We obtain:
15 + 3p + 1 0 + 3F + 1G + 3H + 1/
This is a complete list of the terms that arise from the f 2 configuration. A
check on the correctness of the result can be obtained by counting
wavefunctions. For the f2 configuration the first electron can be inserted in
any one of 14 ways (we have seven orbitals, the spin can be either up or
down); the second electron can be fed in in any one of 13 ways (it cannot
be in the same orbital as the first and also have the same spin). Because the
electrons are indistinguishable we have counted every possibility twice-so,
it and t i have been counted as different. It follows that the total number
of wavefunctions is 14 x 13/2 = 91. This number, 91, is the same as the
number of functions contained in the terms that we generated above:
15 + 3p + 1 0 + 3F + 1G + 3H + 1/
1 + 9 + 5 + 21 + 9 + 33 + 13 ~ 91
It is now time to introduce spin-orbit coupling and this is done using the
procedure that has already been described in Section 11.4 and Appendix 12.
We simply add each component of the spin angular momentum to the orbital
angular momentum. The result is given in Table 11.7, where the final levels
are stacked according to their total angular momenta, according to their j
values, this being given by the subscripts. Now comes an important point;
8 The reader may reasonably object-what about cases such as -3 combined with -3, this
gives us - 6? In fact such cases are included in the discussion in the text. So, for example, the
I term, from the 1 /, comprises 13 (2L + 1; here L ~ 6) different functions, differing from each
other because the z component of the orbital angular momentum spans the 13 values from + 6
through to -6. We do not have freedom to use the -6 function a second time; the angular
momenta which are reflected in term symbols such as S. P, D, F, , are never negative.
256 1 f electron systems: the lanthanides and actinides
~~~ 1 . "
"
/
----1so
45000 s //
I
Fig. 11.10 The electronic energy level diagram I
for Pr111 (4t2). A typical crystal field splitting is I
shown at the right; its effects are much smaller
than those of spin-orbit coupling. As with 20000
Tanabe-Sugano diagrams, the horizontal axis is
taken as the ground state (3 H4 ). This enables ~
'E
__....---10:!
the effects of interactions between levels with
the same total angular momentum to be more " 15000
_1._
clearly seen. Note, for instance, the relative
upward displacements of levels with the j
quantum number 4 (3 H4 , 3 F4 , 1 G4 ).
10000
T
Crystal
field
....-----3H5
0 L-3..:H.:....""""":;::":::/_"_ _ _ _ _ _ _3:_:H4i.__ _ __
Electron repulsion Spin-orbit coupling
levels which are in the same column can interact together because of
spin-orbit coupling; they have the same values of j. In terms of the pictorial
representation of the previous section, the spin orbitals that interact have
the same number of nodal planes containing the z axis; they have the same
symmetry. As is evident from Table 11.7, some levels stand in isolation and
so remain unchanged by spin-orbit coupling (' P1 for instance); for others,
mixing is between functions with the same spin ('S0 and 1 P0 ). However, the
fact that the ground state, 3 H 4 , mixes with 1 G4 ensures that all of the f -> f
electronic transitions of an f 2 ion such as praseodymium(III) are of mixed
spin character (a mixture of spin triplet and singlet), emphasizing the fact
that because of spin-orbit coupling it is not useful to attempt to talk of spin
on its own.
The above discussion has been relatively superficial but has covered all
of the important points. In Fig. 11.10 is given an actual energy level diagram,
which is not relatively superficial, and which shows how the primary
separation between levels is due to electron repulsion, a topic which was
considered in some detail in Chapter 8, in particular. Note the way that
spin-orbit coupling changes the levels. So, although the 3 H 4 is the most
stable level originating in the 3 H, 3 F4 is the least stable of those coming from
3 F; the splitting between 3 F and 1 G is much greater than that between 3 F
4 4
and 1 G. Several other similar patterns can be discerned in Fig. 11.1 0.
This completes our discussion of the excited states off electron systems.
For completeness, however, Table 11.8 gives a more complete list of the
Russell-Saunders terms arising from all f" configurations. It does not include
the effects of spin-orbit coupling, which have to be worked out for each case
individually. The reader may find it a helpful exercise to select a configuration
from Table I 1.8 and use it to construct a table similar to Table 11.7.
Electronic spectra off electron systems 1 257
Of these three types of transition, it is the first which has received, by far,
the greatest study, even though they are the weakest bands. To put this in
perspective, Fig. 11.11 shows the electronic spectra of aqueous Ce 111 , 4f 1, and
Pr 111 , 4f2. Only in the latter are f-> f transitions evident; this is because for
the Ce 111 the only transition is between the two spin-orbit components of
258 1 f electron systems: the lanthanides and actinides
~-L--~
of Pr111 , an f 2 ion, is more characteristic of
lanthanides, showing both charge transfer and
f - f transitions (sharp). The involvement of
f-f electron repulsion moves some of these
transitions into the visible region of the
spectrum. The relevant energy level diagram
is given in Fig. 11.10.
40 24 20 16 10 5 0
cm-1 x 103
(b) Prm f 2
fields are about twice as great for the actinides as for the lanthanides so that
it is less acceptable to ignore them in the way that we shall do.
When a lanthanide ion is at a centre of symmetry, in the hydrated
pcrchlorates where the lanthanide is octahedrally surrounded by water
molecules, for instance, the f--+ f transitions are an order of magnitude
weaker than in low-symmetry complexes, indicating that one intensity-
generating mechanism is the mixing off with other orbitals (presumably d,
because this would lead to an allowed f--+ d component in the transition)
by the low-symmetry crystal field, a mechanism analogous to that discussed
for low-symmetry transition metal complexes in Chapter 8 (where d-p
mixing was invoked). Just as for transition metal complexes, too, it seems
that the mechanism by which the transitions in centrosymmetric complexes
gain intensity is through the dynamic distortions caused by molecular
vibrations (for transition metal complexes this mechanism was discussed in
Section 8.7). Support for this explanation comes from experiments such as
those in which a low concentration of a lanthanide ion is doped into
crystalline Cs 2 NaYC1 6 . The yttrium is an f 0 ion and octahedrally surrounded
by CJ- ions. Particularly with low-temperature samples, it is possible to
observe mixed electronic+ vibrational (vibronic) transitions for the doped
lanthanide ion, consistent with a vibrational mechanism for generating the
intensity.
The point in our discussion has now been reached at which the parallel
with d --+ d transitions is little help; new mechanisms for generating the band
intensities have to be introduced. The reason for introducing such new
mechanisms is that those given above prove inadequate when really tested.
For instance, detailed evidence that a vibrational mechanism is involved in
making f --+ f transitions weakly allowed in centrosymmetric complexes has
just been presented. But what of the band origin which is seen under
high-resolution conditions? It has no vibrational component and so no
vibrational explanation can be given for its appearance. The explanation
which is usually given for it is that it is magnetic dipole allowed, an
explanation which itself calls for an explanation! In the Maxwell model, a
monochromatic polarized beam of light consists of two vectors, an electric
vector and a magnetic vector. These are mutually perpendicular to each
other and both are perpendicular to the direction of propagation of the light
wave. Normally, attention is confined to the electric vector because calcula-
tions indicate that the interaction between the magnetic vector and a
molecule is much weaker than that involving the electric vector. 9 Magnetic
vectors, and magnetic fields, cause circular displacements of charge (think of
the path followed by a charged particle in a conventional mass spectrometer).
Such rotations of charge, like all rotations, are symmetric with respect to
inversion in a centre of symmetry. This is in contrast to an electric vector,
and the linear charge displacements to which it gives rise; they are anti-
symmetric with respect to inversion in a centre of symmetry (Fig. 11.12). So,
magnetic-dipole-allowed transitions are g in nature whereas electric-dipole-
allowed transitions are u. Now, f--+ f transitions are u --+ u and u x u = g.
So, although f --+ f transitions are electric-dipole-forbidden they are magnetic-
9 Not that it is zero-all magnetic resonance measurements rely on the fact that the
Invert in a centre
of symmetry
Q Invert in a centre
of symmetry
Table 11.10 Some f --+ f hypersensitive transitions. In this table it has been recognized
that the Eu 111 7 F1 state is low lying and is thermally populated at room temperature so
that the hypersensitive transition could involve it as the ground state
Of course, the only light wave which is relevant here is that which has an
energy corresponding to the energy of the hypersensitive transition observed.
On this model, the sensitivity of hypersensitive transitions to different
environments arises from the different polarizabilities of the different ligand
environments.
A second mechanism which has been suggested to explain hypersensitive
transitions applies only to low-symmetry molecules, and so, low-symmetry
crystal fields. The mechanism postulates that one of the dipoles is inherent
in the ligand arrangement; the overall ligand environment is so distorted
that it is dipolar. The second dipole required for a quadrupole again comes
from the incident light wave. On this model, different hypersensitivities arise
because of different intrinsic dipoles in the ligand arrangements. This second
model can be modified to cover the case of high-symmetry arrangements,
arrangements which are centrosymmetric and so cannot be intrinsically
dipolar, for example. This modification supposes that the high-symmetry
molecule is distorted by a vibration, such as those vibrations used to explain
the intensities of d -+ d transitions in Section 8.7. These, for an octahedral
complex, are vibrations which destroy the centre of symmetry. Examples,
again for an octahedral complex, are vibrations of T 1 u or T 2 u symmetries.
The vibrationally distorted molecule is then non-centric, at least whilst the
distortion persists (although only for a T 1 " vibration would it have a-
transient-dipole). This, then, is the source of one dipole; again, the second
is that originating in the light wave itself. The hypersensitivity differences
would, on this model, originate in the different vibrational properties of the
different species.
We have then three different explanations for the phenomena of hyper-
sensitive transitions. In general, they are not necessarily mutually exclusive
although the relative importance of each would surely vary from complex
to complex (one could not invoke the second for a centrosymmetric species,
for instance). Unfortunately, although hypersensitive transitions show that
a lanthanide ion is sensitive to its surroundings, until more is known about
the mechanism of hypersensitivity it is not possible to use the phenomenon
to learn more about this sensitivity. One final point, also puzzling: the
actinide ions, despite the greater importance of ligand fields for these ions,
do not seem to exhibit hypersensitive transitions. Even so, some of their
transitions show small, but undoubtedly real, band intensity changes with
change in ligand environment.
ca. 1000 and 2000 em-\ respectively). Although either type of band may be
the lower lying, in the lanthanides the charge-transfer tends to occur at the
lower energy when addition of an electron to an f-shellleads to a half or full
electron shell. The f 6 ion europium(III) and the f 13 ion ytterbium(III) are
species for which this pattern holds. The converse pattern is also general.
That is, when an f-. d transition (a transition which means that the number
off electrons is reduced by one) leads to an empty or half-filled shell then
these f -. d transitions are the lower. The f 1 ion cerium(II) and the f 8
terbium(III) provide examples of this pattern. It seems that the f 2 ion
praseodymium(III) is another example, although this could not have been
predicted. When the lanthanide or actinide is one for which more than one
valence state may be studied, then it is relatively easy to distinguish between
electron-transfer and f-. d bands. As the valence state increases, f-. d bands
move to higher energy with increase in valence state whereas the-ligand to
metal-electron-transfer move to lower. Of course, the spectra of two
differently charged ions will be far from identical but, nonetheless, bands
which appear similar in the two spectra show these relationships. Another
distinction, within a given valence state, is based on the fact that, as one
would expect, the d orbitals are much more sensitive to the ligand environ-
ment than are the f, so that the energies of the f -. d bands are dependent
on coordination number whilst the charge-transfer are not. Conversely, for
a fixed coordination number, the energies of the f -. d bands are less sensitive
to change in ligand than are the corresponding charge-transfer bands.
All of these criteria more-or-less follow simple common sense ideas about
the characteristics of the two types of transition. Together, they enable
distinctions to be drawn in most cases. Although the detailed study of the
bands is in its infancy, it is interesting to note that f-. d transition energies
seem to follow a spectrochemical series, just as do d -. d in the transition
metal ions. So, the following 5f-. 6d transition energies, all x 10 3 em -I,
have been reported for complexes of um and other similar ions:
Ligand: 1- < sr- < cl- < so~- < H2 0 < F-
v(f-. d): 13 ca. 17 ca. 19 ca. 22 25 ca. 25
the most efficient emitters; not surprisingly, their observed radiative lifetimes
are found to be anything up to a factor of I 0 3 greater than those of the ions
at either end of the series. Much more surprisingly, the energies of the excited
states of the ions in the centre of the series are higher than those for the end
members by a factor of two or three-as has been seen from Table 11.4.
With the exception of gadolinium( III), the ions at the centre of the series have
more f electron excited states and so the energy levels are more widely spread.
One might have expected that the more energy there is to lose, the more
readily it would be lost-but, clearly, this is not the case. It seems clear that
a key part in this phenomenon is played by the solvent (water). So, if D 2 0
is used as the solvent in place of H 2 0, the observed excited state lifetimes
are all increased by an order of magnitude. The explanation seems clear; an
important non-radiative deactivation process is one in which the energy in
the electronic excited state is transferred to an overtone of the v(Q-H)
stretching vibration of the solvent water. A lower overtone level is needed
for the deactivation of the end members of the lanthanide series than for the
central members. Because the frequency of the v(Q-D) vibration is only
about three-quarters of that of the v(O-H), a much higher overtone is needed
in D 2 0 than H 2 0. Remembering that the transitions involved are f--> f, it
is evident that this deactivation mechanism indicates something that we have
met several times before, that f--> f transitions are not totally insulated from
their environment.
Amongst the most efficient of current commercial phosphors are those
based on the red emission of europium(III), the green emission of terbium(III)
and the blue emission of europium(III). These ions are at the centre of the
lanthanide series so that it seems that the phenomena found in aqueous
solution extend to the solid state. In addition, gadolinium(III) may well be
a component of a phosphor but this is because of its excited state lifetime,
not because of any emission. The actual matrix into which the lanthanide
ions are incorporated in a phosphor are oxide or glass ceramic-forming
oxyanions-borates, silicates, aluminates and tungstates. The choice is
important because one step in the emission process is transfer of the absorbed
ultraviolet radiation (absorbed by a carefully chosen impurity species,
commercially called a sensitizer) through the crystalline lattice to the emitter.
An ordered lattice facilitates this energy transfer but may simply serve to
allow transfer to an (inadvertent) impurity ion which provides a mechanism
for deactivation without emission. The degree of order of the host lattice has
to be optimized, as too does the concentration of the chosen emitter.
Incorporation of an ion such as gadolinium(III) at quite a high concentration
(up to 20% of total cations) in the lattice is found to facilitate long-range
energy transfer. The current understanding of the processes involved may
best be illustrated by an example. Consider a borate lattice containing a high
concentration of gadolinium, GdB 3 0 6 • A typical sensitizer is bismuth(III)
and europium(III) will be chosen as emitter. Current notation would write
this composition as GdB 3 0 6 :BiEu. The sequence of events seems to be:
Ultraviolet
light Gd(lli) __, energy transfer Eu(ill) --> emission
L T L T
Bi (Ill) __, excitation Gd(lll) --> energy transfer
Magnetism of lanthanide and actinide ions 1 265
iieff = J4S(S + 1)
It will therefore not be surprising that, essentially by replacing S by J in the
derivation, the J-only model gives rise to the equation
1-'eff OC JJ(J + 1)
The equality sign in the spin-only equation has been replaced by a
proportionality because, whereas there can be no ambiguity about how
a given value of S arises-one simply has to count the number of unpaired
electrons and divide by two-as has been seen, a given J value can arise
from a variety of different spin and orbital components. Spin magnets and
orbital magnets are not immediately interchangeable and so the constant of
proportionality in the J-only equation, denoted g, has to reflect the particular
mix involved. The equation for g is
3 S(S + 1) - L(L + 1)
g= 2+ 2J(J + 1)
the ground state by the magnetic field, discussed in Section 9.6) is expected
and has, in fact, been included in the calculated values in Table 11.11.
Secondly, the ions europium(III) and samarium(III) do not give good
agreement. The reason for this is known. These two ions possess very
low-lying excited states which are so low in energy that they are thermally
populated. The J values for these low-lying states differ from those of the
ground states and so the J-only equation above has to be modified to take
account of this. When this correction is made, good agreement with
experiment is obtained.
It is evident from Table 11.11 that several of the lanthanides have rather
high values of Jl.ecc· Physically, this means that when salts of these ions are
placed in a strong magnetic field, they become slightly warm. The system is
stabilized, the salts are attracted into the magnetic field, and the energy of
stabilization has to go somewhere; it appears as heat. At room temperature
this effect is small but at low temperatures it becomes rather significant. This
has led to the use of gadolinium salts, especially Gd 2 (S0 4 h ·8H 2 0, and salts
of dysprosium for obtaining low temperatures. The salts are placed in a
container, cooled with liquid helium and a strong magnetic field applied. As
has just been seen, heat is evolved. When the salts have cooled down to
liquid helium temperature again, the magnetic field is switched off. Cooling-
adiabatic demagnetization-occurs as the salts lose their magnetic orientation.
There is an alternation of natural abundances of the lanthanides, starting
with lanthanum (low), 11 cerium (high), so that both gadolinium and
dysprosium are highs, explaining their selection for this application-
gadolinium is not otherwise, from Table 11.11, the most obvious choice.
Clearly, the total cooling effect depends on having a relatively large amount
of the lanthanide salt available.
11 Low and high here are relative to adjacent elements only. So, lanthanum is four or five
times more abundant than either gadolinium or dysprosium.
f orbital involvement in bonding I 267
12.1 Introduction
In preceding chapters the theories underlying some of the methods which
have been used to characterize coordination compounds have been discussed
in some detail. The present chapter provides a less detailed survey of other
methods, stretching from indicators of complex formation through to some
which provide insights into physical structure, others into electronic structure,
yet others into the forces between atoms and molecules.
If one is solely concerned with the question of whether or not a complex
is formed in a particular system, relatively crude methods often suffice-
although, historically, such methods have played an important part in the
development of the subject. The evolution of heat, the crystallization of a
product, a change in chemical properties- the failure to undergo a charac-
teristic reaction, for instance- are all simple but useful indicators. More
sensitive methods often involve the study of some physical property of the
system as a function of its composition and some of these were outlined in
Chapter 5. Other examples of methods used for non-transition metal
complexes are the measurement of colligative properties such as vapour
pressure, boiling point or freezing point. Less commonly, measurements of
quantities such as viscosity and electrical conductivity have been employed.
Another electrical measurement which has been much used is that of pH
change. Most ligands may be protonated at their coordination site so that,
for example, corresponding to the NH 3 ligand there is the protonated species,
NH.(, the ammonium ion. This means that, if in the same solution there is
free ligand L, an acceptor species A (which could well be a metal ion), and
acidic protons- which for simplicity will be written as H +, there will be two
important equilibria involving L:
270 1 Other methods of studying coordination compounds
and
A+L~[AL]
but the bend only drops by ca. Scm- 1 In contrast, in the [Hg(SeN) 4 ] 2 -
anion, where the seN- anion is S-bonded, the v(e-N) mode rises to ca.
2100 em -•, the v(e-S) drops to ca. 710 em -• whilst the bend drops to ca.
450 em- 1 • It is obvious that these different patterns can be used to determine
the way that a seN- ligand is coordinated. Even more obvious would be
to look for the appearance of v(M-N) or v(M-S) features, but this is not
always easy if there are other ligands present with metal-ligand stretching
modes in the same spectral regions. This explains why there is a particular
interest in ligand-specific modes falling in otherwise relatively clear spectral
regions.
In favourable cases vibrational spectra can be used to determine the ligand
geometry around a central metal atom.2 Best known is the fact that a centre
of symmetry means that bands active in the infrared are not Raman active
and vice versa, although this is just one of many symmetry-spectra
relationships. Such relationships are not infallible. So, a band may exist but
be too weak to be observed separately from other bands ~n the same spectral
region. Alternatively, vibrational couplings may be small. So, if two vibrators
are uncoupled it is irrelevant that they happen to be related by a centre of
symmetry-a single band will appear, coincident in infrared and Raman.
Because of this element of uncertainty, vibrational methods of structure
determination are sometimes referred to as sporting methods-they are not
infallible-although for compounds for which crystals, and thus X-ray
crystallography, cannot readily be made available, they may be the only
methods which it is feasible to use.
As a typical example of the application of vibrational methods, consider
square planar complexes of Pt 11 of general formula PtL 2 X2 , where X is a
halogen and L is a polyatomic ligand. A study of the spectral activity of the
Pt-X stretching modes usually shows whether a particular complex is cis or
trans. If it is cis then the molecular geometry is approximately C 2 v and the
v(Pt-X) vibrations are of A 1 + B 1 symmetries so that both modes are both
infrared and Raman active. It follows that two v(Pt-X) bands are expected
in each spectrum, the bands in one spectrum coincident with those in the
other, to within experimental error. The trans isomer has approximate C 2 •
symmetry and the v(Pt-X) vibrations have A• + A. symmetries. Of these,
the former is Raman active and the latter is infrared. One band is predicted
in each spectrum, at a different frequency in each. The difference in the
spectral predictions between the two isomers is so great that it should be
possible to use the method even if the spectra are incomplete or less than
ideal. An application of these results is illustrated in Fig. 12.1, which shows
the Raman spectra of cis and trans [PtC1 2 (NH 3 h] in the v(Pt-Cl) and
v(Pt-N) stretching regions. The presence of two peaks in the v(Pt-N) region
is sufficient to establish the lower spectrum as that of the cis isomer, without
aid of the infrared spectrum, even though the v(Pt-Cl) only shows one (the
expert might well argue that the greater breadth of the v(Pt-Cl) peak in the
2 This has been exploited in transition metal carbonyl chemistry. Simple metal carbonyl
derivatives often have quite high symmetries and, equally important, a variety of possible
geometric arrangements. The vibrational coupling of v(C-0) groups is so great-and the
resulting bands so well separated-that simple group theoretical methods may often be used
to distinguish between them. For instance, an M(C0) 3 unit could have D31, C3 l, or Cs
symmetries. These could be distinguished because they give rise to one, two and three infrared
bands in the v(C-0) region, respectively.
272 1 Other methods of studying coordination compounds
(b)
2100 2050
'K--------------/( 'K--------------7(
'' '' '' ''
'' '' '' ''
''' ''
' '' ''
''
''
~--------------~ /--------------~
274 1 Other methods of studying coordination compounds
(a)
Fig. 12.6 (a) The in-phase vibrations of two
CO groups in an M(C0) 2 unit subtend a
resultant vector (shown dotted) of 2d cos(0/2).
(b) The out-of-phase vibrations of two CO
groups in an M(C0) 2 unit subtend a resultant
vector (shown dotted) of 2d sin(0/2). In both
cases the vibrations are shown and, to the light,
the corresponding vector sums.
. e
(b)
dSin 2
Resonance Raman spectroscopy 1 275
X 10
6v 1 2v1
?v1
01
~
!!l
i
i
~
Jv3+v1 2v3+V1 v3+V1 v1 lOv3 ev, 8v3 5v3 + v2 7v3 4v3 + v2, 6v3 V3+V2,JV3
. .T2vTr·· ••
·r Ii
I
I
~
g
I
~ cm-1
(b)
Fig. 12.6 (continued) (b) Crystalline [NBu 4 ][0s(C0) 2 Br.:J. The anion has D4h symmetry, the totally symmetric modes are: v1 , v(C=O); v2 , v(Os-C); ''••
v(Os-Br); reproduced with permission from F. H. Johannsen and W. Preetz, Z. Naturforsch (1977) 32b, 625. In both cases, although other modes can be
seen, the totally symmetric modes dominate the spectra.
Spectroscopic methods unique to optically active molecules 1 277
r
~
r
~
/ ~ '
~ ~ ~
>c::::> /
FJc. 12.10 The passage of a beam of linearly
polarized light through an optically active
molecule. Note the rotation of the plane of
x::=::>
c--===='
x::=::>
>
......______
~
~
c::=><
c ><
c::=><
'-.....,
~ ~
polarization and correlate this with the different
behaviour of the two circularly polarized
components INithin the molecule.
(
><==:: eX Optically
active
molecule
~ ex
280 1 Other methods of studying coordination compounds
c:
0
%0
<I)
.c
<(
(a)
4 The requirement for optical activity is that electronic charge displacement occurs along a
helical path, for then optical isomers will be associated wit h bands of opposite helical charge
displacement. Group theore ti cally, this means that transla tion along and rotation about an axis
transform isomorphically - as the same irreducible representation (and, if the irreducible
representation is degenerate, as the same component).
282 1 Other methods of studying coordination compounds
Table 12.1 Some of the nuclei which have been studied by NMR
spectroscopy; spin = ~ unless otherwise stated
Et
•omn-@-o-@:> '
I
0.000~ ~020 1 t ~t
Fig. 12.14 Nickel(ll) aminotroponeimineato
complex.
_ +
+ 0.040
-0.022 2
~a+
- - - - - - - - - a-
to excite the nucleus from a lower to an upper state by application of suitable
radio-frequency radiation. This is a classical description of the phenomenon,
(b) but the essentials are carried over into a quantum mechanical treatment. In
Fig. 12.11 A nucleus (represented by an
practice, the non-uniform electrostatic field is generated by the charge
ellipse) with a quadrupole moment (represented distribution around, but very close to, the nucleus. A more detailed
by the charges within the nucleus) in (a) a more discussion shows that any variation in field gradient is almost entirely due
stable and (b) a less stable orientation within
an applied electrical field gradient. to unequal occupancy of the p orbitals of the atom-so that if there is to
be a field gradient, the p orbitals cannot be symmetry-related. Clearly, this
is a phenomenon which is only applicable to atoms in a low-symmetry
environment (but, note carefully, this does not automatically mean a
low-symmetry complex). The most-studied of the nuclei which exhibit
quadrupole resonance spectra are 35 Cl and 37 Cl. The method is inherently
insensitive and the high concentrations of these isotopes in samples such as
solid K 2 [PtC1 6 ] is a great advantage. Only solids can be studied, anyhow,
because the molecular tumbling in a liquid or gas averages the effect to zero.
The sensitivity of the method has increased in recent years by the advent of
pulse techniques (such as those used in NMR) and also, in suitable cases,
by not looking at the NQR nucleus itself but, rather, at one which is
energetically coupled to it and which is NMR-active.
Other, currently more specialist but of potential wide applicability,
methods include the optical detection of quadrupole resonances-a sample
is laser-excited to an electronically excited state, the return to the ground
state is by phosphorescence; the intensity of the phosphorescence is sensitive
to whether or not concurrent microwave radiation matches an energy
separation in some quadrupole-split intermediate state. Yet another method
depends on correlations between successive f3 or y emissions from excited
quadrupolar nuclei (where the excitation can be achieved by suitable nuclear
bombardment). These do not exhaust the list of current developments-they
have been chosen to illustrate the wide front on which new techniques are
emerging. It is likely that because of these developments the future will see
a wider use of NQR spectroscopy. It is also likely that the interpretation of
the data will become more sophisticated. Traditionally, the experimental data
have been interpreted to give the percentage ionic character of a bond. This
is because, for example, in the Cl- ion all of the p orbitals are equally
occupied whilst in Cl 2 the u bond, if composed of p orbitals only, corresponds
to one electron in the p6 orbital of each chlorine atom, and so Cl- and Cl 2
differ in their resonant frequencies. Interpolation allows a value for the ionic
character of a Cl-M bond to be determined from the chlorine resonance
286 1 Other methods of studying coordination compounds
frequencies in Cl- and Cl 2. Some correction may be applied to allow for the
Table 12.2 Ionic character of the M-el
bond in [MCI 6 ] 2 -. Although these ions
fact that a pure chlorine p orbital may not be involved in the M-el bond.
have Oh symmetry, the symmetry at In this way, Table 12.2 was compiled. When there are two non-equivalent
the chloride is, at most, C4 v (adjacent NQR nuclei in the unit cell of a solid these give rise to separate resonances
cations may reduce this symmetry) which may be resolvable. In this way NQR spectroscopy gives structural
information. Both bromine, 79 Br and 81 Br, and iodine, 127 I, but not fluorine,
Species Ionic character of the
give NQR spectra, as too may 14 N, 55 Mn, 59 Co, 63 Cu, 65 Cu, 75 As, 121 Sb,
M-CI bond (%)
123Sb, 201 Hg, and 2o9Bi.
[PtCI 6 ]"- 44
[PdC16]2- 43
[lrCI 6 ] 2 - 47 12.5.3 Mossbauer spectroscopy
[0sCI6]2- 46
[ReCI 6] 2 - 45 Just as there are ground and excited states of atoms and molecules
[WCiel2- 43 (electronic, vibrational and the like), so too there exist both ground and
[SnCI 6] 2 - 66 excited states of nuclei-they have just been mentioned as sometimes
[TeCI 6 ] 2 - 68 providing a method of measuring NQR spectra. In decaying from an excited
[SeCI 6j"- 56
state a nucleus may emit light, just as an atom or molecule may. In the case
of nuclei, this light is of very short wavelength, it is y radiation. If this y
radiation falls on another, identical, nucleus it may be absorbed, leaving the
second nucleus in an excited state. Nuclei of any one element which are in
different chemical environments will have slightly different energy levels,
but the environment-induced changes are so small that it is possible to
compensate for them with a Doppler shift of the y radiation, achieved by
moving the emitting nucleus either towards or away from the absorber. In
Miissbauer spectroscopy the absorption of·; rays by the sample is recorded
as a function of the velocity of the source. Solid samples are used; the source
may be moved by attaching it to the diaphragm of a loudspeaker driven by
a suitable signal generator. The effect has been observed for relatively few
nuclei at the concentrations at which they occur in most coordination
compounds, of which 57 Fe and 119 Sn have been the most widely studied. A
more complete list is given in Table 12.3. The difference in absorption velocity
and that of a suitable reference standard is called the isomer (or chemical)
shift; it is denoted il and is usually expressed in units of mm s- 1 or em s- 1.
The chemical environment affects the nuclear energy levels through those
electrons which are in orbitals which allow them to make contact with the
nucleus. This means that only s electrons can directly affect isomer shifts
since for all other orbitals the nucleus is contained in a nodal plane (electrons
in p, d, or f orbitals can only influence isomer shifts through their incomplete
shielding of the nucleus, leading to a change in effective nuclear charge which
is felt by the s electrons). Examples of isomer shifts, relative to an iron foil
Table 12.3 Some nuclei which have standard, of 57 Fe in some iron compounds are given in Table 12.4. As this
been studied by Mossbauer spectroscopy
table illustrates, the isomer shift decreases with increase of negative charge
Isotope Natural abundance (%) and increases with increase of coordination number. Both of these general-
izations find application in the observation that the isomer shift of Fe(C0) 5
57 Fe
is greater than that of Fe( CO Ji-. If the same nucleus has two different
2.2
99 Ru 12.0
u9sn 8.6 chemical environments in a compound, these will normally give rise to
121Sb 57.3 separate resonances. In this way it has been shown that the iron atoms in
12sTe 7.0 insoluble Berlin blue, Fe 4[Fe(CN) 6h which is formed by the reaction of
Fem salts and [Fe(CN) 6] 4 - ions in aqueous solution, are not all equivalent
1271 100.0
12gXe 26.2
197Au 100.0
but retain their distinct oxidation states. Indeed, a particular use of Miissbauer
data has been to indicate the valence state of an atom-empirical parameters
Nuclear spectroscopies 1 287
are available which compensate for the effects of change of substitu~nt and
Table 12.4 Isomer shifts for some iron
coordination number. So, isomer shift data have been used to conclude that
compounds•
the n bonding ability of ligands decreases in the order
Compound Isomer shift
(mm s- 1 ) NO+ >CO> CW >SO~- > PPh 3 > N0:2 > NH 3
High spin Fe"' ca. 0.3-0.5 A complication arises in that if the resonant nucleus experiences a non-zero
FeF3 (00 ) 0.49 electrostatic field gradient, a quadrupole splitting of the resonance may be
FeCI 3 (00 ) 0.46 observed (nuclear quadrupoles were discussed in the previous section). In
[FeF4r (Td) 0.30
an octahedral environment, such as that in [Fe(CN) 6 ] 4 -, there is no field
Low spin Fe 111 gradient (the chlorine atoms considered in [PtC1 6 ] 2 - in the previous
[Fe(CN) 6 ] 3 - (00 ) -0.12 section were not in an octahedral environment although in an octahedral
High spin Fe 11 ca. 0.9-1.5 complex), but in [Fe(CN) 6 ] 3 - the extra hole in a t 29 orbital has the
FeF2 (0 0 ) 1.48 effect of introducing a very temperature-dependent quadrupole splitting,
FeCI 2 (00 ) 1.16 which presumably occurs through vibrational distortions of the octahedron.
FeBr2 (0 0 ) 1.12 For [Fe(CN) 5 N0] 2 -, where there is a built-in asymmetry, the quad-
[FeCI 2 (H 2 0)4] COo') 1.36 rupole splitting is almost temperature-independent. The [Fe(CN) 6 ] 4 - and
[FeC1 4] 2 - (Td) 0.90
[Fe(CN) 5 N0] 2 - ions are compared in Fig. 12.17. There is evidence that
Low spin Fe 11 quadrupole splittings can be influenced by relatively distant ions in the
[Fe(CN) 6 ] 4 - (00 ) -0.04 crystal lattice-change a counterion, for instance, and there is a small
a Isomer shifts show a slight temperature dependence change in quadrupole splitting.
which has been ignored in thiS compilation. Data
An excellent example of the use of Mi:issbauer spectroscopy in structure
from N. N. Greenwood and T. C. Gibb M6ssbauer
Spectroscopy Chapman & Hall, London, 1971. determination is provided by Fe 3 (C0) 12 . Although the dark green crystals
of this compound are easy to prepare-dissolve Fe(C0) 5 in aqueous alkali
to give the Fe(CO)~- anion and oxidize this with solid Mn02 to give
Fe 3 (C0) 12 -its structure was uncertain for over 30 yearss Although eight
structures, all incorrect, had been proposed for Fe 3 (C0) 12 , it was the ninth,
suggested on the basis of its Mi:issbauer spectrum, which eventually proved
to be correct. The spectrum is reproduced in Fig. 12.18 (peaks are downwards).
The most evident thing is that it corresponds, approximately, to three peaks
of equal intensity. It would be wrong to conclude that each corresponds to
a different type of iron atom. If the iron atoms were all different then they
could not all be in high symmetry environments and so quadrupole splittings
would be expected on most of the peaks. Given that the chemical nature of
the three iron atoms is so similar-and so similar isomer shifts arc to be
expected-the only reasonable interpretation of the spectrum is that the
outer two lines are the quadrupole-split components arising from a peak of
intensity two, centred at about the same position as the central peak (which
itself has but a small quadrupole splitting and is of intensity one). So, it
seems that there are two equivalent, low-symmetry, iron atoms and one of
high symmetry. The correct structure, since confirmed by X-ray studies-
which were made difficult by disorder in the crystal-is shown in Fig. 12.19.
The high-symmetry iron atom is on the right (it has but two types of bond,
I one to terminal CO ligands, the other to Fe atoms). The low-symmetry pair
I
are on the left. The X-ray crystallographic work showed the bridging CO
' - - - [Fe(CN)6l 4- ligands to be asymmetric, off-centre. This means that the low-symmetry iron
atoms are each involved in four different bonds, two to bridging CO
Shift ligands-one long, one short-one to the terminal CO ligands and one to
Fig. 12.17 The Milssbauer spectra of
[Fe(CN) 5 NO]'- (d 5 • ligand assymetry) and 5 The story makes fascinating reading-see 'The Tortuous Trail Towards the Truth' by
[Fe(CN\ 6 ] 4 - (d 6 , symmetric ligand 1ield). R. Desiderata and G. R. Dobson, J. Chern. Educ. (1982) 59, 752.
288 1 Other methods of studying coordination compounds
c:
0
~
0
<II
Fig. 12.18 The Mossbauer spectrum of ~
Fe3 (C0) 12 .
1.0 2.0
Velocity (mm s-1 )
the unique iron atom. A more recent study of the Miissbauer spectrum over
a temperature range has provided explanations for the asymmetries in the
peak-intensity patterns of Fig. 12.18 which are entirely in accord with the
accepted structure.
Finally, mention should be made of the fact that additional information
can be obtained from measurements made with samples in a magnetic field,
a magnetic field that can even be self-generated because of an inherent
magnetism within the sample-a point of particular relevance to Fe 111, d 5
complexes. In Fig. 12.20 is shown the quadrupolar splitting that occurs when
a 57 Fe nucleus is in a magnetic field, together with an energy level diagram
that both explains the observed spectrum and also allows the relevant
selection rules to be deduced.
3
2
./- + j.
t
t t I I
at
I
bl c dI e[ f
Fig. 12.20 The Mossbauer spectrum of 57 Fe in I I I I
a magnetic field (reproduced courtesy of
Prof. K. Burger). 1
2 -::. /
',
I I
-6 -2 -4 4 6 8
N N N N
FIC. 12.21 The EPR spectrum of originating in interactions between the unpaired electron(s) and nuclei with
5-chlorosalicylaldoxinecopper(ll). In this complex non-zero spin. So, copper has two naturally occurring isotopes, each of spin
there are two N atoms, each with spin 1,
coordinated to the copper, so that the resultant
l For each, we expect (2S + 1) = 4 nuclear levels. The EPR spectra of
N spin can be 2, 1, 0, -1 or -2. It follows copper(II), d 9 , complexes therefore show four peaks rather than one.
that a five-line peak structure is expected. The Actually, the two copper isotopes do not give precisely superimposed lines;
Cu spin of ~ gives rise to a four-line structure
(~. ~. -~. -~) and so the spectrum is
rather, two inter-related sets of four lines each, with intensities determined
interpreted as indicated. Clearly, the unpaired by the relative isotopic abundances, 70:30. The (single) four-line pattern will
electron interacts with both copper and nitrogen be seen at the very top of Fig. 12.21.
nuclei. This spectrum is presented in derivative
form, the normal presentation for EPR spectra.
The interpretation of an electron paramagnetic resonance spectrum may
In this presentation the fine structure is made involve all of the parameters introduced in our discussion of magnetic
most evident. For comparison, Fig. 12.22 is susceptibilities in Chapter 9 and Appendix 10-so that EPR measurements
presented in integrated form {reproduced
courtesy of Prof. K. Burger).
can give information on these parameters. Such measurements have the
advantage that they give the individual components of the g tensor, whereas
magnetic susceptibility measurements normally lead only to an average g
value. In suitable cases, individual g values can also be obtained from EPR
measurements on samples in solution. There are problems, however. EPR
spectra can usually only be observed for ions which have spin degenerate
ground states in the absence of a magnetic field-and that means only for
molecules with an odd number of electrons. The spectra observed consist of
transitions between members of this so-called Kramer's doublet (there is more
discussion of Kramer's doublets in Section 9.10 and in Appendix 10). Further,
there are often peak-broadening phenomena which can only be overcome
by working at very low temperatures, perhaps as low as that of liquid helium.
These peak-broadening phenomena are particularly severe when the ground
state has another orbital level not far above it in energy (not far means
comparable with kT ). Thus, it is very difficult to observe the spectra of
octahedral complexes of titanium(III), d 1, for which the ground state is
derived from 2 T 29 by application of spin-orbit coupling and low-symmetry
Photoelectron spectroscopy (PES) 1 291
crystal field perturbations. Here, the ground state is usually such that it has
two orbital (four spin-orbital) terms not far removed from it because of their
common 2 T 29 parentage. Manganese(Il), ds, however, with its high spin
19 ground state, gives room-temperature spectra. Another broadening
6A
7 The nearest that one can get in practice is to use monochromatic ultraviolet light
from a synchrotron source. Such sources have the advantage of being tunable-so that the
wavelength of the radiation can be chosen to be the optimum for the measurement-but the
disadvantage that they are only available at a relatively small number of national or
international facilities.
Photoelectron spectroscopy (PES) 1 293
12 16 20 24
(a)
ev-
FJg. 12.23 The He" PES of (a) Cr(C0) 6 and
(b) W(C0) 6 . Adapted and reproduced with
permission from B. R. Higginson, D. P. Lloyd,
P. Burroughs, D. M. Gibson and
A. F. Orchard, J. Chern. Soc., Faraday Trans. 2
(1973) 69, 1659.
8 12 16 20 24
(b)
ev-
294 1 Other methods of studying coordination compounds
(a)
ev-
Fig. 12.24 The He' PES of (a) Mg(C5 H5 ) 2 and
(b) Fe(C5 H5 ) 2 • Adapted and reproduced with
permission from S. Evans, M. C. H. Green,
B. Jewitt, A. F. Orchard and C. F. Pygall,
J. Chern. Soc. Faraday Trans. 2 (1972) 68,
1847.
18
(b)
ev-
about 8.5 eV in the PES, ionization from the metal t 29 d orbitals is seen. For
W(C0) 6 this band has a shoulder, clearly resolved in the He' PES. This
splitting is attributed to spin-orbit splitting in the ti. orbitals of [W(C0) 5 ]+.
Between 13.3 and 14.4 eV lie the CO n bonding combinations t 1., t 19 and
t 2 •• The t 29 member of the n bonding set appears within the span of the CO
a bonding combinations e., t .. and a 1 • at lower energies (they vary in going
from Cr(C0) 6 to W(C0) 6 but are centred at ca. 15.6 eV). In the region
17-22 eV lie the oxygen a lone-pair combinations a 19 , t 1• and e9 (these are
not included in Table 10.4). It is to be noted that although the oxygen a
(outward-pointing) orbitals are physically well separated in the molecule,
the involvement of, and mixing with, what are more inner orbitals-of the
same symmetry-is sufficient to cause large splittings.
As a final example consider the case of ferrocene, Fe(C 5 H 5 ),, for which
the He' PES is given in Fig. 12.24 (b) together with that of Mg(C 5 H 5 ), (a).
Although there are marked similarities between the spectra there are also
clear differences. A pair of bands at ca. 7 eV in ferrocene have no counterpart
in magnesocene and are to be associated with ionization from the d orbitals
of the transition metal. The pair of bands between 8 and I 0 eV are closer
together in ferrocene whilst for this molecule the broad band between 12
and 14 eV is more structured. The pair of bands at ca. 7 eV have been
assigned to e2 g(3d) and an a 1 g(3d)-yet in Section 10.7, and in particular
Table I 0.3, we saw that in ab initio calculations on ferrocene these are rather
Evidence for covalency in transition metal complexes I 295
low-lying orbitals and not those that would be expected to be the most readily
ionized. We have a breakdown of Koopmans' theorem. The reason for this
pattern-in terms of the high degree of localization of the electrons in these
orbitals and so the large readjustment energy of the remaining electrons-
has been discussed in Section 10.4. Returning to the photoelectron spectrum
of ferrocene, the 8-10 eV bands have been assigned to e 1u(n-Cp) and
e 1g(n-Cp) whilst the 12-14 eV region is assigned, qualitatively, to ionization
from ligand CT and n orbitals. It is illuminating to compare the experimental
patterns with the alternative sequences given in Table 10.3.8
One final comment: in this section we have been concerned with vacuum
photoelectron spectroscopy. A similar X-ray photoelectron spectroscopy
exists in which the ejected electrons come from inner electron shells. The
ejected electron energy is sensitive to the chemical environment from which
it originates and so gives information on this.
explains its properties without the need to invoke significant d orbital participation on the
sulfur. See J. Cioslowski and S. T. Nixon, Inorg. Chern. (1993) 32, 3209.
Cyclic voltammetry 1 297
investigate its fate. The method also provides information on the reversibility
of electron gain/loss processes, their kinetics and on the strategies that might
be successful in the bulk preparation of species for which there is cyclic
voltammetric evidence.
At its simplest, cyclic voltammetry (CV) consists of applying a saw-tooth
voltage across two electrodes (often of platinum but carbon or mercury are
also used), usually in an aqueous solution containing a supporting electrolyte
as well as the ion under study. The supporting electrolyte is involved in
electrical current transfer across the bulk of the cell but is not involved in
the electrode reactions. In order to ensure that this is so, the voltage range
of the saw-tooth is limited. To avoid complications caused by the electrolytic
reduction of dissolved oxygen, the solutions are degassed; when non-aqueous
solvents are used great care must be taken to remove traces of water. The
concentration of the ion under study is usually less than I% of that of the
supporting electrolyte. This means that as the electroreduction (say) of
A--> B progresses, the layer of the solution adjacent to the electrode becomes
depleted of A and so the electroreduction, ultimately, becomes diffusion-
controlled. If the potential across the cell is now reversed, as it will be if
the applied saw-tooth spans both positive and negative potentials, then,
ultimately, the electrooxidation B --> A will occur. If B has not been reduced
in concentration by some reaction, a similar pattern will be followed~a
maximum current followed by a drop to a diffusion-controlled limit. The
sequence is illustrated in Fig. 12.25.
Although a sequence A --. B; B --. A has just been followed, it is clear that
if some B has reacted en route then the final condition is not the same as
the starting. That is, successive voltammagrams will not be identical. A
further reason for a difference would be if the products of the side-reactions
of B were themselves electroactive. An important variable in studies aimed
at investigating such phenomena is the slope of the saw-tooth waveform.
Rates of up to ca. 100 V s- 1 are common~ which means that a complete
scan takes only about 10- 2 s~although they can be very much slower.
The method is a very useful one for the qualitative, rather than quantitative,
characterization of compounds. It has found common application in bio-
inorganic chemistry, where species often show a family of oxidation-
reduction steps. If one were trying to synthesize a model complex analogous
to a naturally occurring bioinorganic, one question that it would be natural
to ask is, does it have a similar CV behaviour?
Two final comments: first, although the above discussion has been about
a two-electrode process, the real-life experiment involves at least three, the
third being a reference electrode (usually calomel or silver/silver chloride).
If the supporting electrolyte contributes significantly to the current~as may
well happen if, for instance, the pH is well away from 7, then a four-electrode
cell can be arranged in such a way that the contribution from the supporting
electrolyte is cancelled. Finally, as hinted above, not all electrode reactions
are completely or immediately reversible. Such irreversibility is indicated by
the separation between the cathodic maximum current potential (E,) and
the anodic maximum current potential (E.) being separated by a potential
which is greater than expected (which is ca. 0.059/n, where n is the number
of electrons involved in the electrode reaction).
X-ray crystallography I 299
Ec
20
'-'
'6
_g 10
"
'-'
op.-----
-20 ~--~----~--~--~--~
0.8 0.6 0.4 0.2 0 -0.2
Potential, V versus SCE
54 3 210-1-2-3-4-5
La Fem(CN)~-
g Fe"(CNJt'
electrons the stronger the re-emission). So, measurement of the angles of the
diffracted beams gives details of the regular array, and thus to some extent
structure, and measurement of their intensities give the information of
greatest chemical interest, the identities of the atoms and of their positions
relative to other atoms. Clearly, measurement of both angles and of
intensities is important. In most laboratories the process is now highly
automated. A tiny single crystal is selected and mounted on a fibre oflithium
borate glass (low atomic weights, low X-ray scattering). Some preliminary
diffraction measurements enable the computer to orient the crystal and to
orchestrate the collection of data. After some data manipulation, the
intervention of the researcher is generally needed to choose the most
probably correct, but approximate, crystal structure, which is then refined
by the computer and the full data set deconvoluted. It sounds too easy!
Whilst, in favourable cases, little knowledge is needed to solve a crystal
structure, learning how to control the computer can be the major hurdle-
there are traps for the unwary. First, the theory assumes that the crystal is
uniformly bathed in X-rays. No scattering, no diffraction, no absorption.
This is why the crystal has to be tiny, perhaps almost too small to see. This
points to a major problem but one that is not always explicitly addressed,
that of whether the crystal studied is representative of the sample from which
it is selected. Next, heavy atoms dominate the scattering process, so that in
structures containing them it is difficult to accurately place light atoms
nearby and sometimes even to identify these light atoms. In many structures,
hydrogen atoms are not detected (neutron diffraction, best on deutero
materials, solves this problem). Again, at room temperature the atoms in a
crystal will be, thermally, vibrationally excited. This motion 'blurs out' the
structure, reducing precision. However, for some samples, cooling-which
would normally give better precision-shatters the crystal (this was a
problem for ferrocene). Particularly if rather rounded molecules are involved,
disorder can occur in a structure, a disorder which has to be modelled in
some way before the crystal structure can successfully be refined. Rounded
molecules can often be accommodated in several different ways in the
structure, all arrangements being of comparable energy. In such a case
the model would be one in which all of the possible arrangements are
superimposed, each weighted according to its occurrence (the weightings
would be varied to obtain the best fit). A different type of disorder occurs
when solvent molecules occupy some, but not all, equivalent sites. Finally, at
several points in this chapter, as well as elsewhere in the book, mention has
been made of spectroscopic measurement on crystals. Apart from the far
from trivial fact that these often demand crystals which are very much larger
than those for X-ray structure determinations (so that it is not a trivial
problem to check that the large crystal has the same structure as that on
which X-ray work was carried out), it is important to recognize that all
spectral interpretations demand the use of a primitive unit cell. This is because
almost all spectral measurements use wavelengths which are large compared
to typical interatomic separations. Molecules or ions related by pure
translations experience the same electric vector originating in the incident
light wave, for example. To deal with this, the pure translations must be
correctly chosen-and the correct choice is that set which interrelates
primitive unit cells. If, as may well happen, crystallographers find it
Further reading 1 301
12.12 Conclusion
In this chapter some of the more important methods of studying coordination
compounds have been reviewed. The detailed interpretation of the experi-
mental results is often rather difficult, and, for the phenomena of optical
activity and NQR, for example, the theory of the method has only
been worked out incompletely. Discussion has therefore been confined to
the qualitative level, it being considered important that the student should
have a pictorial idea of the phenomena considered. In this way he or she
should both have been made aware of which technique is likely to be of use
in tackling a particular problem and also of some of the difficulties associated
with its application. Inevitably, it has been necessary to be selective and this
has been done on the basis either of techniques which the student may well
meet in the laboratory or of techniques which are of particular importance.
Finally, Chapter 16 will provide examples of the application of some of the
techniques described in the present chapter, as well as a few more which
have proved to be of particular value in the study of bioinorganic molecules.
• NMR, NQR, EPR and Miissbauer Spectroscopy in Inorganic There are two key books on the applications of vibrational
Chemistry by R.V. Parish, Ellis Horwood, Chichester, 1990. spectroscopy: Metal-Ligand and Related Vibrations by D. M.
A book with a minimum of theory which provides an Adams, Edward Arnold, London, 1984; and Infrared and
excellent follow-up for the material in the present chapter Raman Spectra of Inorganic and Coordination Compounds, by
and a bridge to more theoretical treatments. K. Nakamoto, J. Wiley, New York, 1986.
• 'NMR and the Periodic Table' R. K. Harris and B. E. Mann A useful source is Volume I of Comprehensive Coordination
(eds.), Academic Press, New York, 1978; provides an older, Chemistry G. Wilkinson, R. D. Gillard and J. A. McCleverty
but excellent and reasonably readable, specialist treatment. (eds.), Pergamon Press, Oxford, 1987, Chapter 8.1 'Electro-
A very readable review of modern aspects and applications chemistry and Coordination Chemistry' by C. J. Pickett.
of NMR is to be found in the July 1993 (page 589 on) issue
of Chemistry in Britain.
Questions 12.3 In the footnote on page 271 it was claimed that infrared
spectroscopy could distinguish three different M(CO), geom-
12.1 Figure 12.26 shows the infrared spectra of cis and trans etries. Substantiate this claim by working out the predicted
[Pd(NH 3 ),Cl 2 ] in the v(Pd-X) region. Which spectrum corre- allowed bands for each geometry (it will be necessary to use
sponds with which isomer? (Figure adapted from R. Layton, group theory). Could a similar claim be made for Raman
D. W. Sink and J. R. Durig, J. Inorg. Nucl. Chern (1966) 28, spectroscopy?
1965).
12.4 It was found that a complex containing an Fe(CO),
unit, when in solution, gives an infrared spectrum with two
bands of equal intensity (to within experimental error) in the
v(C-Q) region. What is the C-Fe--C bond angle?
13.1 Introduction
One of the earliest applications of crystal field theory was its use to explain
irregularities in thermodynamic and related properties of a series of transition
metal complexes as the transition metal was varied. These applications are
usually dealt with in one of the early chapters of a book such as this; their
consideration has been deferred in order to be able to draw on the
background and additional insights provided by Chapters 6, 7 and II.
Although the notation that will be used implies the use of the crystal field
model, the discussion is not limited to this, as the title of the chapter
indicates. Rather, the concern will be with energy level separations, con-
veniently but not always accurately, thought of as resulting from orbital
energy level differences- hence the convenience of using the language of
simple crystal field theory, even if we do not really mean it! Thermodynamic
aspects of coordination chemistry have already been encountered in this
book- they represent the major theme of Chapter 5. The present chapter
is distinguished from Chapter 5 because here we deal with the fine detail;
the earlier chapter was more concerned with gross effects.
reasonably self-consistent, the radii discussed may not correspond to physical reality. For
example, in the KCI crystal the minimum of electron density along the K-CI axis is ca. 1.45 A
304 1 Thermodynamic and related aspects of ligand fields
X=l 3.25
1.1
+~+ 3.00
<llN
<l>
UN
+
c
I 1.Q "";"~+ § i
.!!? +
Fig. 13.1 The lanthanide contraction and <l>"'
'0 :J
·c :c ' ... -~ +_,"'
'+..... 'ON
X~
metal-halogen distances for divalent first-row
transition and related metal ions. The thin solid "'"'
.£
~-~
~ Lanthanides 2.75 I
II
::;;::;;
lines indicate the probable values for the M-X ro r:: 0.9
_J .Q
distances corrected for crystal field effects.
2.50
0.8
Atomic number ____...
transition metal cation and the surrounding ligands are purely electrostatic.
The theory suggests, therefore, that it should be both valid and possible to
obtain values for the ionic radii of transition metal ions from crystallographic
data. How might these radii be expected to vary from one ion to the next?
If crystal field effects are neglected, the essential difference between adjacent
members of a series of ions such as Ti3+, yH, Cr3+ and so on, is that each
successive member has an extra positive charge on its nucleus and an extra,
compensating, d electron. As a first approximation, therefore, no change in
ionic radii along the series might be expected. Recognizing the incomplete
screening of the additional positive charge by the additional electron,
however, a small decrease in ionic radius seems more probable. These
qualitative ideas find support in the gradual decrease in ionic radius exhibited
by successive trivalent ions in the lanthanide series-the so-called lanthanide
contraction, although here, of course, it is f electrons which are involved.
This particular ionic radius decrease is shown in Fig. 13.1 together with the
metal-ligand separation in halides of the first transition series (in these the
metal is octahedrally surrounded by halide ions). By plotting metal-ligand
separations an assumption about the ionic radius of the halide ion is avoided.
Values for copper(II) compounds have been omitted from this diagram.
These complexes are usually highly distorted, so that Cu-X distances both
greater and smaller than the values predicted by interpolation of the data
for adjacent ions have been reported. The smooth curves which were
anticipated do not appear in Fig. 13.1. There is a simple explanation for this
which becomes evident when crystal field effects are included. In the absence
of a crystal field all of the metal d orbitals are degenerate and so, effectively,
equally occupied. Application of an octahedral crystal field removes the
degeneracy, so that the t 2 " orbitals are preferentially filled. This means that
the d electrons are preferentially placed in orbitals which do not screen the
ligands from the increased (attractive) nuclear charge as we go from one
metal to the next. When an electron is placed in an e" orbital, however, it
has an enhanced screening effect, because these orbitals are concentrated
along the metal-ligand axes. For a d 5 (high spin) configuration all of the d
from the potassium and 1.70 A from the chlorine, compared with the usually quoted
ionic radii of 1.33 A and 1.81 A respectively. However, the (presumably) more accurate radii,
such as that of 1.70 for Cl-, arc not constant but vary from compound to compound and so
the additivity is lost.
Heats of ligation 1 305
orbitals are equally occupied and the screening is the same as it would have
been in the absence of a crystal field. It follows that the additional screening
of an electron in an e, orbital compensates for the deficiencies in the screening
of one and a half electrons in the r2 , orbitals. The deviation of ionic radii
from that given by the simple picture is therefore expected to be in the order:
t~ > t~ > t~eg > ttt > t~ "' t~ei > t~g > t~gei
"' t~eg > t2g "' ttte~ > t~e~ "' t~e; "' 0
Remembering that all of the examples given in Fig. 13.1 arc high spin, it will
be seen that this series is followed quite well. The only exception appears to
be d 7 (high spin) cobalt( II) for X= 1-. However, it has already been
recognized in Section 7.5 that cobalt(II) has the configuration ~~~1 5 e: 115 in
the weak field limit--and I- is a weak field ligand. If the weak-field limit
values are taken, then this configuration for cobalt(II), along with d 2 which
has a configuration ~~~5 e:l', should be placed with the configurations
rt,e~ and t~ 9 e 9 in the above series. In practice, high spin cobalt( II) complexes
have configurations intermediate between ~~:1 5 e: 1/ 5 and ti.e; and so occupy
variable positions in the series. A consequence of this general discussion is
that if the distribution of electrons between t 2 , and e, orbitals changes, then
there will be a synchronous change in ionic radius. So, the ionic radii of
transition metal ions in electronically excited states are expected to differ
from their ground state values (to be greater, because the e, population
increases for spin-allowed d-d transitions). Conversely, when the population
of the r2 , orbitals increases at the expense of thee, a decrease in ionic radius
must be expected. It will be seen in Section 16.2 that just such a change
seems to be involved as a trigger in a mechanism which causes iron(II) ions
some 30 A or more apart in hemoglobin to be sensitive to whether or not
their partners in the molecule are coordinated to 0 2 .
Table 13.1 Crystal field stabilization energies for weak field and intermediate field
octahedral complexes. Where alternative configurations are given, the fractional
values are the weak-field limit and the integer values are the strong-field limit
(between them they give the intermediate field range)
dn Crystal field Crystal field
configuration configuration stabilization energy
do 0
dl t~g -2/51!.
d2 t§.. -4/51!.
t~sel;s -3/51!.
d3 f,g -6/51!.
d4 f,gei -3/51!.
d5 ~.e~ 0
d6
~i -2/51!.
d7 t~ge~ -4/51!.
~;;seil/5 -3/51!.
d" F,ge~ -6/51!.
dg q.e~ -3/51!.
dlO F,ge; 0
-2000
,
Fig. 13.2 Heats of hydration of first row 0
transition metal ions. Experimental values E
(which are subject to varying errors) are shown
by circles. When corrected for the contribution
made by crystal field stabilization energy
(crosses) they conform to a smooth behaviour
(solid line).
-1500~----~----~--~--~~--~--~--~-----
crystal field stabilization energies given in Table 13.1. Indeed, if one corrects
for the crystal field stabilization energies, using this table and the spectro-
scopic values oft., then the resulting points fall on an almost straight line.
For cobalt(II), d 7 , two points are shown, corresponding to stabilizations of
-3/5 t. and -4/5 t., for the reason discussed above. The curve passes
between these limits. This relationship between experimental and corrected
values of the heat of ligation has led to the suggestion that it may be used
as a method of obtaining approximate values oft.. However, our discussion
suggests that the agreement that is found with the spectroscopic values oft.
is somewhat fortuitous, for the ionic radius effect must be superimposed on
the crystal field stabilizations in Fig. 13.2. Further, several assumptions have
been made in the above discussion. In particular, one fact has been
overlooked- that when the randomized arrangement of d electrons changes
to the ground state arrangement there is an increase in the effective
charge seen on the cation by the ligands-the number of e" electrons
screening the nucleus decreases. Accordingly, there is a change in the
electrostatic energy. This emphasizes the interrelationship between effects
which have been discussed separately. However, since the contributions to
the total energy arising from the variation in metal-ligand distances and the
crystal field stabilizations act in the same direction-compare the inequalities
given in Section 13.1 with the relative magnitude of stabilization energies
seen in Table 13.1--the essential point is that the seemingly rather erratic
nature of the experimental data in Fig. 13.1 can be understood.
2800
X~ Cl
~
0'
E X~ Br
£
x~1
Fig. 13.3 Lattice energies of the dihalides,
MX2 , of the first row transition series and
related elements.
2000~----~~--~~----~------L--L--------
Table 13.2 Crystal field stabilization energies for tetrahedral complexes and comparison with high spin octahedral complexes.
As indicated in the text, all quantities involving Ll. are moduli, absolute values without regard to sign (this means that we
do not need explicitly to take account of the fact that t.,.,
and i\. 0 ,, are of opposite sign, although it is convenient to include
negative signs in this table to correctly indicate the fact that it is stabilization energies which are listed)
Table 13.3 Site preference energies, used to predict spinel type. In this table it has been assumed that ~(M 3 +) " 3/2 ~(M 2 +)
M>+ M3+
High spin Octahedral site High spin Octahedral site Difference in site Predicted
configuration preference energies configuration preference energies preference energies spinel type
Ll.oc~(MJ+) = 3/2 Ll.oc~(M2+ ). Using Table 13.2, the data in Table 13.3 are
derived, in which the type of spinel lattice adopted is also predicted. The
prediction is that Mn 3 0 4 (Mn2+ is d 5 ) is normal and _that Fe 3 0 4 (Fe2+ is
d 6 ) is inverted, both as found. Co 3 0 4 (Co2+ is d 7 ) is normal but is predicted
to be inverted. However, the energy difference between the two forms for d 2
and d 7 is only 1/15 Ll. 0 "(M 2 +), so that the prediction is hardly to be regarded
as reliable, particularly when the variability of t 2 , and eg occupancy for d 7
ion is recalled. For Co 3 0 4 and other cases which do not follow the simple
predictions, more detailed analyses have led to agreement with experiment.
It will be noted that high spin configurations have been used throughout
this discussion. The presence of low spin configurations would introduce the
complication of pairing energies.
~G = -Rtlnk
'2
Fig. 13.4 Variation of log10 K1 across the first ~
transition series for the reaction ~
[M(H 2 0) 6 ] 2 + +en~ [M(H 2 0) 4 (en)]2 + + 2H 2 0 ~
13.7 Lanthanides
The recognition that crystal field stabilization energies play a role in the
chemistry of the first row transition elements has prompted a search for
similar effects elsewhere in the periodic table. In particular, the question has
been asked 'do similar effects occur for the lanthanides'? It was recognized
in Chapter 11 that crystal field effects are small for the lanthanides and so,
too, therefore must be any consequent stabilization. In Chapter 7 it was seen
that the seven f orbitals split into a 2 • + t 1• + t 2 • sets in an octahedral ligand
field. If we make the-significant-assumption that this is the relevant
symmetry for lanthanide complexes, then we need the relative energies of
these three sets. We actually have them. In Section 7.5 we found that in the
octahedral weak field limit-surely the limit applicable to the lanthanides-
the 3 F term arising from the d 2 configuration splits into 3 A2 • (6/5 ~), 3 T2 •
(1/5 ~)and 3 T 1• ( - 3/5 ~)components. In that so much of ligand field theory
is symmetry-determined, it is not surprising to learn that the energies of these
components are directly proportional to those of the split f orbitals in an
octahedral crystal field. The differences arise from the fact that we are dealing
with seven f orbitals, not five d. This means that g suffixes have to be replaced
by u and that ~ has to be replaced by -~. This latter point is most
readily seen by considering, the fxyz orbital. This orbital, of a2 • symmetry, is
the most stable f orbital in an octahedral crystal field because it points away
from all the ligands. In contrast, for the d 2 configuration it was shown in
Chapter 7 that the 3 A 2 • is the least stable 3 F component. So, the splitting
of the f orbitals in an octahedral ligand field is as given in Fig. 13.5. It follows
that the lanthanide crystal field stabilization energies will be
Free-atom
f orbitals
Fig. 13.5 The relative energies of f orbitals in
an octahedral crystal field.
10
....
X 5
<I
LU
~
u 0
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
fiC. 13.6 (a) Octahedral crystal field
(a) Lanthanide (+3)
stabilization energies of the tnpositive
lanthanide ions. (b) Enthalpy data related to
the formation of (octahedral) complexes of the
ligand shown in Fig. 13.7 (see also the caption
to Rg. 13.2). 25
~
L
0
E
:>1
0
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
(b) Lanthanide (+3)
valid step for f electrons, where the effects of electron repulsion and
spin-orbit coupling are much more important than those of the crystal field.
Nonetheless, it seems from Fig. 13.6, and certainly has been argued, that
some very small crystal field stabilization occurs for the lanthanide bipyridine-
dicarboxylates. Given the magnitude of any such stabilization, it is not
surprising that no crystal field modulation of the lanthanide ionic radii was
observed in Fig. 13.1.
Fig. 13.7 The 2,2'-bipyridine-6,6'-
dicarboxylate anion.
lengths that are within about 0.01 A, and bond angles that are within about
1°, of those observed. It is not surprising that considerable efforts have been
made to extend these calculations to inorganic chemistry, particularly at a
time when, as became evident in Chapter 2, there is heightened interest in
large organic ligands and ligands with interesting steric properties, both of
which are properties which can already be treated rather well. Applied to
inorganic species, molecular mechanics is not yet able to yield the accuracy
found in organic chemistry. Not that it is without success-an error in a
crystal structure determination was discovered when molecular mechanics
predicted a different ligand conformation. Indeed, progress is encouraging
but the transition metals exhibit a variability of geometry not matched in
organic molecules and the prediction ofthis is a major problem. Nonetheless,
in well-defined areas such as the subject matter of Appendix 1, predictions
are usually in good accord with experiment. Similarly, when the outcome is
dominated by the properties of the organic ligand, good results are achieved.
A common problem, however is the occurrence of different predicted
molecular geometries with energy separations which are within the errors
inherent in the calculations. One criticism of many of the models adopted
is that they make no specific allowance for those factors which, earlier in
this chapter, were highlighted as holding the key to the particular geometry
adopted. Attempts to circumvent such problems are beginning to be made,
for example by the union of molecular mechanics and angular overlap
calculations (see Section 10.6).
13.9 Conclusions
In this chapter some rather crude approximations based on crystal/ligand
field theory were used to make a variety of predictions or correlations.
Although no one piece of evidence is really definitive, taken together they
provide support for the general crystal field approach. However, it must be
emphasized that arguments based on crystal field stabilization energies must
always be used with great care and with due regard for other energetic and
en tropic factors involved in the processes considered. Such factors are at the
heart of the application of molecular mechanics calculations to inorganic
systems. These calculations have minimal computer demands compared with
those of their detailed quantum mechanical counterparts and represent a
rapidly growing field in which significant advances are to be expected.
14.1 Introduction
In the majority of chapters of this book the concern has been with an
understanding of the properties of individual molecules, properties which are
regarded as essentially time-independent. However, no less important are
the chemical reactions of these molecules and, here, changes as a function
of time are of the essence. This chapter is devoted to a review of our present
understanding of some of the reaction types which are characteristic of
coordination compounds. It is as well to recognize the complexity of the
problem. Suppose we are interested in a reaction such as the aquation of an
ion such as [Co(NH 3 ) 5 Cl]2+, a much studied system:
measure is some sort of average. Hopefully, there will be only one, clearly
defined, lowest energy reaction route and the average will be dominated by
this and its minor deviants. However, the three contributions listed above
(and others could be added) will presumably have different temperature
characteristics and so the (average) reaction route will also be slightly
temperature dependent. Fortunately, this complication can usually be ignored
-it seems to be of lesser importance than experimental error. Attention
therefore focuses on a single reaction pathway and, in particular, the different
potential energy profiles associated with alternative pathways. The task of
the worker in the area is to use the experimental data, and, in particular,
rate laws, to deduce the most probable reaction pathway. It is not an
easy task. As will be seen, the existence of pre-equilibria (which mean that
the real reactant species is present in much lower concentration than
expected), the involvement of the solvent (which, because its concentration
scarcely changes, will not be evident from the rate law), the ability to work
over limited temperature and concentration ranges, all impose problems.
Fortunately, more techniques are becoming available; thus, the ability to
study reaction rates in solution as a function of the external pressure applied
to the solution offers information as to whether the reaction pathway
involves compression or extension of the reactant species and, thus, insights
into its molecularity.
Only in one area can theory be said to have led experiment. This is in one
aspect of oxidation-reduction, electron-transfer, reactions, a topic to which
we shall return. First, however, we shall give an overview of the very
important topic of reaction rates. Perhaps the most evident thing about them
is their enormous range. They (or, perhaps easier to think of, the time to
half-completion of a reaction) span a range of at least 10 15 • Clearly, a
theory is to be regarded as acceptable if it succeeds in correctly predicting-
or interpreting-an order of magnitude, exact agreement is too much to
hope for. The range of values emphasizes one thing. For a reaction to proceed
it must be thermodynamically feasible. The range of rate and t 112 values
can arise from variations in mechanism-the availability of a facile mechanism
or the absence of one (thermodynamically, CCI 4 should react violently with
water)-but a given mechanism can be associated with very different rates
also. 1 Ultimately, the hope is that the mechanism can be related to the
electronic properties of the molecule(s) under study and relative rates thus
explained. As will be seen in this chapter, considerable progress has been
made, but much remains to be done.
In Section 2.1 the distinction between inert and labile complexes was
encountered. Several attempts have been made to formalize this distinction,
of which the most popular seems to be Taube's definition: 'if no delay is
noted in the substitution reaction under ordinary conditions (i.e. room
temperature, ca. 0.1 M solutions) the system will be described as labile'.
However, for most chemists inert complexes are effectively those for which
their reactions may be studied by classical techniques, such as monitoring
the change in intensity of a visible or ultraviolet spectral peak with time.
Such reactions are half complete in about one minute or longer at 25 oc for
1 So, the rates of substitution of Ni 11 and Co111 are very different (fast and slow, respectively)
but the mechanisms involved are probably rather similar.
Introduction 1 319
concentration of ca. 0.1 M. In the past, labile complexes were not so readily
studied but current techniques have provided a wealth of data on them.
Reactions of inert complexes are usually studied by mixing solutions of the
reactants and monitoring either the appearance of a product species or
disappearance of a reactant. It is difficult to mix two solutions completely
in less than a millisecond (and only then in small volumes and in some sort
of flow system) and this limits the extent to which technical ingenuity may
be used to apply classical techniques to fast reactions. Some further
extension, enabling the study of some not-too-labile complexes, is possible
by working at low temperatures with very dilute solutions, but for very fast
reactions quite different methods must be used. The simplest has already
been mentioned (Section 5.3): if a stepwise equilibrium constant is known,
together with the rate of either the forward or backward reaction, then
enough data are available to enable the unknown rate to be determined.
Other methods study an equilibrium as the following two examples illustrate.
Suppose that a diamagnetic complex containing coordinated trimethyl-
amine is dissolved in trimethylamine and the proton magnetic resonance
spectrum of the solution studied. If the coordinated trimethylamine exchanges
with the solvent slowly, then two resonances will be observed (provided
that complicating features are absent), one due to coordinated and the other
due to free trimethylamine. If the exchange is rapid, only one resonance will
be observed, at some sort of average position. In favourable cases, if the
temperature of the sample is varied, at one temperature slow, and at another,
fast exchange will be observed (when it will be found that the position of
Fill. 14.1 Dynamic exchange in NMR. the single resonance is a concentration, i.e. peak-area-weighted, average of
The top spectrum is the low-temperature, those of the two) as will be the intermediate region, in which broad peaks
no exchange, limit The bottom is the
high-temperature, rapid exchange, limit The occur (Fig. 14.1). From such measurements the rate of exchange at the
central spectrum was recorded at ca. 200 'C. temperature at which intermediate behaviour is observed may be obtained.
Those immediately below and above were Strictly speaking, the terms slow and rapid in this example are relative to
recorded at ca. 190 'C and 210 'C,
respectively. This particular example arises from the NMR timescale and this depends on the separation between the peaks
two proton environments within a molecule due to coordinated and solvent trimethylamine. If the complex is para-
becoming rotationally equivalent with increase magnetic, additional line-broadening and shifts occur. These may be analysed
in temperature. For the case described in the
text the two low-temperature peaks would be to give rate data but the results tend to be somewhat ambiguous, since it is
of different areas and the high temperature at the slower of two processes which is measured. These are the exchange
the area-weighted average position. processes of interest and that of a paramagnetically induced change in the
nuclear spin state of the protons (for proton magnetic resonance) of the
coordinated ligand. Measurements with a different but similar ligand may
establish which process is the slower, but in such cases one often has to be
content with only being able to put a limit on the rate of the exchange process.
A most important approach to the study of the fast reactions of labile
complexes involves relaxation phenomena. The position of dynamic equi-
librium in a system depends not only on reactant concentrations but also
on such quantities as temperature, pressure and an even electric field gradient
(if present). If one of these is suddenly changed the position of chemical
equilibrium will also change slightly and, with rapid response and sensitive
instruments, this change may be detected. The speed with which the new
equilibrium is reached depends on reaction rates which may thus be
measured. In this way Eigen and his co-workers have determined unimolecular
rate constants greater than I 09 s- 1 , which is approaching the limit for
diffusion-controlled reactions.
320 1 Reaction kinetics of coordination compounds
larger separations are possible. At the time of writing the record seems to
be ca. 40 A. Such long-distance electron-exchange reactions are often referred
to as occurring by an outer-sphere mechanism, that is, as not involving the
immediate coordination sphere of either metal. However, it is reasonable to
expect that electron transfer will occur most readily when the two reacting
species are relatively close together. That this is so is indicated by the
observation that outer-sphere electron-transfer reactions are more rapid for
complexes containing ligands such as o-phenanthroline and the cyanide
anion than for corresponding complexes with ligands such as H 2 0 or NH 3 .
That is, a ligand over which a metal electron may be extensively delocalized
(cf. Section 12.8) significantly reduces the magnitude of the barrier to electron
transfer (one may draw an analogy with a current flowing through a piece
of resistance wire: replacing part of the resistance wire by a piece of copper
wire increases the current).
In the example discussed above, the electron configuration of the iron
atom in the [Fe(CN) 6 ] 4 - anion is tL
and that in the.[Fe(CN) 6 ] 3 - anion
is~~ •. Removal of an iron electron from [Fe(CN) 6 ] 4 - leaves a~~. configur-
ation and addition of one to [Fe(CN) 6] 3 - gives a~~. configuration. It is not
always as simple as this. Consider an electron-transfer reaction between a
molecule of a cobalt(III) complex and one of cobalt(II), both octahedrally
coordinated. Cobalt(III) complexes are usually low spin and cobalt(II) are
high spin, so the electron configuration of the cobalt ions will be cobalt(III)
~~ •• and cobalt(II) t~ 9 e;. After transfer of an electron from cobalt(II) to
cobalt(III) these configurations will presumably become t~ 9 e 9 and t~ 9 e 9
(cobalt(II) and cobalt(III), respectively). However, these are not the ground-
state configurations of the ions. That is, after the electron-transfer reaction
both complexes will be electronically excited (this excess of energy will
rapidly be lost either by radiation or, more probably, it will be converted
into thermal energy). Because this electronic energy contributes to the
activation energy of the process, the rate of electron-transfer reactions
between cobalt(II) and cobalt(III) complexes is much slower than that
between the [Fe(CN) 6] 3 - and [Fe(CN) 6] 4 - anions.
There is yet another reason for this. The original cobalt(III) ion in its
normal octahedral ground state would be 1 A 19 • It becomes an octahedral
cobalt(II) ion with a t~ 9 e 9 configuration. It therefore has a 2 E• term (Table
7.5), but the ground state of octahedral cobalt(II) is 4 T 19 (derived from the
~~.e; configuration). That is, in passing from the as-formed cobalt(II) to
ground-state cobalt(II) the spin multiplicity must change. The only available
mechanism is spin-orbit coupling, operating in a way similar to that shown
in Fig. 8.16. Reference back to the text associated with that figure will show
that such spin-multiplicity changes require special, and rare, conditions if
they are to be facile. It is not surprising that an exchange such as
[Co(NH 3 ) 6 ] 3 + + [Co(NH 3 )a] 2 + -+ [Co(NH 3 ) 6 ] 2 + + [Co(NH 3 ) 6 ] 3 +
should be extremely slow, with time for half-reaction measured in hours. The
reader may find it helpful to reconsider the use of charcoal in the preparation
of cobalt( III) complexes (Section 4.2.2) in the light of the above discussion.
At first sight, an attempt to calculate rate constants for electron transfers
of the sort we have been discussing would seem an impossible task. Charged
species, separated by varying amounts of solvent, a solvent which will react
322 1 Reaction kinetics of coordination compounds
and
The equilibrium constant for this last reaction, obtained from emf measure-
ments, is K 12 . For reactants and products of the same size and charge type
the simplest form of the Marcus cross-relationship is
Electron-transfer reactions 1 323
(If the size and charge requirements are not met, a further factor, / 12, appears
within the bracket on the right hand side.)2 For the above reactions,
k 11 = 4 x 103, k 22 = 1.2 x 10- 2 and K 12 = 1.07 x 10 6 . It follows that the
calculated value of k 12 is 7.2 x 10 3 . This is to be compared with the
Fig. 14.3 The ligand pyrazine. experimental value of 4.2 x 103 • In general, the equation gives values correct
to within one or two orders of magnitude. When a self-exchange or
cross-reaction does not obey the Marcus-Hush predictions, at least approx-
imately, it is a reasonably reliable indication that some complicating
feature is involved-a stabilization caused by hydrogen bonding between
ligands in a cross-reaction, for instance.
So far, the electron-transfer reactions which have been discussed are those
in which only the formal valence states of the metal ions involved changes,
reactions which can occur when there is no intimate contact between the
two reactants (although this absence of contact is a point which it may be
difficult to prove). There is another class of electron-transfer reaction, in
which intimate contact between the two reacting molecules leads to reaction
by a bridge or inner-sphere mechanism. A class of compound which has been
much studied in this connection is one based on a species first studied by
r
Creutz and Taube, [(NH 3 ) 5 Ru(pyz)Ru(NH 3 ) 5 +,in which the two reacting
centres (one is Ru11 and one is Ru 111 are linked by a bridging ligand, in this case
pyrazine (Fig. 14.3) although variation in choice of bridging ligand is
common. Inner-sphere mechanisms often have indicators which reveal their
presence. Reactions of the type
chromium(ll) + cobalt(lll) .... chromium(lll) + cobalt(ll)
have been extensively studied, again, notably by Taube and his co-workers.
Cobalt(III) and chromium(III) form inert complexes whilst the corresponding
divalent ions give labile complexes. This means that if a ligand is transferred
from cobalt(III) to chromium(III) in the reaction it will be possible to show
that this transfer has occurred.
Consider the reaction:
[Co(NH 3 ) 5 CI] 2 + + [Cr(H 2 0) 6 ] 2 +-+ [Co(NH 3 ) 5 (H 2 0)] 2 + + [Cr(H 2 0) 5 Cif+
The reaction is carried out in water and the final cobalt(II) product is actually
[Co(H 2 0) 6 ] 2 +, but this is immaterial to our discussion. It was found that if
e
the solution contained labelled chloride ion 6 Cl-) none of the activity
appeared in the [Cr(H 2 0) 5 Cl]2+ product; the reaction does not involve the
free chloride ions present in the solution. This observation indicates that
there must be an intimate contact between the reacting species, a Cl- ion
being transferred from cobalt(III) to chromium(II) and an electron migrating
in the opposite direction. It therefore seems likely that a species something
similar to that shown in Fig. 14.4 must be involved. Similar transfer of the
ligand X from [Co(NH 3 ) 5 X]"+ to chromium(II) occurs for X= Cl-, Br- ,
N 3, acetate, SO~- and PO!-. That there is a transient intermediate
something like that in Fig. 14.4 is supported by the observation that for
X= Ncs- (the complex having a Co-N bond) the initial product is
2 / 12 is a function of k 11 , k22 , K 12 and a collision frequency which itself is usually taken as
4+
5-
FJg. 14.5 Complex anion containing both
iron(ll) and iron (Ill).
(CN)5Fe-NQ>--<Q>-Fe(CN)5
Mechanisms of ligand substitution reactions: general considerations I 325
Potential
energy
barrier
I
Products
Transition
states
Products
Reaction coordinate -
[Ptlal(] + Y -+ [Ptl3Y] + X
is
Here, for simplicity, the ligands not involved in the substitution have been
represented by L 3 ; however, they need not all be identical. Although in a
particular study only one term of the rate law may be important, changing
the reaction conditions somewhat can lead to the other becoming evident.
The use of polar solvents leads to dominance of the k 1 term, of apolar
solvents to the k 2 • Conversely, when Y is a strong nucleophile the k 2 term
is favoured; weak nucleophiles favour the k 1 • As the ligands L 3 are made
more and more bulky, so both k 1 and k2 are reduced in magnitude. Similarly,
when Y is made more bulky both k 1 and k2 are again reduced. Although,
from the rate expression alone, one might guess that the k 1 term represented
a D mechanism, the evidence just presented strongly suggests that the k 1
and k2 mechanisms are very similar. The detailed data are consistent with
very similar mechanisms indeed; that in k2 Y, and in k1 solvent, are involved
in the rate-determining A step.
NH3 + NH 3
Fill. 14.8 A mechanism for the slow
conversion of cis-diamminodichloroplatinum(ll) I I
into the trans isomer. NH3-Pt-C1 CI-Pt-CI
I I
Cl NH 3
not only exerts a trans effect but can exert a cis effect also. One is left
with the distinct feeling that the trans effect, long a cornerstone in synthetic
inorganic chemistry, merits more detailed investigation into its real nature
and origin.
However, the actual rates show surprisingly little dependence on the chemical
nature of L and other quantities-entropies, enthalpies and, to a lesser
extent, volumes of activation-are also surprisingly constant. This sort of
inconsistency is typical in the field, pointing to a hidden complication. The
data can best be understood by a mechanism proposed by Eigen, Tamon
and Wilkins and nowadays referred to as the Eigen-Wilkins mechanism. This
mechanism is one in which the complex C, and the incoming ligand Y, diffuse
together to form a weakly bonded encounter complex, a rapidly established
equilibrium existing between this encounter complex and the free com-
ponents:
[C]=~ (14.4)
1+K•M
Combining Equations 14.2 and 14.3,
rate= kKE[CJM (14.5)
and, from Equations 14.4 and 14.5, we have the final result:
rate = kKE[CJrM
1+K.M
It can immediately be seen that when the product KE[Y] is small compared
to 1 then the commonly observed proportionality of the rate to both [CJr
and [Y] is explained. It transpires that the-often unexpectedly small-
variations in rate constant mentioned above are the consequence of changes
in KE, not of the intrinsic rate k. So, for nickel(ll) complexes the observed
formation rates with respect to [Ni(H 2 0) 6 ]2+ vary from about 200 to
7 000 000 M- 1 s -~, a factor of over 30 000, but the corresponding k values
only vary by a factor of about 2. This means that the observed reaction rate is
controlled by the amount of encounter complex in solution. Relatively
constant is the dissociation step, the rate at which a ligand (here, H 2 0) leaves
the coordination sphere, to be replaced by the ligand in the encounter
complex-a ligand which can be thought of as lurking, waiting for the
opportunity to insert. This interpretation of the reaction sequence is entirely
in accord with the statement at the beginning of this section-and is a
warning not to jump too quickly to mechanistic conclusions from rate
expressions alone.
Substitution reactions of octahedral complexes I 333
rates of reaction and, again, one is warned of the danger of predicting the
behaviour of one species from a knowledge of that of another. Nonetheless,
quite simple ideas do seem to have some validity. Thus, as would be expected,
for a given formal charge, larger ions exchange more rapidly than smaller.
So, the rate constants vary: Cs+ > Rb+ > K + > Na + > Li+ ~ Ba2+ >
Sr2 + > Ca2+ » Mg2+ » Be2+ and so on; it is also usually true that for a
given ionic size an increase in formal charge is associated with a decrease in
reaction rate. The observed tendency of a change from dissociative to
associative activation on moving down a Group in the Periodic Table
becomes, therefore, rather more understandable.
It is not always a simple matter to follow solvent exchange-the overall
reaction is one in which there is no apparent change. Somehow, a label has
to be introduced by means of which exchange can be followed. The obvious
technique is that of isotopic exchange-for instance, preparing a complex
containing coordinated H/ 7 0 and following the exchange with solvent
H 2 16 0, perhaps by mass spectrometry. Clearly, such a method is only
applicable to slow exchanges for otherwise exchange will be complete before
measurements begin. As we have seen, another, much more widely used, set
oftechniques is based on NMR measurements, where, effectively, a spin label
is used. A nucleus which exchanges between bulk and coordination slowly
relative to the instrumental timescale will give two basic peaks (ignoring fine
structure), the relative intensity of which is dependent on the ratio of bulk to
coordinated solvent molecules (and so a ratio which can be used to obtain
the coordination number). A nucleus which exchanges rapidly relative to the
instrumental timescale gives a single peak at an average position; when
exchange and instrumental timescales are similar, broadened peaks result.
The NMR spectra can be studied as a function of temperature, pressure and
time. The latter-by watching the evolution of an NMR signal over a period
of time-enables the technique to be used for exchanges which are too slow
to be studied by the line-broadening technique. These same techniques find
applicability in the study of fluxional systems, a topic which forms the subject
of a subsequent section of this chapter.
First however, a brief discussion of reaction intermediates. As has been
emphasized, the majority of the reactions of octahedral transition metal
complexes are Id (the most common) or D in type. Let us confine our
discussion to the latter, for this is the more clear-cut. A D-type intermediate
must be five-coordinate, which means that, just as for platinum(II) reactions,
square pyramidal and trigonal bipyramidal intermediates have to be con-
sidered, together with a possible interchange between them. Here, however,
our concern is that of adding a ligand to the five-coordinate intermediate,
not of losing one from it, as in platinum(II) chemistry. Is it possible to say
anything about the geometry of the intermediate? What influence do the
other ligands have? Is there a trans effect? It is clear that ligands other
than that being replaced are important. Indeed, steric effects arising from
these ligands are increasingly being invoked as a determining factor. In
general, ligands trans to that being replaced seem to be no more important
than those cis, although there are exceptions-in rhodium(III) complexes,
for example, although the data are limited.
The series cis and trans [Co LX( enh] 2 +, where X is the leaving group and
L the ligand which is either cis or trans to it (X = Cl-, Br-; L = Cl-, OH-,
Base-catalysed hydrolysis of cobalt(lll) ammine complexes 1 335
orders of magnitude faster than the rate of base hydrolysis. Finally, the failure
of complexes without such protons, bipyridyl complexes, for example, to
undergo rapid base hydrolysis supports the conjugate base mechanism.
When the stereochemistry is studied, for example by working with cis and
also with trans disubstituted cobalt(III) ammines, then it is found that both
scramble; stereochemistry is not retained. This is taken to mean that the
five-coordinate intermediate is close to a trigonal bipyramid. A major
effect of n bonding involving the deprotonated ammine on the behaviour of
this intermediate has been postulated. However, although plausible, the
model is not proven and it does not easily explain all of the available data.
This, then, is another example in which a pre-equilibrium has a major
effect on a rate law-in this case, introducing a proportionality to the
hydroxyl ion concentration. It is not difficult to see how this arises. In the
example given above, let the forward rate be k 1 and the backward k_ 1 :
followed by
fast
[Co(NH 3)4(NH 2 )f+ + H2 0 --+ [Co(NH 3) 50H] 2 +
Now,
d[Co(NH 3)4 (NH 2 )CI]
dt
is equal to
k 1 [Co(NH 3)5CI][OH]- k. 1 [Co(NH 3) 4(NH 2 )CJ][H 2 0]- k 2 [Co(NH 3)4(NH 2 )CI]
(14.7)
M-CH
'\
OI
/ CH" CH
_......CH
CH
unit. If the C 5 H 5 ring moves round by one or two steps then the final
molecule is equivalent to the starting one; if the interchange occurs readily,
as it does, then it indicates that the barrier presented by an intermediate
bonding position, such as
/CH\
QcH
CH
M-1
CH /
""- CH
is not high. Not surprisingly, there has been much discussion on the nature
of this intermediate (it could, for example, be one in which the C 5 H 5 ring
is 11 5 , n-bonded with all CH groups equivalent). In fact, it seems that 1,2
shifts provide a general mechanism. Examples of this behaviour are pro-
vided by Hg(C 5 H 5 h ( =Hg('1 1 -C 5 H 5 h), Fe(C0h('1 5 -C 5 H 5 )('1 1-C 5 H 5 ) and
Cu(PEt 3 h('1 1 -C 5 H 5 ).
A second general example is provided by CO ligands, in complexes
containing this ligand bonded in a variety of ways. Crystal structure
measurements have revealed that there is an almost continuous range of
Photokinetics of inorganic complexes 1 339
H bonding positions for CO, from bonding uniquely to one metal atom
H I H (terminal CO), increasingly being bonded to a second until it is equivalently
"crc'-c/ bonded to two (bridged CO), and then an increasing interaction with a third
I \\ until it is equivalently bonded to three (face-bonded CO). This adaptability
H-C C-H has been recognized in some models of metal cluster carbonyls, such as
\\ I
/c-.....crc, Ir4 (C0) 12 , where a central metal cluster (here a tetrahedron) is surrounded
H I H by a polyhedron of CO groups (here a cube-octahedron), the arrangement
H within the polyhedron apparently as much being determined by packing
(a) constraints within the polyhedron as by metal-CO bonding considerations.
From such models it is but a short step to expect that when not all of the
CO groups are equivalently bonded, as in Co 4 (C0) 12 and Rh 4 (C0) 12 -
~~
molecules which contain three bridging CO groups-then the CO positions
will scramble rather rapidly. Indeed this is just what is found.
Finally, a ligand such as the (cyclic) cyclooctatetraene is too big for all of
6
(b) (c)
Fe
(C0)3
the carbons to be equivalently bonded to most metal atoms and in complexes
the ligand commonly adopts an asymmetric position (Fig. 14.9). Again, not
surprisingly now, the particular carbons bonded to the metal often change
rapidly. The phenomenon has been referred to, rather colloquially but
FIIC. 14.9 (a) cyclooctatetraene, drawn as a warmly, as 'ring whizzers'. There is one question which may be asked-does
planar molecule with four localized double
bonds as might be appropriate for it acting as a the ring rotate or does the metal hop?-which is less meaningful than it
symmetrically bonded ligand. (Actually, the appears. We are not concerned with the translations and rotations of the
molecule is non-planar; if it were planar the entire molecule. It follows that any internal motion must have zero linear
double bonds would be delocalized.) (b) And
(c) are two complexes containing momentum and zero angular momentum. Consequently, the motion involved
cyclooctatetraene as a ligand. in ring whizzers must involve both ring rotation and metal hopping, in
opposite directions-ring rotation on its own would mean that the motion
had angular momentum. The study of such phenomena parallels that
discussed in Section 14.5 for metal hydrates. Nuclear magnetic resonance
spectroscopy is used and the broadening and coalescence behaviour of a
signal as a function of temperature studied. Commonly, the experimental
data are compared with computer-generated model spectra; in many cases
the experimental and theoretical spectra are distinguished only by the noise
in the experimental! In such cases the confidence in a correct interpretation
is high, although this confidence is dependent on the correct assignment of
peaks in the slow exchange limit spectrum. The ability to make such studies
has depended critically on developments in NMR spectroscopy, particularly
Fourier transform techniques which have enabled low-abundance nuclei,
such as 13 C and 17 0, to be studied. Nowadays it is not uncommon for several
different nuclei in one molecule to be accessible for study. Such work has
made it clear that there is often more than one fluxional process occurring
simultaneously in a molecule. We are left with the pattern presented at the
beginning of this section-of molecules that are very floppy, perhaps to be
pictured as rounded lumps of jelly stuck together, as if by surface tension,
but quite free for the various bits to wobble, slide and rotate, rather than as
collections of spheres of various sizes, rigidly locked together.
contrast to this stability there are other inorganic systems which are very
sensitive to the action of light. Although the silver-halide photographic
process is outside the scope of the present book, other photographic
processes fall within it. For instance, iron(III) salts are readily reduced to
iron(II) and the oxalate anion is a useful reducing agent. So, although the
complex anion [Fe(oxlJ] 3 - is easy to prepare-concentrated aqueous
solutions of ammonium oxalate and almost any soluble iron(III) salt give
pale green crystals containing the anion-it is not surprising to learn that
it is not very stable, decomposing under the action of light to give iron( II)
oxalate and carbon dioxide. This behaviour is readily explained by the light
exciting an electron in a ligand-to-metal charge transfer transition (LCMT,
see Section 8.1 0.2). This electron transfer oxidizes the ligand and reduces
the metal. The phenomenon is exploited in two ways. First, by impregnating
paper with a mixture ofK 3 [Fe(oxlJ], or a similar species, and K 3 [Fe(CN) 6 ].
When the paper is placed under a mask and then exposed to light-
ultraviolet light is best-and the paper sprayed with water, the areas not
protected by the mask turn blue because of reaction between the photo-
produced Fe(II) and [Fe(CN) 6 ] 3 - to give the blue pigment Turnbull's blue,
KFe(II)[Fe(CN) 6 ]. This simple technique for producing blueprints is used
up to the present day for engineering drawings and the like because it can
easily be applied to sheets of paper much larger than can be accommodated
by ordinary photocopying machines. The salt K 3 [Fe(oxh] is also used in
chemical actinometry-when an aqueous solution of the salt is exposed to
light under standard conditions, the amount of Fe" produced is a measure
of the total amount of light that has passed through the solution. The
quantum yield of Fe"-the number of Fe" produced for each quantum
absorbed-is almost independent of the intensity of the incident light and
of the concentrations of Fe111 and Fe" species. However, the system is
only useful for light of shorter wavelength than yellow. Photolysis of
[Cr(NH 3 j,(NCS) 4 r, which causes release of NCS-, is effected by all
visible light except deep red and is also used in actinometry. For simplicity,
all ofthe above examples concern complexes containing just one metal atom.
A recent example of a more complicated photochemical system is the anion
[Os 18 Hg 3 C 2 (C0) 42 ] 2 -, a molecule in which, essentially, two Os 9 clusters
are joined by a triangle of mercury atoms. Photolysis leads to the expulsion
of a mercury atom, giving [Os 18 Hg 2 C 2 (C0) 42 ] 2 - where the Os 9 clusters
are joined by just two mercury atoms. In the dark and the presence of
mercury this latter complex adds a mercury atom to regenerate the original
compound.
In recent years, inorganic photochemistry has received much attention
and changed considerably from that typified by the historically important
examples just given. There have been three main directions of research. First,
the study of shorter and shorter timescales. This means using short time-pulse
lasers and exploring their immediate (or, if of interest, their not-so-immediate)
consequences. In this way, the sequence of steps following absorption of
radiation can, hopefully, be followed. In practice, there are such intricate
networks of finely separated energy levels, electronic, vibrational and
rotational, that the energy absorbed when an electron is excited usually
very quickly reappears as heat, the system rapidly changing levels when they
cross, until the excitation is dissipated.
Photokinetics of inorganic complexes I 341
and Appendix 7; these have energies that have approximately the same d
dependence as the ground state; the former is involved in the light emission
process of the ruby laser). Similarly, cobalt(III) complexes, with a 1A,.
ground state, have low-lying spin triplets, in particular a 3 T1., which is the
one usually implicated (again, see Table 7.5 and Appendix A.7). A diagram
which is appropriate to [Cr(NH 3 ) 6 p+ is given in Fig. 14.11. It shows at the
top a (spin quartet) electronic excited state derived from the configuration
t~. e; (and so, remembering that e. electrons are weakly antibonding, with
slightly longer Cr-N bond lengths than the ground state). Associated with
this level is a host of vibrational energy levels. Intersystem crossing from the
spin quartet to a spin doublet derived from the t~. configuration occurs
(because of the spin change, spin-orbit coupling is implicated in this process).
Again, vibrational deactivation occurs until some process to give the ground
state is all that remains. The possibilities for this are discussed above.
A final example of how the relative stability oft~. systems in their lowest
excited state can be exploited is provided by the [Fe(CN)sN0] 2 - anion.
A neutron diffraction study has been carried out on a salt of this anion whilst
the crystal was under intense laser irradiation. The intensity of the light and
the lifetime of the lowest excited state (at liquid helium temperature) were
Further reading I 343
............................................................................................................................................................................................
such that approximately half of the anions in the crystal were in the excited
state throughout the diffraction experiment. After measurements had been
completed, the laser was turned off and the measurements repeated, this time
with all anions in their ground state. Comparison of the two sets of
measurements revealed that in the excited state the Fe-N bond (of Fe-NO)
was lengthened relative to the ground state, all other bond lengths being
unchanged within error. Clearly, the result requires an excited state local-
ization which is not consistent with the octahedral ~~. configuration Uust as
does the ground state structure), but the essential point-that the lowest
excited state involves an electron in a metal-ligand antibonding orbital-
remains valid.
Further reading
Chern. (1989) 34, 219; Electron Transfer Reactions by R. D.
An older book which is both so easy to read and so forward- Cannon, Butterworth, London, 1980.
looking that it deserves mention is Inorganic Reaction Mech- A Nobel Lecture gives both a fascinating and enlightening
anisms by M. L. Tobe, Nelson, London, 1972. A good contem- account: 'Electron Transfer Reactions in Chemistry: Theory
porary text is Kinetics and Mechanisms of Reactions of Transition and Experiment' R. A. Marcus, Angew. Chern. Int. (1993) 32,
Metal Complexes by R. G. Wilkins, VCH, New York, 1991. 1111.
Up-to-date accounts will be found in Comprehensive Co- For a touch of the unexpected: 'Dissociation Pathways in
ordination Chemistry G. Wilkinson, R. D. Gillard and J. A. Platinum(II) Chemistry' R. Romeo, Comments Inorg. Chern.
McCleverty (eds.), Pergamon Press, Oxford, 1987, Vol. 1: (1990) 11' 21.
Chapter 7.1 by M. L. Tobe 'Substitution Reactions'; Chapter An up-to-date and easy-to-read overview is provided by
7.2 by T. J. Meyer and H. Taube 'Electron Transfer Reactions'; Reaction Mechanisms of Inorganic and Organometallic Systems
Chapter 7.3 by C. Kutal and A. W. Adamson 'Photochemical R. B. Jordan, Oxford University Press, Oxford, 1991.
Processes'. A wide-ranging review is 'Ru(II) Polypyridine Complexes:
Other references worth mentioning are ·An appraisal of Photophysics, Photochemistry, Electrochemistry and Chemi-
square-planar substitution reactions' R. J. Cross, Adv. Inorg. luminescence' A. Juris, V. Balzani, F. Barigelletti, S. Campagna,
344 1 Reaction kinetics of coordination compounds
P. Belser and A. von Zelewsky, Coord. Chern. Rev. (1988) Complexes', L.S. Forster, Chern. Rev. (1990) 90, 331.
84, 85. 'Fluxionality of Polyene and Polyenyl Metal Complexes'
Dovetailing well with the general level of presentation in B.E. Mann, Chern. Soc. Rev. (1986) 15, 167.
the present book is 'The Photophysics of Chromium(III)
15.1 Introduction
The study of cluster compounds is an active area of current research. The
scope is widening and systematic methods of synthesis beginning to appear
based, for example, on the isolobal principle first mentioned in Chapter 10
and which will be met again later in this chapter. It seems likely that it will
eventually be found that such compounds may -involve almost any element,
perhaps with the exception of the most electropositive. Some representative
examples are given in Fig. 15.1. Despite all this activity, our understanding
of the bonding in cluster compounds is relatively primitive. Most of them
are beyond the scope of detailed ab initio calculations and resource has to
be made to more approximate methods, the Xot and its refinements probably
being the most reliable (see Section 10.3.1). However, it seems that even the
simplest of methods, in particular the extended Hiickel (Section 10.5), can
give the highest occupied orbitals with a tolerable accuracy, though being
far less reliable for more deeply lying orbitals. Unfortunately, cluster
molecules have such a plethora of bonding molecular orbitals that any
photoelectron spectroscopic data (Section 12.7) can only be interpreted with
the greatest difficulty. So, it is difficult to find any reliable external check on
the available calculations.
Before commencing a detailed study it will be found useful to consider a
well known and simple molecule which may be regarded as a cluster-the
tetrahedral molecule P4 . It will prove helpful to compare this with another
tetrahedral molecule which has been the subject of study- the boron chloride
B4 CI4 , which consists of a B4 tetrahedron with a chlorine pointing radially
outwards from each boron; it is made from gaseous BC1 3 passed through
an electrical discharge. The argument developed will be symmetry-based and
so in Table 15.1 is given the character table for the Td point group. Three
different models of the bonding in P4 will be compared, the object of the
exercise being to show the relationship between the three models. Rather
similar relationships occur between the extant descriptions of the bonding
346 1 Bonding in cluster compounds
The three models that follow all lead to the same qualitative result but
they do so by such different routes that one might reasonably conclude that
they are different. As will become evident, these apparent differences are not
fundamental. Although minor differences may persist, in the things of
Table 15.1 The Td character table
importance, the three models are equivalent. The P4 molecule has been
Td E 8C3 3C2 6S4 6ud chosen because just s and p valence shell atomic orbitals are involved; it is
the additional involvement of d orbitals that leads to the more complicated
Al 1 1 1 1 1 situation in clusters involving transition metals.
A2 1 1 1 -1 -1
15.2.1 'Simple ammonia' model for P4
E 2 -1 2 0 0
In NH 3 , one may think of three equivalent N-H bonding orbitals and of
Tl 3 0 -1 1 -1 an additional lone pair of electrons on the N atom (Appendix 2). The
T2 3 0 -1 -1 1 simplest picture of the bonding in P4 is to think of the N of NH 3 being
replaced by P and each H also being replaced by a P, the resulting overall
Bonding in P4 (and B4 CI 4 ) 1 347
(b) (c)
348 1 Bonding in cluster compounds
P atom \
;=:;::=E
P4 edge
molecular
orbitals
Fig. 15.6. There are four faces of the tetrahedron and so just four orbitals
like that in Fig. 15.6. These orbitals can be used to obtain delocalized
molecular orbitals; the steps in the sequence are detailed in Table 15.3. The
orbitals that result are of A 1(0) and T2 (1) symmetries; again, their explicit
forms may be derived using the methods of Appendix 6 and the numbers
in brackets refer to the number of inherent nodal planes; A 1 is more stable
than T 2 . The final molecular orbital energy level diagram that results is
shown in Fig. 15.7. It is immediately clear that this figure poses problems for
P4 • Placing a lone pair of electrons on each P leaves six bonding electron
pairs to be fed into the four orbitals of Fig. 15.7. It cannot be done. So what
of the statement made earlier that 'the actual electron density distributions
implied by Figs. 15.5(a) and (b) are identical'-if this is so, why do the
FJi:, 15.6 An approximate picture of two of the different choices lead to different results? Clearly, this is a point to which we
four tetrahedron face bonding orbitals of P4 •
Each of these orbitals is totally symmetric with must return. Although the present model seems to have failed for P4 , we end
respect to rotation about the threefold axis on this section on a more encouraging note: the model works for B4 Cl 4 , for
which it lies. which it is a simple matter to see that there are jus~ .eight valence shell
electrons available for the cage bonding. Three-centred face bonding orbitals
Table 15.3 The transformation of the of the type just encountered are commonly used in discussions of the bonding
tetrahedron face orbitals of P4 . in the boron hydrides and related compounds. Such species are cage (or
Characters generated by the four cluster) molecules and their apparent electron deficiency is bypassed by using
orbitals under the operations of the Td such three-centred orbitals. Allocating two electrons to each terminal B-Cl
point group (see Fig. 15.6) which is the bond, in B4 Cl 4 there are two electrons from each boron to allocate to cage
sum of the characters of the irreducible
representations A1 + T2 bonding molecular orbitals, a total of eight. Figure 15.7 is, now, clearly
appropriate. But, conversely, Fig. 15.4, which requires 10 electrons, is not.
E 8C3 3C2 6S4 6u" Or is it? At the end of Section 15.2.1, it was suggested that if the highest E
4 1 0 0 2 orbitals are sufficiently destabilized, then they will be unoccupied. It seems
that this is the case in B4 Cl 4 , leading the three-centred bonding model to be
preferred.
P atom
,P,
P-P
bonding
(triatomic)
'------Al
P4 face
molecular
orbitals
350 1 Bonding in cluster compounds
Zt
T1
T2
n (tangential)
cr (radial) E
T2
(a) A1
Fig. 15.10 (a) A combined schematic
molecular orbital energy level diagram arising
from the radially directed a orbitals (a(radial)) of
a tetrahedron together with the n that are t2 (3) (-1.6)
tangential to the local z axes (n(tangential)).
(b) Fig. 15.10(a) adapted to P4 . The results of
calculations/photoelectron measurements on tl (1) (-1.3)
P4 are indicated in units of eV. Values in
brackets are calculated.
n (tangential)
t2 (2) 10.3
a1 (2) 11.5
~'''"
19.3
cr (non-bonded)
:
(b) a1 (1) 29.0
352 1 Bonding in cluster compounds
e(1)
1t (tangential)
a (radial)
~~~
cr(B-CI) ·~
(c) 81 (1)
(a)
Fig. 15.11 (a) Representative edge (upper) and unoccupied (B 4 Cl 4 ); in Section 15.5 further insights into this E orbital will
face (lower) antibonding orbitals of a tetrahedron. be gained.
The face antibonding orbitals are doubly
degenerate and the figure shows beth
One final point. All three models agree that the lowest lying orbital is of
components, one in the left-hand face and the A 1 symmetry but picture it rather differently. In the edge-bond model it is
other in the right. If both components are to composed of edge bonding orbitals (Fig. 15.3), in the face-bond model it is
refer to a single face one component (either)
should, mentally, be reflected in a vertical
an in-phase combination of all four of the face-bonding orbitals (we have
mirror plane containing the vertically drawn not attempted to picture it because such a picture would be rather convoluted
edge in the lower diagram. and complicated) whilst in the atomic-orbital model it is as shown in Fig.
15.9. Yet all three are attempting to describe the same molecular orbital!
Are the differences real? No, not if all three models are taken to the same
limit-the models differ in the sequence in which interactions are 'switched
on', not in the final result. Our conclusion is clear; different models of cluster
bonding may produce results which, superficially, are very different but these
differences may be illusory.
h h
T2
T2
E E Edge-anti bonding
derived
T2
Edge-bonding
derived
(b)
Fig. 15.11 (continued) (b) A qualitative
energyo level diagram which indicates the result
of combining edge bonding and antibonding
orbitals to give (tetrahedral) molecular orbitals.
(c) The corresponding diagram for 'face' h h
bonding and antibonding orbitals, which
combine to give (tetrahedral) molecular orbitals.
T2
In both upper and lower diagrams the final
\
resu~ is akin to Rg. 15.10(a).
E E
\ T2
Face-antibonding
derived
T2
A1
Face-bonding
derived
(c)
2- 2-
812 Hu Bto Hto
2-
Ba Ha
Wade's rules 1 357
in the same way but with one electron fewer). Following the argument
above, it would be expected that B5 H~- is a closo molecule, and so a
trigonal bipyramid. In fact, this anion has not been prepared. Writing it
as B 3 H 3 (BH)~-. and substituting two BH- by CH, one concludes that
B3 H 3 (CH) 2 , C 2 B 3 H 5 , which has been prepared, is also expected to be a
trigonal bipyramid, as indeed it is.
Application to transition metal cluster compounds starts with the 18-
electron rule, in the sense that we take nine as the number of valence shell
metal orbitals of relevance. We reserve three of these, satisfying the (A 1 +E)
symmetry requirement given above, for metal-metal bonding, leaving six for
bonding of ligands, back-bonding, non-bonding-or anything else. Not
surprisingly, we do not attempt to enquire into the detailed characteristics
of the six orbitals' All we do is to allocate 12 electrons to them. Any electrons
in excess of this number are to be associated with the three orbitals we have
reserved for cluster bonding. Consider the Fe(COh unit as an example. The
neutral iron atom has a ... 3d 6 4s 2 configuration-that is, eight valence shell
electrons. In applying the 18-electron rule to the iron atom we add two (O")
electrons from each carbonyl ligand to give a total of 14 electrons. Of these,
12 are allocated to the six non-cluster orbitals, leaving two electrons to be
associated with the three, A 1 + E, cluster orbitals. We conclude that the
Fe(COh unit, like B-H, has two electrons and three orbitals associated with
cluster bonding. This relationship between B-H and Fe(COh is expressed
in the statement that the two units are isolobal, a term already met in Section
10.4.3. Ideally, isolobal units have the same number of orbitals of the same
effective symmetries, of similar size and energies and, formally, are associated
with the same number of electrons. So, because B.H,;- and C 2 B._ 2 H. have
closo structures, it follows that if we were to add another BH, another CH +,
or another Fe(COh (all three are isolobal) that the resulting molecule would
also have a closo structure. In particular, it is to be expected that species of
general formula C 2 BnHn+ 2 Fe(COh will be closo. Such molecules have been
prepared and, indeed, are closo, the C, B and Fe atoms lying at the apices
of a (slightly distorted) polyhedron with (n + 3) vertices.
As an example of the application of Wade's rules to a purely transition
metal cluster consider the tetrahedral molecule lr 4 (C0) 12 , where each Ir
atom is part of an Ir(COh unit (Fig. 15.13); the preparation of this compound
was briefly described in Chapter 4 and, in more detail, in Question 4.4.
0 °0
c c c
Because the electron configuration of an Ir atom is ... 5d 7 6s 2 , there are
9 + 6 = 15 valence electrons associated with each Ir(COh group (the six
oc,/'\/'
'-I/ coming from COO" orbitals). Placing 12 in the six non-cluster orbitals, three
electrons arc left to be associated with the three Ir orbitals allocated to cluster
bonding. This gives a total of 12 electrons for bonding within the tetrahedron,
exactly the same number as for P4 in Section 15.2. Clearly, the general
arguments used in Section 15.2 can be applied to Ir 4 (C0) 12 , although the
OC- l r - - - - l r - C O composition of the various orbitals involved will be less clear, for each may
oc
/-............
oc-lr-co / "co have a 4d orbital component. In summary, then, there are 60 (4 x 12 + 12)
l electrons associated with the cluster lr 4 (C0) 12 , although only a few of them
c arc formally involved in metal-metal bonding. It is found that the number
60 is a characteristic of tetrahedral transition metal clusters. The recognition
0
Fig. 15.13 The tetrahedral structure of of this fact, together with the generalizations implied by Wade's rules, leads
lr4 (C0) 12 . us to take an interest in the number of cluster valence electrons (often referred
358 1 Bonding in cluster compounds
Table 15.6 Cluster valence electron (CVE) counts for some common
polyhedra
structures then, again, electron counting rules may be applied. Suppose the
fusion occurs because one transition metal atom is common to two cages
(an example is given in Fig. 15.14). Remembering that the 18-electron rule
(Section 10.1) is most probably applicable, the 18 electrons associated with
the common atom have been counted twice~once in each cage. The electron
count for the entire cluster is, then, the sum of those for the component,
separate, cages less 18. When fusion occurs between two transition metal
atoms which are bonded to each other and common to both cages then the
count is the sum of those for the component, separate, cages less 34
(34 =twice 18less 2 electrons for the bond between the two atoms common
to both cages). Similar counts are applicable to more complicated fusion
Fig. 15.14 A schematic representation of
single-atom fusion between two cages. Only the patterns but will not be discussed~the above two examples serve to establish
fusion atom is shown explicitly. the pattern.
A,.. 1 1 1 1 1 1 1 1 1 1
A2Jt 1 - 1 1 -1 1 1 - 1 1 - 1 1
Eg 2 0 2 0 -1 2 0 2 0 -1
r,.. 3 1 -1 -1 0 3 1 -1 - 1 0
T,g 3 - 1 -1 1 0 3 -1 -1 1 0
Alu 1 1 1 1 1 -1 -1 - 1 -1 -1
A2u 1 -1 1 - 1 1 -1 1 - 1 1 - 1
Eu 2 0 2 0 -1 - 2 0 -2 0 1
Tlu 3 1 -1 -1 0 -3 -1 1 1 0
r,u 3 - 1 -1 1 0 - 3 1 1 -1 0
Each threefold axis of an octahedron passes through the centre of a pair of opposite faces and each ud mirror
plane bisects four faces, facts that are apparent in the characters generated by the transformations of the
eight faces of an octahedron under the 48 operations of the Oh point group:
E 6C4 3C2 6C2 8C3 i 6S4 3uh Gad BS6
800020004 0
= All +A2u + T1u + T2t
Similarly, each pair of opposite octahedral edges are bisected by a c; axis; these edges are contained in a
ah mirror plane and are perpendicular to a u0 , leading to the reducible representation
E 6C4 3C2 6C2 8C3 i 6S4 3uh 6ud 8S6
12 0 0 2 0 0 0 4 2 0
= A1.g + Eg + TlLJ + T2g + T2u
(b)
Fig. 15.19 (a) An e orbital combination which Here, S, P and D etc. have their usual meanings but (hopefully), for clarity,
transforms like the d,, of a (hypothetical) atom the number of inherent nodal planes have been added in brackets (the
at the centre of the tetrahedron. It is viewed
perpendicular to the z axis. (b) The same
number is of those that are perpendicular to the surface of the sphere). Again,
conbination but viewed along z. The direction of as usual, the degeneracies are given by (2L + 1), where L is the total number
rotation (see the text) is indicated by small of nodal planes. For each orbital on a cluster atom that is u-like when viewed
arrows. This rotation, by 90', gives (c), a
combination which transforms as d?-r2 of an
from the centre of the sphere, the above sequence holds. So, for each cluster
atom at the centre of the tetrahedron. atom, which we shall take to be a first row transition metal, it holds
separately for 4s, 4pz and 3dz, (z being the axis from the centre of the
polyhedron to the atom). As we have seen, then orbitals (4Pxo 4p, and 3dw
3d,.zl are both more complicated and interesting. In particular, they have
the pairing properties discussed above. Because the n orbitals are inherently
nodal they can never give a node-free pattern. Nodes can be added but not
lost. Now, n orbitals are rather like vectors (a clear analogy can be seen
between a EB-8 pattern and the symbol <--) and so we shall use a vector
arrow representation in our diagrams. It is important to recognize that on the
surface of a sphere, as on any other surface, the vectors in an array do not
have to be either colinear or parallel; they can be rotated relative to each
other. In our case this freedom allows the members of a vector set to turn
round-for one member to point in a slightly different direction to another,
to rotate about an axis-without the intervention of a nodal plane.
Simultaneously, the relative amplitudes of adjacent vectors can vary, reflected
in the size of the arrows, so that pictures of the tensor surface harmonics
tend to appear frighteningly complicated (but are not, unless the reader
panics). Taking just one member of each pair of p, orbitals the following
are spanned (again including the number of nodal planes); note the way that,
because we are dealing with n functions, the label S does not appear
where one node is inherent in each case (we are dealing with p, orbitals)
and the others are global. The degeneracy is again (2L + 1), where L includes
all nodes, be they inherent or not. In Fig. 15.20(a) we show the three members
of the P,(l) surface tensor harmonic. To make these more understandable,
against each left-hand diagram (which applies to all polyhedral clusters, no
Free-electron models I 367
matter how many atoms they contain) is, on the right, given its application
to the octahedral case. We shall discuss this latter case in more detail shortly.
For the moment, before continuing further in this chapter, the reader should
carefully study the relationship between corresponding left- and right-hand
diagrams in Fig. 15.20(a); it may be helpful to look back at Fig. 15.17.
The corresponding, paired, rotated rr tensor surface harmonics are
conventionally distinguished from the original by the addition of a bar above
each symbol and so are
.0.(1l; o,(2l; F,(3l, ...
Again, degeneracies are (2L + 1). In Fig.l5.20(b) is shown the P,(l) set and,
again, at the right, the corresponding octahedral orbitals. For both parts of
Fig. 5.20 on the right-hand side it is p, orbitals of the octahedral cluster
which have been shown~d, could have been used as they cover similar sets.
but the pictures would have been more complicated. In Fig. 15.20 note the
way that for each vector (arrow) diagram in Fig. 15.20(a). in the corre-
sponding diagram in Fig. 15.20(b) the vectors (arrows) are rotated by 90'.
As has been noted above, concomitantly, the octahedral r 1, set becomes r 1".
A similar pattern is shown in Figs. 15.2l(a) and (b). which show the D.(2)
368 1 Bonding in cluster compounds
and the D.(2) functions. In both Figs. 15.20 and 15.21 approximate Cartesian
labels for the individual surface tensor harmonics are given, although there
is no accepted convention for these labels.
To complete the picture, b orbitals (d.,_,,, d.,) on the cluster atoms
have to be included. Again, there is an original-rotated set separation (but,
for these functions the rotation needed to interconvert the two sets is 45°).
Because (j orbitals are characterized by two nodal planes perpendicular to
the surface of the sphere, we cannot have combinations with fewer. The
combinations are therefore
eg
Vyz,
eg
370 1 Bonding in cluster compounds
Vxyu tl I y
X
)Z~
z
Free-electron models 1 371
actually is a shorthand for tensor, just as V was for vector in Figs. 15.20
and 15.21. Beware of confusion: all the functions under discussion in this
section are collectively called 'tensor surface harmonics' but the name
'tensor' also reappears to describe the specific functions with the T label.
Although, within this model, all functions of a given type, the seven F0 (3)
for example, are treated as degenerate, not all of them may exist for a given
372 1 Bonding in cluster compounds
cluster. The actual symmetries spanned by the a, :n: and {J orbitals on the
cluster atoms has to be worked out for each case. Correlations then have to
be made with the tensor surface harmonic functions given above, or rather
their components, in the symmetry of the actual complex. Let us look at a
specific example. In Figs. 15.20 and 15.21 an octahedral cluster was used as
an example and we will continue with this choice. In doing so we will tidy
up some loose ends left at the end of our topological equivalent orbital
description of [Mo 6 CI 8 ] 4 + and [Ta 6 Cl 12] 2 +. To this end two things have
to be done. First, the general discussion has to be applied to the octahedral
case. Second, the symmetry properties of the a, :n: and {J orbitals on the
transition metal atoms have to be determined. The first of these tasks is not
difficult. The data in Table 7.6 has to be applied to the spherical tensor
harmonic functions listed above, although our notation must first be
extended to distinguish between the different sets arising from the various
metal orbitals of :n: symmetry, for instance. This is easy. Instead of talking
about D.(2), for example, we will now call this D~, v:or D~, where we
have added p, d or f to indicate the sort of orbitals on the metal atoms
which are involved (although not much use of D~ is to be expected!) and
compensated by dropping the (2). The results are given in Table 15.9, which
is quite horrendous in appearance. This is because it applies to all octahedral
clusters, including octahedral clusters with dozens of metal atoms. We are
only concerned with the simplest octahedral cluster and so only a small
amount of the information in Table 15.9 is relevant. However, the complete
Table 15.9 has a basic pattern which both makes it worthy of inclusion and
of some study. The second thing which we have to do, the determination of
the symmetry pattern of the a, :n: and {J orbitals, is not too difficult either. We
have already discussed in Sections 6.2.1 and 6.2.2 the a and :n: bonding
associated with an octahedral array of atoms. The fact that it is ligand
orbitals which are discussed in these sections is irrelevant-transformation
properties do not depend on the chemical nature of an atom and so the
results needed are those given there. The discussion has, however, to be
extended to include {J functions. There are two of these on each atom, dx'-y'
and dxy• using the coordinate axes defined in the caption to Fig. 15.22. No
Table 15.9 The symmetry properties of the surface tensor harmonics in the
octahedral group Oh
Metal
en
8
en
8
Fig. 15.25 Density of state (DOS) energy
patterns for (a) vanadium metal and V6 (b)
silver metal and A&;. Adapted and reproduced
with pemnission from G. Seifert and H. Eschrig
Phys. Stat. So/. (b) (1985) 127, 573.
A&;
cluster
en
en 8
8
Energy- Energy-
(a) (b)
oc,,c
0 0
c
I_;CO
M M
FJC. 1&.26 The D3h structure of M3 (C0) 12 •
M = Ru or Os (the iron compound has a
oc,..,I~O/''co
c c c
different, but related, structure which is 0 I 0
oc,.., t'co
discussed in Section 12.5.3). M
towards yes. Both, speaking from their own corner, offer explanations of
Wade's rules. Does this matter? Quite possibly not; it depends on the detail
one wishes to explain. So, for the compounds Os 3 (C0) 12 and Ru 3 (C0) 12
(shown in Fig. 15.26) the Xoc method gives the sequence of highest occupied
molecular orbitals (in D 30 symmetry) as
a;, < a2 < e" < a;, < e" < e' < e' < a;,
a;, < a2 < e' < e" <a;, < e' < e" < a;,
Of these, the highest e' and a~ are largely involved in metal-metal bonding
(note that these are also the irreducible representations generated by the
M-M bonds between the three M(C0)4 groups in M 3 (C0) 12 ). In the
photoelectron spectra (that of Os 3 ( CO) 12 is shown in Fig. 15.27) ionization
from the metal-metal bonding orbitals is associated with the lowest energy
peak, arrowed in Fig. 15.27. Ionization from the other levels listed above
accounts for the two peaks on the immediate right of that arrowed; the
orbitals involved are predominantly d, with significant CO bonding involve-
ment. The intense peaks with energies greater than 12 eV in Fig. 15.27 are
largely associated with ionizations from the CO groups. Looking at spectra
such as that in Fig. 15.27, which is for a small, high-symmetry cluster,
together with the orbital energy sequences listed above, one can understand
the attractions of a density-of-states approach! However, if one wishes to
interpret electronic spectra then a detailed knowledge of the occupied and
lowest unoccupied orbitals becomes essential.
Because of the inherent problems associated with calculations on large
clusters, most of the detailed calculations that have been performed concern
diatomic systems such as the [Re 2 CI 8Y- anion. This anion has a short
metal-metal bond and the two (ReCI 4 ) units are eclipsed (Fig. 15.28(a)).
Since the short metal-metal bond can only increase steric interactions
between the two sets of Cl4 ligands, a staggered configuration might
confidently be predicted! The only evident explanation for the observed
6 12 18 eclipsed arrangement is the presence of significant [J bonding between the
Ionization potential (ev) - two rhenium atoms (Fig. 15.28(b)). Does it occur? The answer given by the
FIJI. 1&.27 The photoelectron spectrum of best calculations (usually Xoc) is yes ... but! The 'but' occurs because of a
Os3 (C0),2 . Adapted and reproduced with problem we have met several times in this book-electron correlation.
pennission from J. C. Green, D. M. P. Minges
and E. A. Seddon, lnorg. Chern. (1981) 20, Indeed, one of the workers in the field has written an article entitled
2595. 'Problems in the theoretical description of metal-metal multiple bonds or
378 1 Bonding in cluster compounds
how I learned to hate the electron correlation problem' (see the Further
Reading at the end of the chapter). Essentially, electron repulsion forces
electron density from bonding orbitals, which may be a, nor b, into low-lying
antibonding orbitals, which may be a*, n* or b* The bonding orbital
occupancy is less than given by the simple model and the antibonding orbital
2.24A occupancy greater. Both vary with change in metal-metal internuclear
distance. Larger bond lengths usually mean a greater occupation of anti-
bonding orbitals (in the limit, where the bond breaks, bonding and anti-
~Re~,---CI bonding orbitals are equally occupied). This can be placed in some perspective
Cl,......... Cl by the outcome of some calculations that have been carried out on the bare
(a) Mo 2 molecule. It might reasonably be assumed that the 12 valence electrons
in the molecule might be distributed over the two a bonding molecular
orbitals (one originating in the s and one in the dz' orbital on each atom),
the two (degenerate) rr, d orbital in origin and the two b (degenerate). The
observed bond length in the gaseous molecule is 1.93 A; in principle a
sextuple bond could be involved! In fact, it was only with considerable
difficulty that the calculations could be induced to show any bonding. They
did so only after inclusion of considerable electron correlation. The absence
of ligands does not seem to be the vital factor; in calculations on the dimeric
diamagnetic molecule chromium(II) acetate the presence of the ligands did
not give results greatly different from those on Cri+; again, a correlation
energy contribution was required to give bonding. Note particularly that, as
has been indicated above, the correlation energy contribution is expected to
(b) increase with increase in metal-metal distance. Yet further perspective can
be added to this by the recognition that the metal-metal n bonding overlap
Fig. 15.28 (a) The eclipsed structure of seems invariably less than the carbon-carbon n overlap in benzene-and
[Re 2Cid] 2 -. (b) The Re-Re ii bond which would
be broken were the anion to adopt the the metal-metal n and c5 are presumably yet smaller. It has even been
sterically-preferred staggered configuration. In claimed, on theoretical grounds, that a Cr-Cr quadruple bond is to be
this case the Re-Re bond is short (2.24 A) and expected to be as weak as a Cr-Cr single bond if there are no bridging
so electron correlation relatively unimportant
(see text). ligands in the molecule (only one compound with a Cr-Cr bond but without
bridging ligands is known). The spectroscopic evidence is unambiguous.
In [Re 2 Cl 8 ] 2 - and related anions the electronic transition from the
b bonding to the b antibonding orbital lies at about 7000 A, an energy
corresponding to a fairly typical crystal field splitting (Chapter 8). So,
returning to larger clusters, where the bond lengths are much greater than
in species such as [Re 2 Cl 8 ] 2 -, must we expect electron correlation to be of
any importance? It seems that the answer is yes, although the delocalized
nature of most orbitals, a phenomenon which means that electron density
is spread out and so the effects of repulsion reduced, offers hope that the
importance is limited. Nonetheless, it is the difficulty of including electron
correlation which is a limiting factor in carrying out accurate calculations
on cluster compounds. However, one must proceed with extreme caution.
Not only is the problem a difficult one but it has been found that agreement
with experiment is not invariably a reliable guide to the accuracy of a
calculation. So, as has been mentioned in Chapter I 0, there are cases known
in which a calculation gives good apparent agreement with experiment-a
bond length is accurately predicted, for instance-but, when the calculation
is improved, perhaps by the inclusion of more orbitals on some atoms or
the inclusion of more excited states (which can interact with and thus modify,
the ground state) the apparent good agreement is lost. This is not all. There
Clusters and catalysis, a comment 1 379
16.1 Introduction
Coordination compounds are central to living processes. Although the
cations of Group I (Na +, K +) are usually thought of as very mobile, as
Section 5.7 shows, it is possible to have relatively stable complexes of
them when there is a matching between their size and that of a cavity in a
ligand; change the cavity size and the stability changes. Perhaps, in part,
this lies behind the function of these mobile cations in biology, for example
in the transmission of nerve impulses. Group 2 cations are also important,
presumably also through complexes that they form - Ca 2 + in physical
structures such as teeth, skeleton and shell, as well as functioning as a trigger
in neurotransmission, as a messenger in initiating hormone action and in
initiating blood clotting. The presence of Mg2+ in chlorophyll is well known .
Zn 2+ is present in many enzymes; sometimes it seems to have a structural
role, maintaining a protein structure in a particular conformation for
example, in others it is intimately involved in the chemistry. Complexes of
transition metal ions are of vital biological importance. They provide
mechanisms for electron storage, for electron transport and for catalysis-in
particular involving small molecules and ions such as 0 2, N 2, NO, NO:Jand
SO~-. Is it understandable that bioinorganic chemistry is currently one of
the fastest growing parts of inorganic chemistry 1 In the present chapter this
growth is not the focus of attention. Rather, attention is focused on the way
that the subject throws up fascinating problems which offer a great challenge
and require careful and, often, great ingenuity and subtle methodology to
solve. This, then, is a problem-solving chapter. We start with a general
overview of the subject which will provide a background for the specific
problems that will be considered.
1 A very readable paper which develops and enlarges the theme of this paragraph is 'The
Chemical Elements of Life" by R. J. P. Williams, J. Chern. Soc., Dalton Trans. (1991) 539.
382 1 Some aspects of bioinorganic chemistry
compounds and much water, these are only temporary expedients. Much
better would be the use of metal complexes without the side-effects. Although
the buck-shot approach (synthesize a lot of likely-looking compounds and
try them) can be used, the desired goal is more likely to be reached by a
detailed understanding of the mode of action of a drug, the in vivo
transformations that it undergoes and thus, ultimately, the nature of the
species that is actually leading to the desired effect. This calls for all the
apparatus of bioinorganic chemistry-the identification, isolation, handling
and study of biologically active species, the preparation of model compounds
and so on.
As already indicated, it is not the objective of the present chapter to
attempt to give either a comprehensive account of, or even an overview of,
bioinorganic chemistry. Rather, we shall explore the application of subject
matter in earlier chapters to the topic. From this viewpoint, bioinorganic
chemistry provides a series of novel applications of this subject matter.
Further, some advances in bioinorganic chemistry have. involved specialized
techniques and ideas which, to date, have been less applied to more
traditional areas of coordination chemistry; we shall look at some of these
also. In that such applications will surely become more commonplace,
bioinorganic systems provide an indication of the results that will be
obtained and of the information to be learned from them.
We start with four important points, some of which contain lessons in
their own right. First, although it seems that seldom, if ever, does a metal
ion occur in a biological molecule and not have a function, it should not be
thought that the sole function available is a catalytic one, although this aspect
is often emphasized. Quite the contrary, for instance, a number of cases are
known where the function of Zn 2 + is classified as structural. In some of
these, Zn 2 + forms rather strong bonds to sulfur ligands (usually from the
amino acid cysteine, HS-cH 2-cH(NH 2 }-COOH) which are part of a
protein amino acid sequence. By so doing, the protein chain, or several chains
together, are locked into a particular geometrical arrangement. It seems that
the resulting global, quaternary, structures are favourable to the action of
the protein (and this action could be a catalytic one) and are thus stabilized.
Other non-catalytic functions of metal ions include the transmission of
impulses along nerve fibres (each impulse is associated with a sudden change
from a 'high K + concentration inside the fibre, high Na + outside' pattern.
Immediately after the impulse, a pump mechanism restores the original
concentrations). Another function is that played by Ca 2 + as a messenger in
nerve action (from one nerve cell to the next) and in blood clotting. It seems
that Ca2+ serves such functions because the complexes it forms in these
situations are rather balanced-not too weak, not too strong; not too labile,
not too inert.
The second point is related to the first and echoes one that has already
been mentioned. Metal ions in biological systems are often in rather unusual
environments. This is not just because of the geometrical arrangements of
the coordinating atoms but because of the influence of what one might think
are the more distant parts of the coordinating molecule. Typically, these are
convoluted around the metal ion and its immediate ligating atoms and,
effectively, play the role of the solvent in simpler systems. Except, of course,
that the solvent is a rather rigid one, one that severely limits access of
384 1 Some aspects of bioinorganic chemistry
molecules to (and egress from) the metal ion; seldom, if ever, is it as innocent
as it may appear! Further, the physical bulk of the main ligating molecule
both limits access to, and the mobility of, ligands attached to the metal ion.
By the hydrophobic or hydrophillic surfaces that it presents, it to some extent
selects the species that can approach the metal ion. These features combine
to lead to a rather different chemistry. A link with classical coordination
chemistry is forged by the observation that the coordination equilibria of
complexes of polyfunctionalligands are very sensitive to changes in solvent
composition.
Thirdly, we note that several of the ions mentioned in this paragraph are
not too easy to study. Zn2+ and Ca 2 +, in particular, lack convenient and
sensitive spectroscopic 'handles'. They have no low-lying electronic transitions
by which they can be conveniently characterized; they have no high-
abundance isotope which can be studied by NMR; they do not give
Mossbauer spectra, and so on. The problem is addressed by a technique
which would be almost unthinkable in conventional coordination chemistry.
Such ions are studied by replacing them with a similar ion which does have
suitable spectroscopic characteristics. The reason for this that, although the
biological macromolecule may distort a little when a look-alike ion is
substituted, the distortion is small and good insight into the environment of
the original metal ion may be obtained. So, ifbioinorganic species containing
Zn2+ are treated with an aqueous solution of 1,10-phenanthroline (phen,
Table 2.3) the zinc is extracted because it forms a very strong complex with
phen. If the zinc-free macromolecule is separated it is then readily coordinated
to another cation. Co2+ is a popular replacement ion, but Mn 2 +, Cu2+ and
Cd2+ have also been used. Replacements which have been used for Ca 2 +
are lanthanides, particularly Nd3+ and Eu3+.
The fourth and final point is that species containing two or more metal
ions, linked by bridging atoms, are fairly common in bioinorganic chemistry.
The two ions can be of different metals, so that one is immediately faced
with the question of whether chemistry takes place at both of them, perhaps
simultaneously, or whether one fine-tunes the chemistry of the other.
Commonly, the ions are of transition metals and it is then usual to find that
interaction occurs between what would otherwise be unpaired electrons at
each centre. We shall not develop this particular topic here because all the
important points have already been covered in Section 9.11.
(b)
one with a low affinity in order that the oxygen can readily be made available
where it is needed, within a muscle, for example. Hemoglobin has such
properties, but the final transport is carried out by myoglobin. Myoglobin
consists of a protein with a molecular weight of just under 18 000 and has
just one iron- porphyrin unit, a heme (Fig. 16.1). Hemoglobin, however, has
four heme units, each with its own protein chain, the whole intermeshed to
give a single molecule (Fig. 16.2). The unusual oxygen-affinity properties of
hemoglobin arise from a cooperative behaviour between the four heme
groups, a cooperation that must surely be mediated by the protein chains,
for the heme units are from 25 to 40 A apart. The heme cooperation in
hemoglobin means that when one or two 0 2 molecules are bonded, the
oxygen affinity ofthe other (0 2 free) heme sites is enhanced. The more oxygen
it has, the more it wants! Clearly, this is a useful property to have when
scavenging for oxygen in the lungs. Conversely, when transfer of oxygen to
myoglobin is under way, the less oxygen the hemoglobin has, the less it
wants. Just the necessary characteristics! Oxygen is absorbed by attachment
at iron atoms in myoglobin and hemoglobin (Fig. 16.3), interestingly, neither
uniquely u nor n bonded, but rather at an intermediate, angular orientation
386 1 Some aspects of biolnorganic chemistry
prepared, tend to become brown after having been stored for years;
the iron(II) salt used in volumetric analysis as a reducing agent, ferrous
ammonium sulfate is so used, in part, because it is much less prone to
oxidation than most iron(II) alternatives. So, too, with iron(II) porphyrins
(and the iron(II) found in myoglobin and hemoglobin is in a porphyrin ring,
Figs. 16.1-16.3). They are rapidly oxidized to their corresponding iron(III)
compounds by oxygen. Such oxidation can occur, although much less readily,
with myoglobin and hemoglobin, when met-myoglobin and met-hemoglobin
are formed (these names should not be written with a hyphen; one is
included here solely to facilitate a correct pronunciation). Metmyoglobin
and methemoglobin do not transport oxygen! About 3% of the total
hemoglobin in the blood is oxidized to methemoglobin each day, but
mechanisms exist within the red blood cells for regenerating hemoglobin
from methemoglobin. The reason that the loss is so low-and 3% is low-is
reasonably well understood, as follows. Kinetic studies show that the rate
of oxidation of Fe" porphyrins is second order in the iron(II) species and
first order in oxygen. The most probable explanation is that a 1: l complex
is first formed between an iron(II) porphyrin molecule and 0 2 which
subsequently reacts with another iron(II) porphyrin to give an 0 2 bridged
complex-an example of such a compound of cobalt is given in Table 2.2.
The decomposition of the bridged dimer leads to the oxidation products. In
myoglobin and hemoglobin the protein provides a protective shroud for the
0 2 molecule such that formation of a dimeric species is scarcely possible,
whilst remaining sufficiently open to enable ready 0 2 access and egress. The
low rate of formation of the met derivative is thus explained.
The object of models of the myoglobin molecule has been to provide a
similar environment to that found in the natural material; some examples
are shown in Fig. 16.4. One of these ligands, the picket fence molecule, was
given in Table 2.4 and it will be given again in detail in Fig. 16.5. What
characteristics do the models have to attempt to emulate? In myoglobin and
hemoglobin the iron(II) is in a square pyramidal environment, the four
square-planar coordination sites being associated with the porphyrin ring
and the axial with a histidine (see Fig. 16.1(c)) from the protein chain. Oxygen
enters the sixth site, completing an octahedral coordination around the
iron(II). Whereas the five-coordinate iron is high-spin-four unpaired
electrons-the 0 2 complex, oxyhemoglobin, is low-spin, no unpaired elec-
trons. The reduction in ionic radius of iron(II), high-spin, to low-spin
(Section 13.2) is sufficient to enable the iron(II), initially a bit too big to fit
in the porphyrin hole and so sitting slightly out of the ring and towards the
histidine ligand, to drop more or less into the plane of the ring in
oxyhemoglobin. The consequent tug on the histidine connecting the iron(II)
to the porphyrin presumably contributes to the cooperative oxygenation
effect seen in hemoglobin. How well is this behaviour of myoglobin mimicked
by the model compounds? To those shown in Fig. 16.4 we have to provide
a ligand, for obvious reasons a histidine derivative, to occupy the open
coordination site at the bottom of each species shown. With this addition,
comparison is possible. Of course, the models can offer no information on
the detailed involvement of the protein chain-this problem is being explored
by site-directed mutagenesis (creating mutants in which specific amino acids
are changed-and this is more than just an academic exercise; a number of
388 1 Some aspects of bioinorganic chemistry
R R
R
R
Myoglobin and hemoglobin 1 389
Soret
Oxyhemoglobin (human}'
"582 fl
544 418
Picket fence model 548 429
a The data vary little with source of oxyhemoglobin.
Table 16.4 Oxygen uptake of myglobin and a picket fence model compound
400 450
Wavelength (nm) -
16.4 Peroxidases
The peroxidases are heme enzymes that catalyse the oxidation of organic
substrates, the actual oxidant that is catalysed being hydrogen peroxide. In
a typical peroxidase-horseradish peroxidase is one of the best studied-
it seems that iron(III) is coordinated by a heme ligand, a fifth coordination
site being occupied by the nitrogen of a histidine. The action of hydrogen
peroxide on this enzyme gives an intermediate, usually called compound I,
which carries out part of the oxidation of the organic substrate, itself being
reduced to compound II. Compound II completes the oxidation of the
substrate, the native enzyme being regenerated. The question is: 'what are
compounds I and II?' Before seeking experimental evidence, it is sensible to
make some intelligent guesses. First, it seems likely that they contain both
heme and histidine. If the histidine were lost during the oxidation cycle then
it would have to be replaced before the cycle could repeat. Such a need to
scavenge for histidine would introduce a bottleneck which is scarcely
compatible with an effective catalyst. Secondly, they are likely to contain the
iron and/or heme in an oxidized form. For the iron this means Fe'v or
perhaps Fev, both unusual and therefore interesting valence states. Finally,
since hydrogen peroxide is the ultimate oxidant, it may be that some
fragment of the H 2 0 2 molecule is present in either or both of compounds
I and II.
The sixth coordination site around the iron is formally vacant in the native
enzyme and seems an obvious place for interaction between the enzyme and
H 2 0 2 to take place. Of particular interest in this field have been studies of
the absorption of white X-rays. In this context X-rays may best be thought
of in the sequence
visible light -+ ultraviolet -+ vacuum ultraviolet -+ X-rays
All excite electrons, the X-ray excitation being the most energetic. As is well
known, X-rays are emitted when an outer electron drops into a hole in a
low-lying incompletely filled orbital (this is the basis of everyday X-ray
394 1 Some aspects of biolnorganic chemistry
Energy-
- - Wavelength
Energy-
- Wavelength
table-N and 0; S and Cl. Finally, it is at its best when there is only one
type of individual atom present, one sort of Fe, for example. When there are
several chemically different Fe atoms, it will only yield an average structure.
Another technique that has been used to probe the structures of com-
pounds I and II is magnetic circular dichroism (MCD). Circular dichroism
(CD) has been met both earlier in this Chapter and also in Chapter 12. It
has the disadvantage that it is a phenomenon confined to molecules which
are optically active. MCD is a similar phenomenon but is applied to
molecules which are not optically active. The trick is this. All molecules
distort very slightly in a magnetic field and the distortion is always one that
changes the optical activity of a molecule. A previously inactive molecule
has no alternative, it becomes slightly optically active. The magnetic field
distorts the molecule so as to destroy any centre of symmetry and any mirror
planes. So, in the intense magnetic field of a superconducting magnet, optical
activity is the norm. MCD is particularly useful when the field-free molecule
has some symmetry, for then the magnetic field splits degenerate electronic
energy levels. This means that it is particularly useful for exploring transition
metal ions and their environments. In MCD there are two cases, A and C,
of particular importance:
Because of the theory used to describe the phenomenon, the spectral patterns
resulting (Fig. 16.12) are described as A term and C term, respectively. As
the reader may have guessed, B terms also occur but are much less important
(they relate to magnetic-field mixing of wavefunctions and so are akin to
TIP, mentioned in Section 9.3). Figure 16.12 shows the energy level splittings
which are associated with A and C terms. Because the ground-state splittings
tl:_
A term
-c:::C
II ~
no spectrum is observed in the absence of
a field). In the C term case the thermal
populations of the split levels differ leading to
bands which, although of opposite sign, are of
unequal intensity.
No field
-c:::C Magnetic field
C tenm spectrum
Peroxidases 1 39 7
of C term spectra are small, the thermal populations of the split levels vary
significantly with temperature; the lower the temperature, the stronger the
spectral bands. In contrast, A terms are essentially temperature-independent.
MCD appears to be a somewhat esoteric technique but, in fact, it is of
great value. A complementarity between CD and MCD should be noted. CD
explores a static optical activity and this usually means exploring a property
of the ligand system around, but relatively remote from, a metal ion; MCD
offers the possibility of exploring the properties of the immediate environ-
ment of the metal ion for this is, in part, responsible for degeneracies. Further,
it is not surprising that for paramagnetic species, this paramagnetism is
reflected in their MCD spectra (the technique, after all, is concerned with
the splitting of degeneracies and this is what Chapter 9 was all about, too).
Indeed, MCD can be used to measure magnetic properties and, for this, has
advantages. It is much less sensitive to impurities than are the methods
described in Appendix 8. EPR (Section 12.6) also provides magnetic data
but, really, only for systems with an odd number of electrons. MCD provides
data for both even and odd. Just as a detailed understanding of the magnetic
properties of a complex is none too simple (Appendix 9 gives the simplest
case), so too with MCD. No attempt, therefore, will be made to give a
detailed analysis, but simply to give in Fig. 16.13 the low-temperature MCD
spectra of compound I and compound II. The differences, believed to result
from the one extra electron in compound II, are obvious. Just as for CD, it
is possible to carry out time-resolved MCD measurements and to use the
technique to follow the effects of photolysis, for example. It is in the area of
using spectroscopic techniques to follow such changes that many advances
in bioinorganic chemistry are presently being made.
500 750
Wavelength (nm)
398 I Some aspects of bloinorganic chemistry
0
N-11-N
,;;}e•v1+
Fig. 16.14 The current view of the catalytic //1\ Compound I
~
cycle in which peroxidase (top left-hand) is 1
oxidized by hydrogen peroxide to give compound H is
I. This oxidizes the substrate AH 2 to give AH"
(dots indicate free radicals), H+ and compound
11. This similarly oxidizes more substrate to
/\ 'AH2
regenerate the starting material (His = histidine).
0 •AH+W
N\11 /,N
V/ Fe'I\v
N - -1-N
7 Compound II
His
long-range electron transfer reactions, so that the interplay of Cu11 and Cu'
is of the essence. Although crystallographic work has confirmed the ultimate
conclusion it is interesting to trace the non-crystallographic path leading to
the structure of the site of the copper ion. The compounds are called blue
because of their intense colour, about 400 times more intense than that of
most copper(II) salts. The conclusion drawn was that the Cu11 in the protein
is in a tetrahedral site, because as seen in Section 8.8, such a site makes more
allowed d-d transitions because of mixing of d and p orbitals by the ligand
field. The conclusion was right but the reasoning wrong. The blue colour is
due to a charge-transfer transition, not d-d, although the suggestion of a
(very) distorted tetrahedral geometry was correct. Evidence for the charge-
transfer assignment comes from a comparison of the electronic spectrum
with the MCD spectrum (Fig. 16.15). The electronic spectrum can be
decomposed into a sum of up to eight individual transitions, as many charge
transfer as d-d (there is separate excitation from each of the components of
d sets which would be degenerate in T• symmetry), a spectrum which has
reasonably well been interpreted by XIX calculations. The lower energy bands
are the ones showing the greatest MCD activity and so are likely to be the
most metal-localized. The blue band also persists when the Cu" is replaced
by Ni" or Co", but it moves into the near-ultraviolet, consistent with a
low energy ligand-to-metal charge-transfer assignment.
In almost 30 blue copper proteins, all of which contain about 100 amino
acids, the amino acid sequence has been determined. About a quarter of the
amino acid sequence is common to all and so this sequence almost certainly
contains the ligands which bond to the copper. In this way, the amino acids
4000
2000
(a)
+0.025
Fig. 18.15 A comparison of (a) the visible and
(b) the MCD spectra of the blue copper protein
plastocyanin.
-0.025
..::; -0.075
0 0 0
-NHCHC-
II II II
-NHCHC- -NHCHC-
1 1 1
CH2 CH2
~N
Fig. 18.18 Chemical structures of amino acids
bonded to copper in blue copper proteins. I I
SH CH2
I
SMe
NH_jj
Histidine Cysteine Methionine
histidine (two), cysteine and methionine (Fig. 16.16) have been implicated
as involved in the coordination. This conclusion has been confirmed by X-ray
studies, studies that have also shown that the site at which the copper is
located changes little when the copper is removed. The copper is at a site
with a ligand-imposed geometry (Fig. 16.17). This at once provides one
reason why it has proved difficult to devise really good model compounds.
Copper(II) is rapidly reduced by thiols
2Cu 11 + 2RS- -+ 2Cu 1 + RSSR
and it is necessary to arrange that this fate does not befall the cysteine, which
contains an RSH unit. Some apparently promising model ligands form stable
complexes with Co" but react with Cu11 •
Another interesting characteristic of the blue copper proteins is the
visibility of the histidine ligands in their EPR spectra. Like many copper(II)
species, the blue copper proteins give strong EPR signals. When there is a
nitrogen ligand bonded to Cu" it is usual for its presence to be indicated by
nitrogen hyperfine structure (see Section 12.6) and this is seen in the EPR
Fig. 18.18 1he EPR spectrum of plastocyanin;
the wiggle pattern at the centre-left of the spectra of blue copper proteins (Fig. 16.18). Finally, EXAFS measurements
spectrum is the nitrogen hyperfine structure. on the proteins provided further evidence of the environment of the copper
Nitrogen fixation 1 401
····························································································································································································
ions at about the same time as the results of the first crystal structure
determination became available. Information on the long-range electron-
transfer properties of the blue copper proteins has come from several sources;
the use of [Ru(NH 3 )s(OH 2 )]3+ is particularly interesting. It has been found
that this ion binds to the surface of the protein surrounding the copper in
at least one blue copper protein, coordinating to an exposed histidine
(presumably losing the H 2 0 ligand) and thus being held almost 12 A
from the copper. Flash photolysis of [Ru(bpyiJJ2+, added to a solution
of the blue protein/rutheniumammine adduct gives [Ru(bpy)J]J+ and the
[Ru(NH 3 ) 5 ]2+ -protein adduct. The [Ru(NH 3 ) 5]3+ adduct is reformed by
electron transfer from the Cu11, which becomes Cu', the protein losing its
blue colour, providing a convenient method of measuring the electron-
transfer process. The observation of a rate constant (about 2 s- 1 ) which is
temperature-independent over the range studied indicates the absence of any
significant activation energy although there is an entropic factor, probably
associated with reorganization of the water molecules.-around the Ru site;
the site of the copper is optimized for electron transfer.
I s--J--Fe
Fe·---s,.. 's(Cys)
subunit being through two cysteines. The other molecule contains two pairs
of clusters, the members of one pair themselves each contain a pair of
ferredoxin [Fe4 S4 ] clusters, bridged through two of the cysteines (Fig.
(Cys)S/
16.20(a)). The other pair of clusters in the second protein also each contain
(c) two bridged clusters, one a modified ferredoxin [Fe4 S4 ] cluster (not all of
Fig. 16.19 The typical environment of Fe in the sulfur-containing ligands attached to Fe are cysteine) and the other a
(a) rubredoxins, (b) plant ferredoxins and
(c) bacterial ferredoxins. 3 J. Kim and D. C. Rees, Science (1992)257, 1677.
402 1 Some aspects of bioinorganic chemistry
i'
Cys
I
/s'-.....
I " ,,
s ---Fe Fez---s
Cys-S~ ...,.s-Cys
,s,.....
c - s ,,,,,\\ """'" ", '- ·.!''''''
ys ' ',, ' ', ""S-Cys
s----Fe Fe~·s
Fig. 16.20 (a) The structure of the FeMe
cofactor Fe8 cluster in nitrogenase. (b) The I
structure of the so-called P-cluster (Fe 7 Mo) in (a) Cys
nitrogenase (the identity of Y is not known).
similar cluster but with one of the Fe atoms replaced with a Mo. The Mo,
perhaps deceptively, is located at one corner of this latter cluster and seems
well away from the site of likely action in the nitrogen fixation process (Fig.
16.20(b)). To what extent had these structures been anticipated by the
spectroscopic studies on the system? Rather well-consider the more
complicated Fe 7 Mo cluster (Fig. 16.20(b)). It had proved possible to isolate
this species, along with enough of its surrounding groups to preserve its
integrity, although it has proved impossible to obtain the crystals needed for
a structure determination. Has the extraction process destroyed the cluster?
EPR spectra of the native protein and the extracted cluster were both similar
and unusual; the cluster had survived. Analysis of the cluster species gave a
Fe:Mo ratio of not less than 6:1 and not more than 8:1, in accord with the
observed 7:1. Mossbauer spectra showed that most, if not all, of the Fe atoms
were involved in the interactions which gave rise to the unusual EPR signal.
Measurements related to the EPR showed that at least five different types
of Fe atom are present; in the proposed structure each of the seven is unique,
although this does not mean that they will necessarily be distinguishable.
Finally, EXAFS not only showed that the nearest neighbours of the Mo are
S and Fe but that it is also coordinated to two atoms which are either N
or 0 (as mentioned above, EXAFS cannot distinguish). For the (average)
Fe, EXAFS revealed, correctly, that each Fe is bonded to three S atoms.
Actually, six of the seven iron atoms present are three-coordinate, a very
unusual situation which must surely be connected with the coordination of
N 2 as the first step of the fixation process. Not surprisingly, this is another
case in which no model compounds have yet been prepared, although one
may be confident that they will. This does not mean that the biological
Protonation equilibria in bioinorganic systems 1 403
K1 K2 K3
[A, B, C] :;=:[A, B, C]H :;=: [A, B, C]H 2 :;=:[A, B, C]H 3
constant kA
fr~:-f? ~
/
A~ ~ tA ft~:/\\ tAH
t -
~
k8 BH B _
k "\
k" -
tAH
BH
•'~ [: /,:~ r~ /
CVCHkA CH
CcH CcH
It follows that
K1 = kA + ks + kc
K2 = 2(kA + k8 + kc)
K3 = kA + ks + kc
Other useful sources are o 'Active-site properties of the blue copper proteins' by A.G.
Sykes, Adv. Inorg. Chern. (1991) 36, 377.
o 'Magnetic circular dichroism of hemoproteins' by M. R.
Cheesman, C. Greenwood and A. J. Thomson, Adv. Inorg. o 'Long-range Electron-transfer in Blue Copper Proteins' by
Chern. (1991) 36, 201. H. B. Gray, Chern. Soc. Rev. (1986) 15, 17.
o 'Natural and Magnetic Circular Dichroism, Spectroscopy on o 'The Synthetic Approach to the Structure and Function of
the Nanosecond Timescale' by R. A. Goldbeck and D. S. Copper Proteins' by N. Kitajima, Adv. Inorg. Chern. (1992)
Klinger, Spectroscopy (1992) 7, 17. 39, 1.
o 'Calculating Equilibrium Concentrations for Stepwise Binding
o Biochemical Applications of Raman and Resonance Raman
of Ligands and Polyprotic Acid-Base Systems' by E. Weltin,
Spectroscopies by P. R. Carey, Academic Press, New York, J. Chern. Educ. (1993) 70, 568.
1982.
o 'X-ray Structure Analysis of FeMo Nitrogenase-Is the
0 EXAFS Spectroscopy, Techniques and Applications B. K. Teo Problem ofN 2 Fixation Solved?' D. Sellmann, Angew. Chern.,
and D. C. Joy (Eds.), Plenum, New York, 1981. Int. Ed. Engl. (1993) 32, 64.
o More recent is a review article 'EXAFS' by H. Bertagnolli o 'The Iron-Molybdenum Cofactor of Nitrogenase' B. K. Bur-
and T. S. Ertel. Angew. Chern., Int. Ed. Engl. (1994) 33, 45. gess, Chern. Rev. (1990) 90, 1377.
o 'Electronic Structures of Active Sites in Copper Proteins: 0 Biocordination Chemistry: Coordination Equilibria in Bio-
Contributions to Reactivity' by E. I. Solomon, M. J. Baldwin logically Active Systems K. Burger (Ed.), Ellis Horwood,
and M. D. Lowery, Chern. Rev. (1992) 92, 521. Chichester (1990).
Questions
16.1 Mushrooms of the Amanita family can contain sur-
prisingly high concentrations of vanadium (in fly agaric up to
325 mg kg- 1 ). It was originally believed that it was present as
the compound shown in Fig. 16.22(a) but, currently, the
compound shown in Fig. 16.22(b) is favoured. Given adequate
supplies of the pure material, what measurements would enable
a distinction between the two alternatives?
16.2 Two spectroscopic methods that have been used in
bioinorganic chemistry but not discussed in the text are
time-resolved infrared spectroscopy and time-resolved resonance
Raman spectroscopy. Speculate on the types of investigation
that might be made with these methods.
16.3 Starting with the A and C term spectra in Fig. 16.12
suggest how each would change as (a) the temperature of
measurement is reduced, (b) the temperature of measurement
is increased and (c) the magnetic field used in the experiment
is increased.
17.1 Introduction
There have been many points in this book at which it has been apparent
that the properties of aggregates of transition metal ions can differ from
those of isolated molecules. This, in some measure, was a theme in the
previous chapter and very much that of Chapter 15; it was met when
discussing magnetic properties (Section 9.11), visible spectra (Section 8.11)
and vibrational spectra (Sections 12.2.1 and 12.2.3). What, then, of the logical
limit, the solid state? There has been an increasing interest in inorganic
chemical aspects of the solid state in recent years, although it is a topic which
is sometimes called material science. The subject covers the whole of the
periodic table, not just transition metal species. Fortunately, the basic
understanding and theory are common to all. It is the purpose of this chapter
to give an introduction to this theory. Although it is a topic widely taught
and studied, it seems the theory of the solid state is not always well
understood by chemists. There are several reasons for this. First, crystalline
materials have a great deal of symmetry, that of the translation operations
that interrelate all the basic building blocks in the crystal (the term building
blocks is preferred to molecules because not all crystals are molecular solids).
The impact of group theory on chemistry is now so great that, given
symmetry, chemists will look for the corresponding character table. Where,
then, are the character tables appropriate to crystalline lattices? They are
rather difficult to find. Not that they do not exist-a fragment of one is
shown in Table 17.1-but the fact that, for an (idealized!) infinite crystal
there is an infinity of translation operations, coupled with the fact that all
character tables with which the chemist is familiar are square (as many rows
as columns), means that we might well expect the full character table to be
of dimensions infinity by infinity! Secondly, usually, chemists have learnt
about crystals from crystallographers. This means that they have ad~pted
408 1 Introduction to the theory of the solid state
T E
0 1 1 1 1
1 eikt e2ikt e3ikt
k
-k 1 e-ikt e -2ikt e -3ikt
1
1
0-C- +
0 nodes 0 nodes
B--E-
+
1 node 1 node
$-[- +
-it-:{ - +
1 node 2 nodes
+
3 nodes 5 nodes
410 1 Introduction to the theory of the solid state
present case, in which our concern is with the solid state, the 'molecule' is an
almost infinite lattice the ends can safely be ignored-as was done in Fig.
17.2. This figure contains the nodal patterns for a linear polyene and these
will now be used to obtain the nodal patterns appropriate to crystals. The
theory we shall develop is applicable to more than orbitals and electronic
structure but, for simplicity, we will confine our discussion to this case and
relate the crystal orbitals to the molecular orbitals of polyenes. Although the
molecular orbitals of polyenes consist of p, atomic orbitals there is no reason
to expect that the crystal orbitals will also be :rc, but for the moment it is the
nodal plane patterns that are of interest, not the precise nature of the orbitals
involved. In the simple picture of a linear molecule presented above, all the
nodal planes were parallel. So, too, in a crystal. Crystal molecular orbital
nodes lie in parallel sets. Consider the two-dimensional crystal shown in Fig.
17.3(a). Despite its clearly finite size, it will be taken to be infinite; its
rectangular shape is meant to indicate that the primitive translation vectors
(the vectors that relate adjacent translationally related points) along the two
different axis directions are not of the same size.
If a crystal molecular orbital had just one nodal plane it could be
represented as shown in Fig. 17.3(b ). Extending this pattern, two nodal
planes could occur in the way shown in Fig. 17.3(c). Equally, however, a
b-y----'
X
(b)
(a)
+
FIJI. 17.3 Some of the possible nodal patterns
in a two-dimensional crystal. The crystal is
indicated in (a); the translation vector in they
+
direction has a magnitude which is one and a
half times that in the x direction. If a phase +
(+ or -) were added to (a) it would show the
no-node combination. One-node combinations
are shown in (b) and (d), a two-node in (c). (c)
(d)
However, nodal patterns are not limited to
those with nodal planes perpendicular to x or y;
a three-node inclined pattern is shown in (e).
\-
\
+
(e)
Nodes, nodes and more nodes 1 411
+ + + + +
~\~_
• • • •
( :::~ +
0,2
~ '----------1
.------------.
+
+
-
+
(a)
Fig. 17.S (a) Some of the nodal pattems of represents a unique nodal pattern and this, of course, is just what irreducible
the two-dimensional crystal of Fig. 17.3 with all representations do also. It is not surprising to learn that there is a one to
nodes perpendicular to either x or y, together
with a dot-map representation of them. (b) The
one correspondence between the dots and the irreducible representations of
dot map of Fig. 17.5(a) extended to include the translational group of the crystal. We may not have full details
nodal pattems inclined to x andy. (c) Figure of these irreducible representations but, at least, we have a way of representing
17 .5(b) corrected for the fact that the
magnitude of the translation along y is 1.5
them; each irreducible representation corresponds to a dot in a figure such as
times that alongx in Fig. 17.3. Here, the Fig. J5.5(c). It is easy to see, in a general way, how to extend the pattern of
separations are in the inverse ratio, 0.67:1. Fig. 17.5(c) into three dimensions. We shall return to this task shortly, when
it will lead us to the important concept of the Brillouin zone.
Travelling waves and the Brillouin zone I 413
1 For a linear polyene such as that shown in Fig. 17.2, some symmetry species appear more
than once. It is then necessary to sum over all functions transforming as the same irreducible
representation before all carbon P. orbitals are found to make equal contributions.
414 1 Introduction to the theory of the solid state
I other. Now, an important point. Although the waves are closely related, they
do not transform as a pair (if they did, they would have a pattern that in a
point group would carry an E irreducible representation label). Remember,
we are working with translations and only translations. There is no
combination of pure translations which either interchanges or mixes the two
wave motions. Functions which transform as doubly degenerate irreducible
representations of a point group are always mixed or interchanged by at
least one operation of a group-this is why they are degenerate? It follows
that the two travelling waves, although related, transform as different
irreducible representations of the translation group.
Secondly, the problem posed by a difference in the number of nodes along
two perpendicular axes disappears. This is demonstrated in Fig. 17.7 which
shows some time-sequenced pictures of a travelling 4 + 2 nodal pattern, most
readily seen for a finite crystal, although the general pattern can clearly be
extended in principle to an infinitely large crystal. The meaning of the
additional points included in Figs. 17.5(b) and (c) compared with Fig.
17.5(a) becomes clear; they correspond to travelling waves moving at an
angle to both x and y axes. An important lesson, therefore, is that in a
crystalline solid one should think in terms of travelling waves, not stationary
ones, because it is possible to attach symmetry labels to travelling waves but
not, in general, to stationary waves.
Fig. 17.7 Time-sequenced pictures of a
travelling wave in the crystal of Rg. 17.3. Such
travelling waves can be along axial directions as 2 The case of separable degeneracy is not covered by this statement but it is true for
all of
well as inclined to them. the point groups that normally concern the chemist.
Travelling waves and the Brillouin zone 1 415
•
•
•
• . .• +
-
• •
•
•Can there ever be stationary waves in a crystalline solid? The answer is
yes, under very special circumstances. So, if, along an axis, the nodal planes
•
exactly bisect interatomic planes, there is no need to invoke travelling waves
•
•
• . .
• • + • •
- •
•
to obtain equal atomic orbital contributions. Each atom is at a point of
•
greatest amplitude (which may be positive or negative) and so all orbitals
appear with the same coefficient although this is either positive or negative,
matching the sign of the amplitude. Pictures for three such waves in the
two-dimensional case are shown in Fig. 17.8. All three satisfy the requirements
imposed on functions which transform as an irreducible representation of a
space group. The example considered is that of a two-dimensional space
group but the conclusion is general. Standing wave irreducible representations
exist only when there is a matching between the nodal separation of the
wave and the magnitude of the corresponding translation vector.
An important point follows. Within the context of the present discussion,
it is meaningless to have wavelengths shorter than those shown in Fig. 17.8 in
• • • • • •
any of the three nodal plane directions shown there. A shorter wavelength
mean that somewhere in space we would have a pattern akin to that
• +
• • +• • •+ would shown in Fig. 17.9. However, in Fig. 17.9, for two of the four regions with
• • • • • • positive amplitude there are no atoms within the region and so no orbitals
• • • • • •
that could have this amplitude! So, those amplitudes are meaningless, and
with them the nodal pattern that gave rise to them. Another way of looking
Fig. 17.8 Some standing-wave patterns for a at this is to compare Fig. 17.9 with Fig. 17.10. Qualitatively, at least, the
two-dimensional crystal. The nodal planes lie at atomic orbital amplitudes are the same in Figs. 17.9 and 17.10, and a detailed
exactly half the distance between atomic layers. study shows that it is not difficult to ensure that they are also quantitatively
the same. If the pattern in Fig. 17.9 is not meaningless, it certainly is
• -1 I. • .I •
redundant. We see, then, that nodal patterns such as those shown in Fig .
17.8 are the limit that we need to consider, be it two-dimensional or
• .I • • • • three-dimensional space about which we are talking. Such nodal patterns
+ + + + can be incorporated in dot maps of the type given in Fig. 17.5(c)-enormously
• • • • • • extended-and represent the limit of the dot patterns. Further dots can be
• • • • •I • added outside this limit but, within the present context are either meaningless
or redundant. 3
Wave patterns in very similar directions will terminate at very slightly
displaced, adjacent, points. A family of such points will define a surface which
Fig. 17.9 A wave pattem in which the
separation between nodes is less than the turns out always to be planar, except where two or more such planes
separation, in the same direction, between intersect. Most important is the three-dimensional case, which we now
atoms. The result is that in two regions
(arrowed) there are phases but nothing to which
consider and which follows, without complications, by extension of the two
these phases can refer. dimensional. The three-dimensional limiting surfaces define some rather
beautiful shapes, the shape depending on the particular lattice under
consideration. Examples of some of them are given in Fig. 17.11. The shapes
• • • • • • shown in Fig. 17.11 are pictures of the Brillouin zones ofthe solids. A Brillouin
zone can be thought of as containing a multitude of closely packed points,
•+ • • +
• • • + each with a unique set of labels and each point corresponding to a unique
• • • • • •
• • • • • • 3 The present context is one in which there is a single set of points, presumably representing
atoms, interrelated by translation vectors. Had there been units made up of several points,
Fig. 17.10 A wave pattern which gives the grouped together and interrelated between themselves by, say, point group operations, then the
same amplitudes as those in Fig. 17.9 but which 'empty spaces' in figures. such as Fig. 17.9 would not be empty. In such a case there would be
has an internode separation which is greater a distinction between Figs. 17.9 and 17.10; the additional node patterns would be neither
than the interatomic separation in the same meaningless nor redundant. They are associated with the second, third, and so on, Brillouin
direction. zones.
416 1 Introduction to the theory of the solid state
@~
equivalent of an irreducible representation in a point group character table,
so the Brillouin zone, containing all points, is, in a sense, the equivalent of
a translation group character table. As we have seen, the further that one of
the closely packed points is from the origin (the origin is at the centre of the
Brillouin zone and is the no-node combination), the shorter the distance
OJ@
between nodes in the corresponding travelling wave pattern in real space.
Because of this reciprocal relationship one commonly talks of reciprocal
space; all Brillouin zones are in reciprocal space. Conventionally, the vector
drawn from the origin to a particular point is called the 'k vector' for that
point. As we have also seen, for every vector k there exists a vector - k. The
origin is often referred to as k = 0 in phrases such as 'At k = 0, .. .'. It is
important because k = 0 is the equivalent of the totally symmetric irreducible
p@
representation of a point group. As a result, just as for point groups, the
spectroscopic selection rules for crystals correspond to non-zero integrals
and this means integrals that they transform as k = 0, which is why phrases
such as at k = 0, ... will be encountered. The label k, with a suitable plethora
of subscripts to indicate the point in the Brillouin zone to which it refers,
is the equivalent of a point group irreducible representation label such
as A" B 2 ., ••• with which the reader will be familiar. Not surprisingly,
instead of reciprocal space the term 'k space' will be encountered. Just to
FJC, 17.11 Drawings of the shape of some add a bit of physics, and remembering that we are talking about travelling
Brillouin zones: (a) face centred cubic, waves, the name momentum space will also be found in books on solid
(b) trigonal, (c) tetragonal, (d) orthorhombic, state physics.
(e) monoclinic and (f) triclinic. In each case an
axis of highest symmetry is arranged It is to be emphasized that in the above discussion we have been talking
approximately vertically in the plane of the about translations, and only translations. A particular pattern of translations
paper. In principle, there is one general shape may well lead to the automatic generation of rotational axes. So, it takes
of Brillouin zone for each of the 14 Bravais
lattices but, because of the freedom allowed by nothing but translation operations to generate a cubic lattice. However, when
changes in relative axis size for the lower we look at the lattice so generated we find that it also has threefold, fourfold
symmetry groups, there is a total of 24 and twofold rotation axes. These are something additional. For instance, the
qualitatively different shapes.
way that a cubic lattice is brought into coincidence with itself by a C4
rotation, the way that the points map onto one another, cannot be replicated
by any combination of translation operations applied to the lattice. Inclusion
of operations other than pure translation operations is important for those
interested in the details of the theory of space groups, but is not essential
for our discussion. The lattices are always of the highest symmetry compatible
with any crystal type,4 be it cubic, tetragonal, orthorhombic, hexagonal,
trigonal, monoclinic or triclinic; addition of point group operations can never
increase the symmetry although the absence of such operations can lower
it. Clearly, it is convenient to restrict discussion to the high-symmetry cases.
These cases have given us the Brillouin zone and a myriad ofk vectors-and
that's enough! The point group operations serve to interrelate k vectors that
are otherwise distinct. So, in any cubic lattice, invariably of o. symmetry, a
k vector in a general position ink space has 47 symmetry-related equivalents
(there are 48 operations in the o. point group). Diagrams showing energy
levels across a Brillouin zone (which will be met shortly) would only show
one of the 48.
4 For instance, all lattices are centrosymmetric, contain centres of symmetry, although not
all crystals are centrosymmetric.
Band structure 1 417
I I I I
~~~~~~
~~~~~ + I +
i
I I
I~
+ + I +
~~~ ~
I
~I~
k = x* edge; a,(x)
~
example described in the text. To the left is the
I
zone-centre pattern and two zone-edge patterns X I
to the right. The type of bond1ng (u or n) is
indicated together with whether the interaction I
is bonding (,) or anti bonding(,) and the axis k = y* edge; aa(X) + 1ta(Y)
along which these interactions are directed.
~ ~ ~
~ ~ ~
i ~ ~ ~
-x• y*-
Density of
Brillouin k=O Brillouin states (DOS)
zone edge zone edge
Band structure 1 419
distinguish them from x and y.6 Because the change from bonding to
antibonding is expected to have larger energetic consequences for u than 1t
interactions, the slope of the plot to the x* zone edge is greater than that to
they*. For simplicity, the interpolations from zone centre to edge have been
made linear, although in reality they would be curved. To the right of Fig.
17.13 is shown the resulting density of states plots (DOS; these were first
met in Fig. 15.25). It is clear that what was an individual level in the isolated
atom has become a band in the solid-a somewhat unrealistic rectangular-
sectioned band, but a band nonetheless. Further, the energy varies across
the band (in jargon, the 'dispersion of the band'); the energy varies with
position in k space. The solid-state physics literature is full of plots of
dispersion curves, usually along directions in reciprocal space that correspond
to high-symmetry directions in the crystal; we shall meet examples shortly.
It is evident that the greater the interaction between adjacent translation-
related orbitals, the broader will be the resulting band. Inner orbitals will
usually give narrow bands, outer orbitals will tend to give broad bands.
Increase in pressure on a crystal will tend to increase the interorbital
interactions and thus to broaden the bands, this not only affecting the
outermost orbitals but, with sufficient pressure, the inner ones too. So, it is
calculated that the ls orbitals of atomic hydrogen give rise to a band such
that hydrogen becomes a metal if it is subjected to ca. 10 6 atmospheres
(10 11 N m - 2 ) pressure. Returning to our example, if the Px orbital of A was
empty, so too will be all levels of the band that have been generated from
it. If the Px orbital contained a single electron the corresponding band would
be half-filled (if the bottom half of the band was filled then the electrons
would be paired up). If the A Px orbital were filled, so too will be the band.
Unless, of course, something happens to change the situation.
There are two important things that can happen, the first of which will
have to be discussed at some length. This arises from the recognition that
bands adjacent in energy can overlap with each other. So, if a band
corresponding to a doubly filled A orbital overlaps with that corresponding
to an empty A orbital, what results is a single band (really, composed of two
overlapping bands) which is half-filled. Although such overlapping is likely
to be of common occurrence, as we shall see, one has to be careful-
overlapping can be avoided. In the example considered above it was seen
that interactions between Px orbitals changes from antibonding to bonding
with increase ink along the x* axis. As shown in Fig. 17.14, the corresponding
change in s orbitals is the opposite, from bonding to antibonding. Now, it
is generally true that s atomic orbitals lie below the corresponding p, so the
different behaviour of s and Px along x* means that they are potentially set
on a collision course! In fact, they interact; this interaction can lead to
a mutual repulsion, a so-called avoided crossing, as shown in Fig. 17.14 (the
energy levels do not cross although the wavefunctions behave as if they did).
In such cases, increase in pressure increases the separation between two
bands. Although avoided crossings are not uncommon, more general is band
overlap. Such overlap is not limited to pairs of orbitals-there is no limit
on the number of energy bands that can overlap with each other. Just as in
6 In the general case, where x and y are inclined at an angle of other than 9W, care has to
be taken in defining the corresponding axes in k space.
420 1 Introduction to the theory of the solid state
Zone Zone
centre edge
Zone edge
along y*
Zone edge
along x*
commonly referred to as Bloch functions , although the use of this term is not
confined to electronic wavefunctions- it applies to any function which
transforms as an irreducible representation of the translation group) have the
same symmetry. That is, they have the same k vectors. It follows that such
interactions are not required to be constant across the Brillouin zone. For
example, the energy difference between two Bloch functions at k = 0, the
zone centre, may be very different from the corresponding energy difference
at a point on the surface of the Brillouin zone. It is worthwhile at this point
to emphasize that a Brillouin zone is not, of itself, an energy surface, although
such an interpretation is sometimes encountered. Equally, it is not a hollow
shell, although sometimes drawn in this way. As we have seen, it is perhaps
best regarded as a compact- and perhaps solid - mass of points of the sort
shown in Fig. 17.5(c) in which each individual point represents a unique
irreducible representation.
The next example is more complicated; it is copper, each atom having six
valence shell orbitals (five d and one s~the configuration of the isolated Cu
atom is 3d 1 0 4s 1 ) to be included together with, of course, 11 electrons. In
the language of Chapter 7, there is only one hole to consider. Copper has
the face centred cubic (fcc) lattice, the Brillouin zone of which was shown in
Fig. 17.11(a). This is repeated in Fig. 17.17(a) with the labels of special
positions indicated. In Fig. 17.17(b) are shown the results of calculations on
the energy bands in copper and their occupation at 0 K; the Fermi level is
indicated. Figure 17.17(b) merits close inspection. Most evident are the peaks
at W and K; they correspond to the antibonding combination of the 4s
orbitals at the zone surface. The band of levels lying together originate in
the 3d orbitals, although the way that they get mixed with the 4s is evident
on close inspection. Thus, the line representing the peak at W is not
continuous with that representing the peak at K; an avoided crossing has
occurred. Even this is not simple~the single low-lying level at r involves
the bonding combination of 4s orbitals. Of particular interest is the way that
at L, all nearby points in the Brillouin zone lie below the Fermi surface, the
only labelled point in Fig. 17.17(b) apart from r so to do. The consequence
of this is evident in Fig. 17.17(c), which shows the Fermi surface for copper.
424 1 Introduction to the theory of the solid state
--
magnetic field is orientated as indicated as
shown in (b). The two periods of oscillation may
be correlated with the concurrent existence of 1
the two different circulatory pathways shown. (a) H
A different orientation of the crystal with respect
to the magnetic field would lead to a different
oscillation pattern and thus provide information
on the topology of the Fermi surface (adapted
from Ashcroft and Mennin).
will be made and then the above discussion will be applied to some topics
covered in earlier chapters. Clearly, when we turn to complexes, we have
to start with the AB pattern, covered in the previous section, and develop
it. The relevant situation is one in which either or both of A and B are
coordination compounds. Were we to follow the pattern developed so far in
the chapter, we could consider each atomic orbital of every atom of A and
do the same for B. This, however, would not be sensible. A major theme of
many earlier chapters has been the exploitation of the molecular orbitals
appropriate to A and/or B when these are coordination compounds. The
question which immediately faces us is the extent to which these discussions
have to be modified by the introduction of interactions between translationally
related sets of A and also of B. It is not difficult to see the general
outcome. When an orbital of a coordination compound, A, is largely located
on a central metal atom, then the intervening ligands serve to insulate these
426 1 Introduction to the theory of the solid state
orbitals from each other. The orbitals do not interact and the discussion of
this chapter becomes largely irrelevant. The crystalline material behaves like
a set of superimposed atoms, each unaware of the others' existence. Such a
behaviour is much less likely to occur for molecular orbitals delocalized over
the whole molecule. Here, the contribution from orbitals on the periphery,
on the 'outside' of the complex, is important. Such orbitals enable an inter-
action between adjacent molecules, the interaction giving rise to dispersion
patterns of the type that have been discussed. Evidently, some-external-
ligand orbitals will be the most affected.
We can at once understand two things. First, why ligand field theory
works so well (in a typical coordination compound the d orbitals of the
transition metal ions are physically well separated). Secondly, why our
understanding of ligand orbitals is much less detailed and less general than
that of metal ions. The energy levels that ligand orbitals subtend are likely
to be environment-sensitive. The (complex) species B will only partially
insulate the ligands of one (complex) A molecule from those of another.
Change B and these residual A-A interactions will also change. Further,
interactions (strictly, over k-space) between ligand orbitals of A and external
orbitals of B may occur.
With this discussion in mind, it is helpful to reconsider the distinction
between charge-transfer bands introduced by Robins and Day, the class 1,
class 2 and class 3 dis<inction that was discussed in some detail in Section
8.11. An example of class 2 behaviour (clear interaction but chemical identity
retained) is provided by the coordination compound KFe"[Fem(CN) 6 ).
It can at once be understood why the Fe" and Fem ions retain their
individuality. They are insulated from each other by the eN- ligands.
However, a little consideration suggests that it is a delicate matter. They
have t~ 9 and t~ 9 configurations, respectively, and, following the discussion
of Section 6.6, these t 29 electrons are partially delocalized into the CN-
n antibonding orbitals. It follows that they have a non-zero probability of
being at the periphery of the complex. A mechanism for metal-metal
interaction does exist; evidently, its effect is small. Of course, although this
explains in outline why this compound is class 2 and not class 1 (no
interaction), there is ample evidence for small metal-metal interactions of
the type that have just been considered. Their effects are most important in
magnetochemistry and were discussed in Section 9.11. A typical class 3
compound is Ag 2 F (Table 8.4). The contrast with KFe[Fe(CN) 6 ] is clear.
In the crystal, the orbitals involving the Ag and Ag + ions must be delocalized
over the entire 'molecule' (Ag 2 F has an extended, not a molecular, structure).
Interactions between the silver ions and atoms will surely occur, leading to
dispersion, the observed broad band absorption and a black compound
results. Any characteristics of Ag and Ag+ are completely lost.
There are aspects of the discussion in the last paragraph which can be
extended well beyond the compounds covered by the Robins and Day
classification. Consider the following problem. NiO, when pure, is a pale
solid. TiO and VO both have the same structure as NiO but are black and
are almost metallic conductors of electricity, whereas NiO, when pure, is an
insulator. That there is some electronic explanation for this pattern is made
evident by the observation that when suitably doped with an impurity (Li 2 0
is the one usually cited) NiO also becomes black and highly electrically
Solid state and coordination compounds I 427
conducting. What is the explanation for this pattern? To avoid the problem
of the electronic levels introduced by an impurity we will consider just the
origin of the differences between TiO, VO and pure NiO. All three
compounds crystallize in the NaCl, rock salt, structure-which means that
each metal ion is octahedrally surrounded by oxide anions, so that the
ideas of simple crystal field theory should be applicable. The electronic
configurations of the three ions are d 2 , d 3 and d 8 , respectively. In an
octahedral crystal field these will become ~~., ti. and ~~.e;. Only the first
two have incompletely filled t 2• shells. As Fig. 17.19 shows, in the TiO
structure overlap can occur between such t 2 • orbitals, leading to an
incompletely filled band structure and an explanation for the electrical
conductivity. For NiO, a comparable band structure will be less available.
First, the partially occupied d orbitals are e., which do not overlap
each other in the same way-they point towards an oxygen, not a nickel; the
electrons are more likely to be localized on the metal atoms. Secondly, the
band structure in NiO will involve participation of the 0 orbitals. Finally,
it turns out that the Ni d orbitals are a bit more contracted than are the Ti
or V, reinforcing the idea that they will tend to be relatively insulated from
one another. We can confidently predict that NiO will have a lower electrical
conductivity than the other two oxides. As we will see, the difference in colour
also follows. One important point: because it was important to show the
interaction between several metal atoms, Fig. 17.19 does not show just a
single primitive unit cell. When multiples of the primitive unit cell are shown
in this way it is a good idea to check whether the interaction between
adjacent units is in-phase (as here). If so, then moving to the zone boundary
the interactions will become out-of-phase and so the interaction shown in
Fig. 17.19, at k = 0, at r, is at the bottom of the band (assuming that overlap
with other bands can be neglected).
Some generalizations have been made about those features which enhance
the probability of band formation of the type just discussed in transition
metal compounds:
9 A problem arises for some members of a series of related crystal structures for some
compounds- ZnS and CSi (carborundum) are well-known examples-in which translational
vectors of up to thousands of Angstroms can occur (so-called polytypism). However, the
long-range communication between unit ce11s which must be involved may be rather differently
mediated although the mechanism is not known (for a speculation on one possibility see S. F. A.
Kettle, J. Cryst. Spect. Res. (1990) 20, 59).
10 It is important to remember that the theory developed in this section concerns only pure
translation operations. It contains no rotation or reflection operations; inclusion of these into
the theory is required to explain the anisotropies associated with the absorption and emission
of radiation by real crystals.
430 1 Introduction to the theory of the solid state
t
E
(c)
0 ki~
(Ground (b)
state)
-k;-
(Excited individual structural units becomes important (as in TiO and VO) the band
(a) state)
structure of the solid means that quite a different approach has to be adopted,
Fig. 17.20 (a) The rather different dispersions one that takes account of the dispersion of the energy bands across the
of the ground and excited states involved in an Brillouin zone. Clearly, success in this latter endeavour is more likely to come
electronic transition, together with (b) the for simple species-and these are just the ones most likely to show the phenom-
corresponding density of states. The resu~ing
spectrum is shown as (c); the observed maxima ena which require it. Not surprisingly, the work published by those concerned
and minima can be correlated with features in with simple species tends to be rather different in content and interpretation
the dispersion curve. ln principle, similar to that published by those working with large coordination compounds.
patterns apply to all tonns of spectroscopy
carried out on solids, not just electronic Throughout this chapter the assumption of an infinite, perfect and pure
spectroscopy. crystal has been made. In general, these assumptions are sufficiently valid
for the general results obtained to apply to real-life crystals. Two important
exceptions are to be noted, however. The first has already been mentioned
but not developed (it would require a chapter of its own). This is the presence
of impurities. If these, usually by intention but sometimes by accident, have
energy levels which interleave the band gaps in the host material then the
properties of the host material are dramatically changed. The model
developed in this chapter is an adequate starting point for a discussion of
these changes but, of itself, is not adequate to deal with them. The assumption
of an infinite and perfect crystal is usually valid. When the size either of the
crystal or of the perfect blocks within a crystal become too small, however,
dramatic changes occur. For instance, cadmium phosphide, Cd 3 P2 , is a
semiconductor with a band gap of only 0.5 eV. Not surprisingly, it is black
(the dark colours of many main group element sulfides similarly reflect
small band gaps). If small particles of the compound are prepared as colloids
and then separated the colour of the product depends on the particle size.
So, at 30 A they are brown; as the particle size diminishes further, in rapid
succession they become red, then orange, then yellow and finally white at
about 15 A. The band gap has increased to 4 eV. These colour changes have
nothing to do with the lighter colours commonly displayed by coloured
materials when they are finely ground (this is a purely optical effect which
may be cancelled by changing the refractive index of the material surrounding
the finely ground crystals). Such particle size effects, and the breakdown in
the simple model presented in this chapter, are likely to become of increasing
importance as the search continues for molecular electronic devices.
Questions 1 431
Fig. 17.21 Question 17.1. 17.4. The Brillouin zone for the simple (primitive) cubic
lattice is a (simple!) cube. The Brillouin zone for the face-
17.2. When studied in the crystalline material the Raman centred cubic lattice is pictured in Fig. 17.11; that for the
band associated with the totally symmetric (breathing) v(Cr...CO) body-centred cubic is shown in Fig. 17.16. Show that all three
mode of Cr(C0) 6 was little different from the same band when have 0, symmetry.
Appendix
c c
Q:,b r1;.0
\I M \ I
c c
=P- _ \'_
FIJI. AL2 The 2 and b ring configurations.
L-CJ---~---~' L--~-~--~---~\
A
Fig. AL3 The A and 6 configurations of three
chelate rings in an octahedral complex.
( J
configurations is obtained, all of which must be assumed to be present in a
solution of a complex such as [Co(enh]3+:
A(J.,V,) A(Mb)
A(J.J.b) t.(bbJ.)
A(lM) A( b).).)
A(bM) t.().I.J.)
The first member of each pair (those on the same line) is the mirror image
of the second (they are enantiomorphic pairs). Generally speaking, members
of families of such compounds rapidly interconvert. However, they have been
extensively studied, usually by NMR (often with somewhat more exotic
ligands than en) because they provide an opportunity of investigating the
434 I Appendix 1
Comparison with the geometries which are discussed in Chapter 3 will show
that for main group elements these predictions are rather good. So,
[BF3 · NMeJ , in which there are four valence electron pairs around the
boron, is approximately tetrahedral, the distortion being consistent with the
436 1 Appendix 2
fact that one of the valence electron pairs (that involved in the B-N bond)
would not be expected to have exactly the same spatial distribution as the
other three, so that (B-N)--(B-F) bond repulsions will be slightly different
from (B-F)--(B-F) repulsions. Similarly, extending the discussion to a
compound which would not normally be regarded as a complex, the oxygen
atom in water is surrounded by four electron pairs (two from the 0-H bonds
and two from the lone pairs on the oxygen atom). Repulsion between these
electron pairs leads to a tetrahedral distribution and, in accord with this,
the H--()-H bond angle in water has roughly the tetrahedral value, the
deviation from regularity being in accord with the suggestion that the centre
of gravity of the electron density in the 0-H bonds is further away from the
oxygen atom than is the case for the lone pair electrons. Such arguments
lead to the prediction that the relative magnitude of the electron repulsion
is in the order:
lone pair-lone pair> lone pair-bonding pair> bonding pair-bonding pair
Table A2.1 Change in energies between planar and most stable pyramidal geometries of NH 3 and NF3 . Positive quantities mean
that the pyramidal form is the more stable. All energies are in atomic units (1 au = 2626 kJ mol- 1 )
barrier-and hence the reason why the molecules are pyramidal rather than
planar, for each of the three energy terms identified in this appendix-bonding,
electron-electron repulsion and nuclear-nuclear repulsion-the change in
going from a planar to a pyramidal molecule was calculated. These data, the
numbers in Table A2.1, can then be used to understand why NH 3 and NF3
are pyramidal rather than planar. In Table A2.1 a positive quantity means
that this contribution stabilizes the pyramidal molecule relative to the planar.
The first thing to be noted about Table A.2.1 is that the entries in the
'change in bonding energy' column are non-zero. Contrary to the VSEPR
model's assumption, bonding is a function of bond angle. Even more striking
is the different pattern of the signs of the entries in the 'change in bonding
energy' and 'change in repulsive energy' columns. The lesson is a salutary
one. The sum of the two entries for NH 3 and also for NF3 is positive-the
pyramidal molecule is the stable one-but the two molecules achieve this
result in quite opposite ways. Pyramidal NH 3 is stable because this geometry
minimizes the repulsive forces-bonding is a maximum in the planar
molecule. In contrast, NF3 is pyramidal because this geometry maximizes
the bonding; repulsive forces are a minimum in the planar molecule, in
contradiction to the VSEPR model's predictions. The most charitable thing
that can be said about the second set of entries in Table A2.1 is that they
show that electron-electron and nuclear-nuclear repulsion energy changes
are of comparable importance in determining molecular geometry (for NH 3
the nuclear-nuclear are the more important). It can be argued that the
electron- electron terms in Table A2.1 include all the electrons in the
molecules and not just those which are the focus of the VSEPR model.
However, such an objection misses a key point. If other electron-electron
repulsion interactions are important and yet ignored by the model then this,
too, is a weakness. One should not argue too much from one example, yet
that which we have given is a dramatic one-in both cases considered,
electron-electron repulsion makes a smaller contribution to determining the
molecular shape than do either bonding or nuclear-nuclear repulsion. One
is forced to remember a difficult lesson; a model that leads to a correct
prediction is not necessarily a correct model. A further indication of the
existence of concealed assumptions within the VSEPR model is provided by
some other molecules related to ammonia. The C-N-c bond angle in
triisopropylamine, N(CHMe 2 )J, is 119.2°-the N-c 3 framework is essen-
tially planar, presumably because of steric interactions between the isopropyl-
amine groups. In trimethylamine, N(CH 3 )J, the C-N-c bond angle is 111 o
but in the corresponding fluorinated compound N(CF 3 h it is 118°. This
change is surprising because fluorination has little effect on bond angles in
alkanes. In N(C 2 F5 h the C-N-c bond angle is 119.3°. Clearly, there are
factors of importance in determining these angles that are not included in
the simple VSEPR model. Nonetheless, the value of this model as a usually
correct predictor of the molecular geometries of compounds of main group
elements is unquestioned. Further, the sort of analysis which has been given
above for NH 3 and NF3 is not the only one possible. It has been much more
common to ask a different question: as to whether there is theoretical support
for the electron repulsion inequality sequence:
Most workers who have considered this question have concluded that there
is. Crucial is the way that a single entity, the electron distribution in a
molecule, is divided up to give distinct lone and bonding pairs. What has
usually been done is to adopt some criterion which seems to lead to a division
as close as possible to that made, more qualitatively, by the VSEPR model
itself.
We have reached a point at which opinions differ; the reader is entitled
to his or her own. Those who would like the benefit of a view which is
intermediate between the two extremes presented in this appendix might find
it helpful to read an article by A. Rodger and B. F. G. Johnson in Inorg. Chim.
Acta (1988) 146, 37. The most recent exposition of the VSEPR model is to
be found in The VSEPR Model of Molecular Geometry by R. 1. Gillespie and
I. Hargittai, Prentice-Hall (Allyn and Bacon), New York, 1991.
Appendix3
Introduction to group theory
Figure A3.1 shows a square planar complex and something of the symmetry
which it possesses. Most evident are the pictorial representations of mirror
planes- reflection in any of these infinitely thin, double-sided mirror planes
turns the square planar complex into itself. The solid black shapes in the
figure indicate rotational axes-- shaped for twofold and • for fourfold.
Although Fig. A3.1 shows symmetry elements- mirror planes and rotational
axes-in group theory it is the corresponding operations which are of interest;
the consequences of the act of reflecting or rotating. If the operations
indicated in Fig. A3.1 are counted one finds a total of 16. In Table A3.1
these 16 operations are listed across the top. Some of the operations which
are very closely related are grouped together- rotation by ± 90° about the
c4 (fourfold rotation) axis; rotation by 180° about equivalent c2 (twofold
rotation) axes; reflection in equivalent mirror planes (denoted a.). Symmetry
operations which are grouped together are said to be in the same class. Each
number in the table is known as a character and the whole table is called a
character table. This group of 16 symmetry operations is called the D4 , group,
and the character table is the D 4 , character table. Down the left-hand side
of the character table are listed some of the symmetry symbols used in the
text (in Table 7.2, for instance). All such symbols originate in character tables.
Each symmetry symbol is associated with a unique row of characters-no
A'<~ 1 1 1 1 1 1 1 1 1 1
A2S 1 1 1 -1 -1 1 1 1 -1 - 1
B'<~ 1 -1 1 1 -1 1 -1 1 1 -1
B2g 1 -1 1 -1 1 1 -1 1 -1 1
Eg 2 0 -2 0 0 2 0 -2 0 0
Alu 1 1 1 1 1 -1 -1 -1 - 1 -1
A2u 1 1 1 -1 -1 -1 -1 -1 1 1
Blu 1 -1 1 1 -1 -1 1 -1 -1 1
B2u 1 -1 1 -1 1 -1 1 -1 1 -1
Eu 2 0 -2 0 0 -2 0 2 0 0
Appendix 3 1 441
A pair of equivalent
mirror planes, a ' ll
Horizontal
mirror plane, "•
The other C2
A pair of equivalent
A pair of equivalent two-fold axes, C2. two-told axes, Ciz. The second is
The other is perpendicular to this one. on the other side of the diagram.
FIC. A3.1 Symmelly elements and operations two rows are identical, although inspection reveals that the characters block
associated with a square planar complex. There into four 5 x 5 squares, the characters within one square being very simply
are two other symmetry operations, not shown
on the diagram. The first, apparen~y trivial related to the corresponding characters within another. This fundamental
operation, is 'leave everything alone ' and is building unit of the D4 , character table is, in fact, the character table of the
denoted E. The other is the operation of D4 group (which has only the E, 2C 4 , C 2, 2c; and 2C~ operations of the
inversion in the centre of symmetry (which is at
the central atom) and is denoted i. D4 , group). In Table A3.1 any one character corresponds to a symmetry
symbol, (its row) and to one or more symmetry operations (its column). The
character has the property of telling us how something which carries its
particular symmetry symbol behaves under its particular symmetry operation.
So, something of A 2 • symmetry is multiplied by I (i.e. turned into itself) both
by a C 4 rotation (be it clockwise or anticlockwise) and also by the C 2
rotation. However, it is multiplied by -I (i.e. turned into itself with all signs
reversed) by the c; and q rotations (any of either pair of operations). The
442 I Appendix 3
&u
z
p, orbital shown in Fig. A3.2(a) has A 2 u symmetry and the reader should
check that it behaves under these operations in the way that has just been
described. Those symmetry symbols which, for the leave-alone operation (E),
have characters of 2, describe the behaviour of two independent things
simultaneously. The number 0, for these species, means either that the two
things are interchanged in some way or that one is changed into itself ( + 1)
and the other into the negative of itself ( -1) so that, together, a character
of zero is obtained ( + 1 - 1 = 0). The number -2 means that under the
symmetry operation each of the two objects is turned into itself with all signs
reversed. So, Px and p, (Fig. A3.2(b)) together have E" symmetry. The reader
should check that the D4 • character table gives an accurate description of
their behaviour under the symmetry operations.
The direct product of two symmetry species, say A 29 x B 19 , is formed by
multiplying together in turn their characters under each symmetry operation,
thus
A2g 1 1 1 -1 -1 1 1 1 -1 -1
a,. 1 -1 1 1 -1 1 -1 1 1 -1
A2g x a,. 1 -1 1 -1 1 1 -1 1 -1 1
Comparison with Table A3.1 then sht>ws that the set of characters produced
is, in fact, that labelled B 29 • We say that 'the direct product of A 29 and B 19
is B 2 ;. ForE-type symmetry species direct products are a bit more difficult.
Consider E 9 x Eu:
Eg 2 0 -2 0 0 2 0 -2 0 0
Eu 2 0 -2 0 0 -2 0 2 0 0
Eg X Eu 4 0 4 0 0 -4 0 -4 0 0
Appendix 3 1 443
Alu 1 1 1 1 1 -1 -1 -1 -1 -1
A:~g 1 1 1 -1 -1 -1 -1 -1 1 1
alu 1 -1 1 1 -1 -1 1 -1 -1 1
a2u 1 -1 1 -1 1 -1 1 -1 1 -1
gxg=UXU=g; gxu=u
Notice that in Table A3.2 A 1 only appears on the diagonal and that it appears
on every diagonal element; had we been working in the group D 4 h then it
would have been the A 19 irreducible representation which would have
behaved in this way. This is important; it can be seen from Table A3.1 that
A 19 is the only symmetry species which never goes into its negative under a
symmetry operation. If we were interested in an integration over all space
and the mathematical function being integrated had A 19 symmetry then it
o. A, A2 a, a2 £
A, A, A2 a, a2 £
A2 A, A, a2 a, £
a, a1 a2 A1 A2 £
a2 a, a1 A, A1 £
£ £ £ £ £ (A 1 + A2 + a 1 + a2)
444 I Appendix 3
would not automatically equal zero. If it had any symmetry other than A,.
Table A3.3 The 0 character table
then the negative and positive contributions to the integral would exactly
0 E 6C4 3C2 6C2 8C3 cancel (this is easy to see for all except the Es in Table A3.1; for all the other
As and Bs there are as many -I entries as there are +I; the statement is
Al 1 1 1 1 1 also true for the Es). This conclusion is very relevant to quantum mechanics
A2 1 -1 1 -1 1 for there one is frequently interested in integrals over all space. So, if we are
E 2 0 2 0 -1
1 -1
interested in an integral such as
rl 3 -1 0
T2 3 -1 -1 1 0
f 1/1 1 1/12br
which is an overlap integral, we can discuss its symmetry properties (in
particular, whether it is zero or not) by considering the direct product of the
symmetry species of the two orbitals concerned. In this particular case it
leads to the important result that only orbitals of the same symmetry species
have non-zero overlap integrals. So, in those places in the text that direct
products have been used, it is because, really, we are interested in products
of orbitals; this is true even in places such as Section 8.7 where, ostensibly,
the subject is the intensities of spectral bands.
Finally, in Table A3.3, is given the character table for the octahedral group
0. It can be used to check that tables such as Table 6.3 are correct. The full
octahedral group, which has symmetry species with g and with u suffixes, is
denoted o. and is four times as large. The two groups are related in that
the character table for the group o. is built up from that of the group 0 in
the same way that the character table of the group D4 • is built up from that
for the group D4 .
The reader will find simple applications of group theory in Chapters 10
and 15. A non-mathematical introduction to the subject is given in Symmetry
and Structure by S. F. A. Kettle, Wiley, Chichester and New York, 1995. Also
relevant, useful and/or relatively non-mathematical are: Symmetry in Coordi-
nation Chemistry by J. P. Fackler Jr., Academic Press, New York, 1965;
Symmetry in Chemistry by H. H. Jaffe and M. Orchin, Wiley, New York, 1965;
Molecular Symmetry and Group Theory by A. Vincent, Wiley, London, 1977;
Symmetry in Chemical Bonding and Structure by W. E. Hatfield and W. E.
Parker, Merrill, Columbus, Ohio, 1974; and Group Theory for Chemists by
G. Davidson, Macmillan, London, 1991. More mathematical is Chemical
Applications of Group Theory by F. A. Cotton, Wiley, New York, 1990.
Appendix4
Equivalence of dz2 and dx2 _y2
in an octahedral ligand field
y2
- z2 = - .1(x2
2 -y 2) - ...}3
2 x.J13 (2z2
- x2- y 2)
(hint: show that simplification of the right-hand side expression gives the
left-hand side) and
1 ...}3
- (x 2 - y 2 ) - -1 1
2 2 2 2 2
- (2x - y - Z ) = X - (2z -X - y 2)
...}3 2 2 ...}3
That is,
and
In other words the new orbitals are simply linear combinations of the old.
The readers should check that this is also true of d,' - x' and d,, (they
only have to change two of the signs in the right-hand side expressions
given above).
A final word, for the reader of Appendix 6, who has returned to this
appendix in search of a character of -I. As shown above, when dx, _,, is
rotated into d,, _,, its contribution to the latter is --!- Similarly, d,,
contributes --! to dx'· As the reader may well have discovered, the same
result is obtained if rotation into d,, -x' and d,, is considered. In both cases
the aggregate character is -I.
Appendix 5
Russell-Saunders coupling
scheme
in the way we have just described, the arrangements will also be related.
Note that if we wish to hold the z axis fixed in space we have to apply some
fixed axial perturbation, an electric field, for example.
So far, the nature of the vectors has not been specified. They represent
angular momenta (there is more on this topic in Chapter 9) and one talks
about the orbital and spin angular momenta, their magnitudes and their z
components. The quantization of orbital angular momentum, introduced
arbitrarily in the Bohr theory of atomic structure, appears in present-day
quantum mechanics in the way just described. The various z components
of orbital and spin angular momentum are, within our approximation,
energetically unimportant. The energy of a particular arrangement is deter-
mined solely by the absolute magnitude of the orbital and spin angular
momenta, and is independent of their z components. The reason that the z
components have been introduced is the following. For each of the (2L + 1)
z components of an orbital angular momentum vector and for each of the
(2S + 1) z components of a spin angular momentum vector there is an
individual orbital or spin wavefunction. Since a complete wavefunction
contains both orbital and spin parts and any of the orbital functions may
be combined with any of the spin functions, (2L + 1)(2S + 1) distinct
combinations exist. All of these wavefunctions are degenerate. It is the
removal of this degeneracy by a crystal field which is discussed in Chapter
7, by a magnetic field (amongst other things) in Chapter 9 and by spin-orbit
coupling in Chapter II. An individual energy level of a free atom or ion could
be characterized by specifying (2L + I) and (2S + I )-they are both integers
-or, alternatively and more simply by specifying L and S. L is always an
integer butS may not be, so a combination of the two is used. Land (2S + 1)
are given. Rather than quoting Las a number it is replaced by a letter: L = 0
is indicated by S; L = I by P; L = 2 by D; L = 3 by F; L = 4 by G;
L = 5 by H; L = 6 by I and so on. A parallel with orbitals should be evident;
for example, there are five d orbitals; similarly, the label D indicates a
(2 x 2 + 1) =five-fold degeneracy. The value of (2S + 1) is given as a
number, written as a pre-superscript to the L symbol. Thus, 1 S, 3 P, 1 D, 4 F,
etc. When one is talking one says singlet, doublet, triplet, quartet, quintet.
So, 2 D is 'doublet dee'. One talks of a 2 D term. When there are, say, two
electrons distributed within a set of three (degenerate) p orbitals, this is
spoken of as a p 2 (pee two) configuration, the fact that all three p orbitals
are involved being implicit. Such a configuration will usually give rise to
several terms. The p 2 configuration gives rise to 1 D + 3 P + 1 S terms.
There is a simple check that can be applied to a list of terms such as this,
which it is claimed, arise from a given configuration. Consider the p 2
configuration. The first of the two electrons can be fed into the set of three
orbitals in any one of six ways (allowing for spin) and the second in any
one of five (it can only go into the same orbital as the first electron with the
opposite spin). Recognizing that we have counted each arrangement twice
because j t is the same as t j, the number of distinct arrangements, and
therefore wavefunctions, is (6 x 5)/2 = 15. Now, the 1 D, 3 P and 1 S terms
correspond to 5, 9 and I wavefunctions respectively, again a total of 15. The
number of wavefunctions is the same whichever way it is calculated, as it
should be. In Table AS.! the process is worked out in some detail for the d 2
configuration, because this is of importance in Chapter 7. The z component
448 I Appendix 5
Ti n ll ii n H Ti n H ii H H
22
2 1 io
2i i6
20 i -1
26 i -i
2 -1 i -2
2 -i i -2
2 -2
2 -2
06
21 0 -1
2i 0 -i
2o 0 -2
26 0 -2
2 -1 6 -1
2 -i 6 -i
2 -2 6 -2
2 -2 6 -:2
1i
10 -1 -i
16 -1-2
1-1 -1 -:2
1 -i -i -2
1 -2 -i -:2
1 -2 -2 -:2
It is to be emphasized that this table is a book-keeping exercise only; it cannot be assumed that the actual functions are as indicated in it (the actual 1S function,
for instance, is a linear combination of all functions listed with resultant spin and orbital angular momenta of 0).
In this appendix are presented three different methods of obtaining the linear
combinations of Table 6.2, repeated as Table A6.1, the ligand u orbital
symmetry-adapted linear combinations for an octahedral complex. There is
much to be learnt from a careful comparison of all three methods.
Method 1
The first method is scarcely a derivation. Knowing the symmetries spanned
by the ligand u combinations (A 19 + E9 + T 1u), one finds an object of each
of the symmetries in turn and asks what ligand combination matches the
nodal pattern of the object? The most obvious reference objects to choose are
the appropriate orbitals of the central metal atom. Labelling the ligand u
orbitals as in Fig. 6.7, repeated as Fig. A6.1, and referring back to Figs. 6.8
(s orbital), 6.9 (p orbitals) and 6.10(a) (d orbitals) it is evident that the
matching combinations are of the form
p,: (u 3 - us)
remembering that the full label of d,, is d 2 z, _•' _,,. These apart from
normalization, and in one case a -I factor, are the expressions found in
Table A6.1.
Method 2
When we have two equivalent atomic orbitals, 1/1 1 and 1/1 2 say, the former
on atom I and the latter on a symmetry-equivalent atom, 2, symmetry
demands that the electron density at a given point in 1/1 1 is equal to the
electron density at the corresponding point in 1/1 2 . That is, 1/!i = 1/1~ at these
450 I Appendix 6
1
~ 2 (u1- Us)
1
~2 (<12- <14)
1
~ 2 (a3- <1s)
points. It follows that t/1 1 = ± t/1 2 . In other words, t/1 1 and t/1 2 may be either
in-phase or out-of-phase with each other. If t/1 1 and t/1 2 combine to form a
molecular orbital, r/J say, then an analogous argument shows that r/J may
either have the form (t/1 1 + t/1 2 ) or (t/1 1 - t/1 2 ), all other combinations leading
to unequal electron densities on atoms 1 and 2. Normalizing, that is,
multiplying by a factor so that the sum of the squares of the coefficients
multiplying t/1 1 and t/1 2 equals unity, but neglecting overlap between t/1 1 and
t/1 2 , gives the orbitals
and
Appendix 6 1 451
B d
1 2 which, if atoms 1 and 2 are part of a larger, polyatomic, system are called
group orbitals.
In what follows it will, for simplicity, be assumed
1. That all orbitals labelled 1/1 are of a type with respect to a suitably placed
4 3 metal ion, although the discussion has to be little modified to cover 1t
and c5 type interactions.
FIJI. A8.2 Numbering system adopted for
ligand u orbitals of a square planar complex. 2. That all such orbitals have the same (positive) phase. That is, that
1/1 2 has the opposite phase to 1/Jt because of the negative sign, not because
it is inherently of opposite phase.
What if we have four identical a orbitals 1/Jt, 1/1 2 , 1/1 3 and 1/1 4 on atoms
arranged at the corners of a square? The a orbitals on the ligands of a square
planar complex are an example of this situation. The simplest way of
discussing this case is as follows. Label the orbitals cyclically as in Fig. A6.2
and consider the pairs 1/Jt and 1/1 3 , and 1/1 2 and 1/1 4 . We can treat each trans
pair as if the other were not present and use the discussion above to obtain
group orbitals for each pair. These are
1
t/>1 = ../2 (1/11 + .Pal
1
t/>2 = ../2 (1/11- .Pal
1
4>a = ../2 (1/12 + 1/14l
1
4>4 = ..}2 (1/12- 1/14)
All that we have to do now is to form suitable combinations of 4>t• </> 2 , </> 3
and </> 4 to obtain the ligand group orbitals that we are seeking. But how do
we know which combinations to take? The answer is that the transformations
of the final group orbitals must be described by the irreducible representations
of the D4 h point group. If the language of the last sentence seems obscure,
reread Appendix A3. This requirement boils down to the fact that any nodal
planes contained in <Pt. </> 2 , </> 3 and </> 4 , must be compatible with each other
for any interaction to occur. For example, neither 4>t nor </> 3 contain any
nodal planes; because they are equivalent to each other we simply, again,
take their sum and difference
1
ll1 = ../ 2 (t/>1 + 4>al = !(1/11 + 1/12 + .Pa + 1/14)
1
82 =- (t/>1- 4>al = !(1/11- 1/12 + .Pa- 1/14)
..}2
452 1 Appendix 6
However, ¢ 2 contains a nodal plane which passes through 1/1 2 and 1/1 4 and
so ¢ 2 cannot combine with the group orbitals involving these atoms, ¢ 3 and
¢ 4 . Similarly, ¢ 4 has a nodal plane which means that it cannot combine with
either ¢ 1 or ¢ 2 . We therefore conclude that the ligand group orbitals of a
square planar complex are
1
111 = 2 (1/tl + l/t2 + l/t3 + l/t4)
1
e2 = 2 (1/tl- l/t2 + l/t3- l/t4)
1
Eu(1): </>2 = .J 2 (1/tl- l/t3)
The reader is urged to check that these orbitals have the symmetries indicated
by using the D4 , character table given in Appendix 3.
Finally, the problem of the ligand group orbitals appropriate to an
octahedral complex ion. It is convenient to regard the octahedron as a
synthesis of a square planar complex, with the combinations given above,
and two additional ligands with orbitals 1/1 5 and 1/1 6 , which give rise to group
orbitals
1
<l>s = .J 2 (1/ts + 1/ts)
1
<l>s =-(1/ts- 1/ts)
.J2
Applying the nodal plane compatibility requirement it is found that only 11 1
and ¢ 5 , neither of which has a nodal plane, combine together. What are the
correct combinations? It is a general rule that, for the type of ligand group
orbital which we have considered, there is always a combination
1
- (ljt1 + l/t2 + · · · + 1/tn)
.Jn
which has the symmetry given by the first row of the appropriate character
table (the totally symmetric combination, a row of Is). In the present case
this means that one combination is
In summary, the ligand <I group orbitals of an octahedral complex ion are
1
.jB (1{11 + o/2 + 1{13 +f.+ fs + 1{16)
1
T1u(1): - (1{11- 1{13)
.j2
1
r1u(2): .j2 (1{12- "'·)
1
r1u(3): - (fs- 1{16)
.j2
(just the labels 1 and 5 have to be interchanged between the two sets).
Method 3
The final method is group-theoretical, a full derivation has been given
elsewhere. 1 That which follows is the shortest known to the author; we use
Fig. A6.1. From the full octahedral group choose a set of operations
consisting of complete classes, one of which is the identity, such that the
total number of operations in the full group (48) is an integer times the
number in the selected set. Further, choose a set such that the number of
operations in the set is an integer times the number of objects to be
considered (here, we have <I 1 --> <I 6 so this latter number is 6). This selection
1 SeeS. F. A. Kettle, J. Chern. Educ. (1966) 43,21 or S. F. A. Kettle, Symmetry and Structure,
50
50
t./8 AjB
5T2g 1
T1g
1A2g
i3A2g
"'i:U- 40
:~~
11 3Eg
1A1g
80r-------------,---------- .
1Eg
1T2g
5Eg
3T1g
•Eg
50 •r2g
!VB 4 A1g, 4 Eg
2A1g
10
w 0.02
10 000 20 000
cm-1
Rg. A7.4 d5 ; y ~ 4.48 and the spectrum of [Mn(H 2 0) 6 j>+ (the ground
state for all the transitions in the spectrum is 6 A.tg).
Appendix 7 1457
80
'Eg
lEg
3Eg
3A2g
1A2g
1T2g
"'i:iJ 40
30,11 1T1g
3T2g 80
3TlJI
12A2g
4 T2g
1AlJI
50
!:./8
"'i:iJ 2r2g
2r1g
1TlJI<-1AlJI
w 50
0 •rlJI
10 000 20 000 30 000
cm-1
50
t:.jB
10 000 20 000
80,-----------------------,
50
li/B
The most general selection rule for electric-dipole transitions, the type of
transition that is relevant to almost all of the bands observed in the electronic
spectra of transition metal complexes, is a group theoretical statement of the
excitation process described in qualitative terms in Section 8.1. It is the
following:
An electronic transition is electric-dipole allowed if the direct product of the
symmetries of the ground and excited electronic terms of a molecule contains
the symmetry species of one or more coordinate axes, both ground and excited
terms having the same spin multiplicity.
This selection rule sounds somewhat formidable but in practice it is not
difficult either to understand or apply. The coordinate axes come in because
a +--dipole has the same symmetry as a coordinate axis. Direct products
arise because we need to consider, simultaneously, the ground and excited
terms together with the mechanism of excitation (electric dipole). A triple
direct product is therefore needed. The real selection rule is that this triple
direct product must contain the totally symmetric irreducible representation-
that for which all of the characters are +I in the character table. For the
case of octahedral complexes this is A 19 • However, the totally symmetric
irreducible representation only arises when (any) irreducible representation
is multiplied by itself. This means that we can split the triple direct product
into two parts and compare them to see if the two parts contain the same
irreducible representation. The most convenient division is to form the direct
product of the ground and excited terms and to compare the result with the
symmetry of the coordinate axes. Let us consider a specific example, the
3T
29 +- T 19 (F) transition in the d
3 2 case. Here the ground term orbital
oh A1 Az E Tl Tz
Al Al A2 E Tl T2
Az A2 Al E T2 Tl
E E E A1 +A2 +E Tl + T2 Tl + T2
Tl Tl T2 Tl + T2 A1 + E + T1 + T2 A2 +E+T1 +T2
T2 T2 Tl Tl + T2 A;,+E+T1+T2 A;,+E+T1+T2
This table is actually the product table for the 0 point group; for the point group Oh the usual suffiX rule applies:
gX g= U XU= g; g XU= U.
This final direct product contains T 1 • three times so, as T 1 • is the symmetry
species of the coordinate axes to which the selection rule directs our attention,
the intervention of the T 1 • vibration not only makes the 3T29 <-- 3T 19 (F)
transition allowed but also splits it into three separate transitions. These will
differ slightly in energy and so contribute to the breadth of the observed
peaks.
The picture is further complicated when it is recognized that there are
two distinct sets of T 1 • metal-ligand vibrations in an octahedral complex
and, further, there is a set of T 2 • vibrations which may also give intensity to
the d-d transitions by way of four subpeaks. We therefore conclude that we
must expect the 3T 29 <-- 3T 19(F) transition to be composed of no less than
ten (3 + 3 + 4) subpeaks, which may be separated by energies of the order
of magnitude of the energy of metal-ligand vibrations (ca. 200 em- 1 or
more). Each of these peaks is broadened by the vibrational modulation of
~ discussed in Section 8.7 and also, probably, by spin-orbit coupling and
by the Jahn-Teller effect. The large half widths observed in d-d spectra are
understandable!
Append·xg
Determination of magnetic
susceptibilities
Suspension -
Filling mark. scribed on
the sample tube
Draught-proof enclosure,
fitted with a themnometer
pocket (not shown)
they show an apparent increase in weight- the only exception being when
the inherent diamagnetism of the sample swamps its paramagnetism.
Experimentally, the sample is either a solution or a finely ground solid
within a silica or Perspex (Lucile) tube (ordinary glass contains paramagnetic
impurities). The sample length is usually from 5 to 10 em and its diameter
2 to 10 mm. It is suspended from the beam of a balance in a suitable
draught-proof enclosure. More complicated arrangements are used to study
magnetic behaviour over a temperature range. The sample tube is filled to
the same height for all measurements and, for solids, every effort must be
made to ensure that it is uniformly filled (it has been shown that loose
packing can lead to the crystals twisting around in the magnetic field so that
the assumption of random orientation is not valid; this can have a significant
effect on the results). A compound with several unpaired electrons per
molecule would be expected to show a greater paramagnetic effect than a
compound with only one, assuming that the same number of molecules of
each is studied. That is, the quantity
weight increase in the field ~w
would be expected to be larger for the former. The quantity on the right-hand
side of this equation is proportional to the molar magnetic susceptibility of
the compound, XM:
y~w x M
XM = - - -
w
The constant of proportionality varies from apparatus to apparatus- the
more powerful the magnet used, the greater .1w. Hence, y is called the tube
calibration constant. One small addition has to be made to the equation for
XM-the sample tube is diamagnetic and will lose weight in the field. The
actual magnitude of this correction is determined by a trial run with the tube
empty, when its weight drops by J. The value of J must be added to the
experimental values of .1w so as to make .1w more positive (or less negative).
464 1 Appendix 9
Note that this equation refers to the whole molecule. To find out the number
of unpaired electrons on the transition metal atom in a complex an allowance
must be made for all of the paired electrons present in the molecule. These
effects are approximately additive and are listed for some of the more
common species found in complexes in Table A9.1. A more complete list is
given in a reference at the end of this appendix. To make the correction
one sums the appropriate values. So, for [Ni(H 2 0) 6 ]2+Soi- · H 2 0, the
corrections are: Ni2+ 12; SOi- 40; 7H 2 0 7(2 x 3 + 5) = 77-a total of 129
( x 10- 6 /g atom). The correction calculated in this way, D, say, is added to
XM and the sum is x;.., the susceptibility corrected for diamagnetism. That is,
XM = ZM +D
The diamagnetic correction can be more elaborate than indicated above, but
in practice-and particularly when a complicated ligand is involved-it is
best to measure the diamagnetic correction, either directly on the ligand
(using Pascals constants for the remainder) or by measuring the susceptibility
of an analogous but diamagnetic complex. The reader is reminded that in
some cases (Con provides the most important example) it is better not to
make a diamagnetic correction unless a similar allowance is being made for
temperature independent paramagnetism (TIP), see the end of Section 9.3.
With a value for x;.., Jl,rr is obtained using a relationship derived in
Chapter 9
Jlert = Jn(n + 2)
Bringing all this together, gives
References
A good review of reasonably recent developments is C. J. O'Connor, Prog.
Inorg. Chern. (1982) 29, 203. There have been several useful articles dealing
with the measurement of magnetic susceptibilities in J. Chern. Educ. (1972)
49, pages 69,117 and 505; (1983) 60, pages 600 and 681.
AppendixlO
Magnetic susceptibility of a
tetragonally distorted t~g ion
The object of this appendix is to expand the discussion in Section 9.9 on the
magnetic susceptibility of a t1 9 ion such as Ti 111 • It has been written in the
expectation that the reader will be more interested in the general arguments
than in the detail. As far as is possible at a simple level, the individual
mathematical steps will be justified, although the reader should allow neither
the mathematics nor the sometimes off-putting equations to inhibit a general
scan of the appendix. The treatment is algebraic, although today a computer
would be used to treat the problem and would do so more accurately than
does the presentation in this appendix. However, the technical language used
in discussions of magnetism is based on the algebraic approach and so an
understanding of this approach is essential.
The example chosen is almost the simplest possible-a single d electron
confined to the t 29 set of orbitals in a tetragonally distorted octahedral
complex. The effect of the octahedral ligand field is to give three degenerate
orbitals, each of which may be occupied by an electron with spin up or spin
down. There are six wavefunctions, all degenerate. A tetragonal distortion
is now applied and, as Table 6.5 indicates, this separates the d orbitals into
the degenerate pair (dzxo d,,) and dx,, taking the axis of distortion as the z
axis. If the tetragonal splitting is t then retention of the centre of gravity
requires that the energy of the degenerate pair be -t/ 3 and that of dx, be
2t/ 3, assuming that the distortion is such that the latter orbital is destabilized.
The octahedral ligand field stabilization is the same for all three, 2/ 511, and
is omitted from this point on. Including spin, we have four wavefunctions
of energy -t/3 and two of energy 2t/3. Each successive step at each point
in the development of the energy-level argument involves including the next
largest interaction and at this point this is the spin- orbit coupling. The form
of the result of including it follows from the rule given in Section 9.5.
Spin- orbit coupling causes the degenerate orbitals d,x and d,, to interact
and thus to lose their degeneracy. Further, each of these orbitals interacts
with dx, · Detailed calculations give the energies below-check that when (
(zeta, the spin-orbit coupling constant) = 0 they give the energies above.
Once spin-orbit coupling has been included it is no longer possible to refer
to orbitals or to use labels such as d,x·
Appendix 10 1 467
(doubly degenerate)
Each of the above levels is doubly degenerate, and each pair is a Kramers
doublet, with a degeneracy that can only be removed by a magnetic field.
The effect of such a magnetic field on this set of energy levels is conveniently
divided into two parts. The first part is the effect on each individual
degenerate pair; the second part is the magnetic-field-induced interactions
between pairs, interactions which mix them. They are called the first- and
second-order Zeeman effects, respectively. The first-order Zeeman effect leads
to the following energy levels, where, for simplicity, it has been assumed that
the field is applied along the z axis. The quantity k is the orbital reduction
factor.
1 [w(k + 2)] 2 2 2
+ (t2 - tl; + ~1,"2)1/2 x 1 + w2 x f3 H
w2 (k+ 1)
£1 (b) = ~[~C + ~~- (t2 - tC + ~C 2 ) 112] + - X PH
1+ w2
w2 (k+1)
£1 (a) = ~[~C + ~~ - (t2 - tC + ~C 2 ) 112 ] - - x PH
1 +w2
1 [w(k + 2)]2 2H2
- (t2 - tC + ~C2)1/2 x 1 + w2 x p
These equations are quite horrific in appearance and we have considered a
particularly simple case! The important thing to remember is that the
second-order Zeeman effect is characterized by terms which depend on the
square of the field strength. The splitting of the six t 2 • functions we have just
discussed are shown schematically in Fig. AIO.l, which has previously
appeared as Fig. 9.2.
The final step is to consider the thermal population of these energy levels.
Clearly, given the fearsome equations we have obtained, it makes sense to
simplify the problem-we have seen what the real-life algebraic problem
looks like, and that is enough. First, we revert to an octahedral ligand field,
i===-'"/ ==:=<<=======
< E3(b)
E3 (a)
Magnetic field
Appendix 10 1 469
E2 (b) = -~~
E2 (a) = -~~
4{J2H2
E1 (b) = -~~ - - -
31;
4{J2H2
E1 (a) = -~[ - - -
. 31;
These expressions were first derived by Kotani (by a less circuitous route!)
and the simple theory which neglects distortion and the orbital reduction
factor is called Kotani theory. Note that levels which are split apart in the
more detailed analysis are degenerate in the Kotani treatment.
We now have to derive an expression for the temperature variation of the
magnetic susceptibility in terms of the thermal population of these levels. To
do this, the energy levels are conveniently rewritten in the form
En = En(O) + En(1)H + En(2)H 2 + · · ·
Table AlO.l gives the above set of energy levels in this format. We will return
to this table when we have the mathematical expressions that require the
data in this form.
An important property of the magnetic moment of a molecule, Jln, is that
it is proportional to the decrease in energy of the molecule with increase in
the applied magnetic field. In atomic units the proportionality constant is
unity:
where the sum is over all molecules. Because the population of the set of
energy levels will be governed by a Boltzmann distribution law, it follows that
-En(O)) ( -En(1)H)
=exp ( ~ xexp ~ ···
and putting this, together with the expression for Jln given earlier, into the
equation for XM gives, approximately,
We now expand the numerator but first note that when the magnetic field
is zero the average magnetic moment per molecule must be zero and so the
numerator must be zero also. It follows that
L -En(1) -E (0)) = 0
x exp ( _n_
n kT
We have, then,
XM =
HL:exp _n_
n kT
Cancelling the Hs and neglecting the term 2HE.(l)jkT, which will be very
Appendix 10 1 4 71
1/JM =
L:exp -"- (0))
n kT
2 3RTxM 3kTXM
/lett= N2f32 = N/32
f32L:exp (-E(O))
2
!lett=
_n_
n kT
These expressions for XM and Jl;rr are known as the van Vleck relationships.
It is a simple matter to apply these equations to our problem; we have
only to sum over the six different E, states using the values for E,(O), E,(l)
and £.(2) given in Table AIO.l. This gives
2
3kT[2(~-
kT
B~)
3(
exp(-.£) + exp(i_) + +8~() exp (i_)]
kT
2(0 -0)
2kT
2(0
3 . 2kT
llett=
P exp( -~) + exp(~) + expC~r) J
2[ 2 2 2
where the factors of 2 arise because E 1 (a) and £ 1 (b), for example, make the
same contribution to the summation. This expression simplifies to
(~-8)exp(- 3 ')+8
ll~tt = k~ ex{d): 2J
kT
[
2kT
2 (3x-8)exp( -3~)+8
/lett= x[exp( -3~)+2]
another equation first derived by Kotani and discussed in Section 9.9. This
is the equation used to give the theoretical plot in Fig. 9.4(a); for Fig. 9.4(b),
the corresponding but complete expression, including k and t, was used.
Appe ixll
High temperature
superconductors
At the present time, several different, but frequently related, classes of high
temperature (high T,) superconductors are known . The first class that will
be considered is based on La 2 Cu0 4 . This compound has a structure which
may be thought of as consisting of planes of octahedra of oxygen atoms, the
octahedra sharing corners, with a copper atom at the centre of each
octahedron, as shown in Fig. All.l. The lanthanum ions are sandwiched
between the oxygen/copper octahedra layers. The system becomes super-
conducting when Sr 2 + ions are substituted for La 3 + To maintain overall
charge neutralityl one Cu 2+ has to become Cu3+ for every La3+ substituted
by Sr 2 +. Superconductivity occurs, it seems, because of the presence of Cu 3 +
ions in the copper- oxygen planes. It also seems that the superconductivity
is localized in these planes.
The second class of superconductors, the so-called 1,2,3 compounds, are
of approximate formulae YBa 2 Cu 3 0 7 , although the oxygen content is again
variable. Neither yttrium (Y3+) nor barium (BaH) have variable valency so
charge neutrality requires that, for the above formula, of the three copper
ions, two are Cu2+ and one Cu 3 + This class of compound is structurally
more complicated than the first although they seem to have in common that
the superconductivity is associated with the coexistence of Cu2+ and Cu3+
in an oxide lattice. In the 1, 2,3 compounds the lattice is three-dimensional.
Further, there are two distinct copper sites. One, in which the copper ions
are square planar, seems an obvious place for the dB ion Cu3+ (the d B ions
1 The present discussion neglects the fact that the substituted compound may lose oxygen,
although the evidence is that this loss may be important in the phenomenon of high temperature
superconductivity. At the present time, the oxygen loss is regarded as a fine-tuning mechanism;
for every oxygen atom lost two Cu 3 + must revert to Cu 1 +, to compensate for the fact that
o' - ~o.
Appendix 11 I 473
Pt 11, Pd" and, to a lesser extent, Ni 11, characteristically form square planar
complexes). In the second site the copper ions are five-coordinate, being
surrounded by oxygens at the corner of a square pyramid. Such an
environment is not uncommon in copper(Il) compounds and so this seems
to be an obvious site for the Cu 2 + ions. The ratio of occurrence of the square
to square pyramidal sites is 1:2. This is also the ratio of Cu3+ to Cu 2 +,
supporting the model presented for the location of the two different copper
ions.
The relative complexity of the 1,2, 3 and other, yet higher temper-
ature, superconductor structures has meant that the majority of theoretical
work has been concerned with superconductors based on La 2 Cu0 4 , the
La 2 _xSrxCu0 4 superconductors, and only these will be considered in this
appendix (the highest temperature superconductor reported at the time of
writing this appendix contains Sr, Ca and Bi as well as Cu and 0-it is
reported to superconduct at -23 'C). However, to highlight the problem of
explaining the superconductivity of the high T, materials we shall give an
outline of model which for the past 30 years has been the accepted model
of superconductivity. This is the Bardeen, Cooper and Schrieffer (BCS) model.
The simplest explanation of the BCS model is an anthropomorphic one.
Suppose that the reader (an electron) enters at one corner of a large, but
extremely crowded, room in which an informal party is in progress. The
guests (atoms) are making small, to some extent random, movements (atomic
vibrations). The reader (electron) is presented with the task of crossing to
the opposite corner of the room (this simulates an electron's contribution to
an electrical current). In crossing the room the reader will find themselves
frequently deflected-this simulates electrical resistance. Now, suppose that
an ultra important person (UIP)-much more important than a YIP-
together with their partner enters the room, in place of the reader. As the
pair cross the room they will experience no resistance, the guests part to
allow them free passage. Their movement models the movement of super-
conducting electrons through a lattice-these electrons also move as pairs,
so-called Cooper pairs. The essential of the BCS model is a coupling between
the movement of the superconducting electrons and the vibrational motion
of the lattice through which they move. Normally, BCS theory is applied to
materials which are superconducting at very low temperatures-at these
temperatures the vibrational motion of the atoms will be very small anyhow.
At higher temperatures, such as those relevant to the present discussion,
the apparently2 random motion of the atoms will be of much greater
amplitude. Does the BCS model still apply? Apparently not, and therein lies
the problem.
It is easy to demonstrate the problem of explaining the superconductivity
of the high-temperature systems. In Fig. A11.2 is shown a copper-oxygen
plane in La 2 Cu0 4 ; there is an additional oxygen above and below each
copper which, for simplicity, are omitted. Each copper(II) ion in Fig.
Al1.2 has a single unpaired electron and these are antiferromagnetically
coupled, just as is described in Section 9.11, giving the arrangement shown
2 In fact, the atomic excursions may be expressed as a linear sum of thermally populated
lattice modes combined with appropriate phase differences. In addition, there will be the zero
point vibrational amplitude of each mode to be considered, but this contribution becomes less
important as the temperature rises.
474 1 Appendix 11
I
0 0 0 0
I2 I2 I2 I
- 0 - Cu • - 0 - Cu • - 0 - Cu •-o- Cu 2• - 0 -
I I I I
0 0 0 0
I I I I
Fig. A11.2 The copper-oxygen plane in - 0 - Cu 2•-o- Cu 2 •-o- Cu 2 •-o- Cu 2•-o-
La2 Cu04 •
l I I I
0 0 0 0
I2 I2 I2 I
- 0 - Cu + - 0 - Cu + - 0 - Cu + - 0 - Cu 2+ - 0 -
I I I I
0 0 0 0
I I I I
schematically in Fig. All.3. When Sr2+ substitutes for La3+ this substitution
occurs between the copper-oxygen layers-and this is rather remote from
where the superconducting action takes place, or, rather, is believed to take
place-remote from the copper-oxygen planes of Fig. A11.2. When Cu2+
becomes Cu3+ an electron is lost; consider just one such substitution, a
substitution which leaves the structure shown schematically in Fig. A11.4.
In all forms of electrical conductivity, superconductivity not excluded,
electrons have to move from site to site; in the present case the obvious
electron hop, such as that indicated by a curved arrow in Fig. A11.4 leads
to the arrangement shown in Fig. A11.5. The key point is that as far as spin
arrangements are concerned, Figs. A11.4 and A11.5 are not identical; parallel
spins occupy adjacent sites in Fig. A11.5 but not in Fig. A11.4. The discussion
of Section 9.11 makes it quite clear that the arrangement of Fig. A11.5 is
energetically unfavourable. If it is correct, then superconductivity requires
energy to make it occur. Of course, this is just the opposite of reality and
so something is wrong with the ideas that led up to Figs. A11.4 and Al1.5.
Appendix 11 1 475
(b)
The model is wrong; the problem is how to correct it. At the present time
the answer is not known, although it is possible to make some relevant
comments.
First, the problem posed by Figs. All.4 and Al1.5 could have been
avoided had the conduction electron hopped across the diagonal of a square
of copper ions in Fig. A11.4 rather than along an edge. This alternative can
be excluded because the theory is unambiguous-there has to be some
overlap between the orbital of the before and after sites. The distance
involved in a diagonal hop is too great, particularly when it is remembered
that the orbitals of a Cu3+ ion will be contracted compared with those of
a Cu2+. Secondly, we have adopted a one-electron model. At several points
in this book it has been recognized that a one-electron model may not
provide an adequate explanation; is this another example? Almost certainly,
the answer is yes, but there is no agreement on the model to be adopted.
Perhaps the simplest two-electron model is illustrated in Fig. All.6, which
parallels Figs. A11.4 and A11.5. Two electrons hop synchronously, along the
476 I Appendix 11
In Section 11.3 the problem of combining spin and orbital angular momenta
to give the correct resultant was encountered. Put another way, how do
terms such as 3 H, as occurs for an f 2 ion such as Pr 111 and 4 I, appropriate
to an f3 ion such as Nd 111 , split when the spin and orbital motions couple
together? The solution of the general problem will be indicated by considering
these two examples in detail. This is done in Tables A12.1 and A12.2. In
each table down the left hand side are given all of the components of the
orbital angular momentum; at the top are listed all of the components of
the spin angular momentum. The entries in the body of the tables are total
angular momenta, obtained by adding the appropriate spin and orbital
momenta (those at the top and side of the table). In Table A12.1 it is not
difficult to see that all of the components of the state with total angular
e
momentum of 6 H 6 ) are covered by the left-hand column and final row
entries.
If these entries are deleted from the table, the components of the state
with total angular momentum 5 eH5 ) are similarly obtained from the table
remaining. Repeating this process gives the components of the total angular
e
momentum 4 H4 ) state. That this pattern is not accidental is evident from
Table A12.2, where it is repeated. Note particularly that each set of total
angular momenta (the entries within the shaded panels in Tables A12.1 and
A12.2) contain an entry from the figures in the first row of the table. Further,
that these entries, together, are a complete listing of the magnitudes of the
permissible angular momenta. It follows that, as described in Section 11.4,
the allowed values of the total angular momentum are obtained by taking
the maximum value of the orbital angular momentum and adding each
component of the spin angular momentum to it.
One final word of caution, one that also appears in Appendix A14 and
which explains the appearance in the tables of the word ' schematic'. What
has been described in this appendix is really a method of counting; it is not
a method of obtaining wavefunctions. So, it is not true that the component
of the total angular momentum = 6 state which has an angular momentum
of 0 along the z axis is the function with orbital angular momentum = -I
combined with spin= I, although this is the association made in the table.
The correct wavefunction is a linear combination of all three functions with
478 1 Appendix 12
1 0 -1
5 6 5 4
4 5 4 3
3 4 3 2
2 3 2 1
Orbital 1 2 1 0 . _ 3H4
angular 0 1 0 -1
momentum - 1 0 -1 -2
-2 -1 -2 -3
-3 -1 -3 -4
- 4 -3 -4 - 5 +-- 3Hs
-5 -4 -5 -6 ,._____ JHe
3/ 2 1/2 -1/ 2 - 3/ 2
6 15/2 13/2 11/2 9/2
5 13/2 11/2 9/2 7/2
4 11/2 9/2 1/2 5/2
3 9/2 7/2 5/2 3/2
2 7/2 5/2 3/2 1/2
1 5/2 3/2 1/2 - 1/2 +--- 4/9!2
Orbital 0 3/2 1/2 - 1/2 - 3/2
angular - 1 1/2 - 1/2 - 3/2 - 5/2
momentum - 2 - 1/2 - 3/2 - 5/2 - 712
-3 - 3/2 - 5/2 - 7/2 - 9/2
-4 - 5/1 - 7/2 - 9/2 - 11/2 +--4111/2
-5 - 7/2 - 9/2 - 11/2 - 13/2 +--4/up
-6 -9/2 -11/2 -13/2 - 15/2 .._4,15/2
Problem A12.1
Construct a table similar to Tables A12.1 and A12.2 for the 5 I (f4 ) case and
show that the levels 5 I 4 , 5 I 5 , 5 I 6 , 5 I 7 and 5 I 8 are given by the procedure
described in this appendix (see also Section 11.4).
Bonding between a transition
metal atom and a Cn Rn ring,
n = 4, 5 and 6
In this appendix are merged two closely related approaches to the bonding
between a cyclic C,R, (n = 4, 5 and 6) unsaturated hydrocarbon and a
transition metal atom. The first approach is that developed in Chapter 10.
The axis between the centre of the C, R, ring and the metal atom is taken
as z and the orbitals of the metal atom classified according to the number
of planar nodes which they have and which contain the z axis (cutting the z
axis does not count). The 1t orbitals of the C,R, ring are classified using the
same criterion. Interaction- be it bonding or antibonding-only occurs
between orbital pairs with the same nodal count. All interactions between
orbitals with different nodal counts are zero. This approach is summarized
in Figure A13.1 . In this figure the metal orbitals are drawn in a direction
perpendicular to the z axis (thus enabling their identity to be established, s,
p, and d,, look much the same when viewed down z). However, the
hydrocarbon n molecular orbitals are viewed down z, because this enables
their nodality to be clearly seen. As is the common pattern-from orbitals
through to acoustics- the more nodes, the higher the energy. Of course, for
this simple statement to be true the vibrations under discussion must be
immediately related to each other. So, the statement is applicable to the n
orbitals of hydrocarbons, because different combinations of carbon p,
orbitals are involved. It does not apply to the metal orbitals because these,
s, p and d, have no such simple common basis. For unsaturated hydrocarbons
it means that the orbitals of highest nodality- of highest energy- are empty
in the isolated molecule. These higher orbitals, then, are regarded as 1t
electron acceptors in a transition metal complex and the (occupied) orbitals
of lower nodality are 1t electron donors. Schematic molecular orbital
diagrams based on these arguments are given in Figure A13.2. We do not
indicate the final occupancy of the molecular orbitals because this will
depend on which particular metal atom is involved.
The second, equivalent, approach is group theoretical (see Appendix 4).
The local symmetry of the transition metal- C,R, unit in a complex is C,v·
The transformation under the operation of the C,v character table of the n
carbon p, orbitals in the ring gives the irreducible representations subtended
480 I Appendix 13
z z
¢~ ~ Pz
0 nodal planes
z z
N~~m
0-0 6=0 0--Kb
_cp±y2
6-t-0
0~~~. -
- +
+ +
- -
+
~~
FJg. A13.1 (b) (continued).
~1Wr
'--r----'
2
0-e=
~
. . : I .
(b)
482 1 Appendix 13
(n+ l)p
(n+ l)s
(0)
(a)
Fig. A:l3.2 Schematic molecular orbital energy The connection between the two approaches described above is also
level diagrams for transition metal complexes of shown in Table Al3.1. At the right-hand side of each character table is listed
(a) C4 R4 , (b) C5 R5 and (c) C6 R6 • The ligand
and final molecular orbitals are classified by the number of nodal planes containing the z axis associated with each type
their symmetry species under the relevant point of interaction. There is something to be said in favour of each model in
group. The final molecular orbitals are also comparison with the other. Thus, the 'number of nodal planes' criterion does
classified by the number of nodal planes they
have which contain the z axis. In all cases, for not discriminate between B 1 and B2 (in the C4 • and C6 • cases), yet only one
simplicity, only interactions between d orbitals is involved in the bonding. Conversely, although in each case the number
and the ligand orbitals are shown. However, as of nodal planes quoted is the lowest, for each case higher numbers exist also
Table A13.1 shows, p orbitals must be expected
to be involved in the (1)-node interactions and (we give these for the A 1 case). In such cases, the number of nodal planes
both p and s orbitals in the (0)-node interactions. criterion provides a distinction that the symmetry criterion does not. Only
orbitals with the same number of nodal planes interact.
Appendix 13 1 483
TableA13.1
(a) C4 R4 : P. orbitals of C4 R4 transform as A1 + E + 81 (taking the C4 atoms to lie in the 2crv
mirror planes)
C4v E 2C4 c. 2u. 2a~ Metal orbitals C4R4 T<? Nodal planes
c•• E 2C6 2C3 C2 3av 3a~ Metal orbitals CsRsT<? Nodal planes
A1 1 1 1 1 1 1 S, Pz• dz2 yes 0 or 6
A2 1 1 1 1 -1 -1 no 12
81 1 -1 1 -1 1 -1 yes 3
82 1 -1 1 -1 -1 1 no 3
E1 2 1 -1 -2 0 0 Px• Py; dzx, dyz yes 1
E• 2 -1 -1 2 0 0 dx2-y2• dxy yes 2
Hole-electron relationship in
spin-orbit coupling
!
0
0
0
0
0
0
3 2 1 0 -1 -2 - 3
5 +--- 3Fo
3 6 4 3 2 1 0
2 5 4 3 2 1 0 - 1 ._3Ft
Orbital 1 l!\ 3 2 1 0 -1 -2 +--- 3F,
-3 +-- 3F4
momentum -1 2 1 0 - 1 -2 -4
- 2 1 0 -1 -2 -3 -4 - 5 + - 3Fs
0 -1 -2 -3 -4 -5 -6 ._JF6
-3
Appendix 14 1 485
The entries in this table can be classified according to the levels that result.
So, the 7 F6 level has components with resultant angular momentum of 6, 5,
4, 3, 2, 1, 0, -1, -2, -3, -4, -5 and -6. Deleting entries with these values
from the table leaves 5 as the highest value. This must come from a 7 F5 level,
with components 5, 4, 3, 2, 1, 0, -1, -2, -3, -4 and -5. Proceeding in
this manner it is concluded that the 7 F6 , 7 F5 , 7 F4 , 7 F3 , 7 F2 , 7 F, and 7 F0 levels
of the 7 F term are obtained. Had the 7 F arising from the f 6 configuration
been under discussion, we would have been working with electrons and the
7 F0 level would have been the most stable-the orbital and spin magnets
Creutz-Taube species 323, 341 electroneutrality 116 group theory 95, 122, 440
cross-reactions 322 electronic spectra 156 Guoy method 462
crown ether 3, 14 electronic spectra off electron systems 257
cryptosolvolysis 333 electronic spectra of V111 and Ni 111 157
cryptands 14, 91 enantiomers 46, 433 H
encounter complex 331 hapticity 28
crystal field splittings 123
crystal field stabilization energies 306, endohedral species 70, 215 hard acids 82
309, 337 energy level diagram for f 2 256 hard and soft acids and bases (HSAB) 81
hard bases 82
crystal field stabilization energies (of enthalpy 87
lanthanides) 312 entropy 87 heats of ligation 305
entropy of activation 327 hemoglobin 305, 384
crystal field theory 121
ethane 17 heterogeneous catalysis 21
crystal molecular orbitals 410
ethenetetracarbonyliron 232 hexacarbonylchromium 229, 293
cubic setoff orbitals 241, 242
exchange coupling and spectra 182 hexadentate 13
Curie-Weiss law 203, 205
excited states off electron systems 254 high spin 128, 130
cyclic voltammetry (CV) 297
exohedral 216 high temperature superconductors 208, 472
cylindrical symmetry 348
extended Hucke! method 223,232,345,376 highest occupied molecular orbital
extended X-ray absorption fine structure (HOMO) 3, 83, 212, 423
D
(EXAFS) 394, 400, 402 hole-electron relationship (in spin-orbit
d orbitals in an octahedral crystal field 124
coupling) 484
D (dissociative mechanism) 326
F
holes 133
D3, character table 120
f electron charge-transfer transitions 262 Hund's rules 138, 142
D4 direct product table 443
f electron systems 238 hydrate isomerism 45
D4 , character table 440
f orbital covalency 219 hyperfine structure (in EPR) 289
de Haas-van Alphen effect 424
hypersensitive transitions 261
density functional method 222, 267, 376 f orbital involvement in bonding 267
f orbital shapes 240 hypha (cages) 355
density of states (DOS) 376, 418
determination of magnetic susceptibilities f orbitals in an octahedral crystal field 126
462 f--> d transitions 257, 262
diamagnetism 186 f--> f transitions 257, 260 1 (interchange mechanism) 326
didentate 10 face bonds 349 I, (associative interchange mechanism) 327
differential thermal analysis (DTA) 62 face centred cubic 408 '• (dissociative interchange mechanism)
direct product 138, 442 Faraday method 462 327
dispersion 419, 429, 430 Fenske-Hall method 223 icosahedral 216
dissociation reactions 51 Fermi surface 420, 423 incongruent solubility 55
distorted octahedral crystal fields 153 ferredoxins 401 inelastic neutron scattering 270
distortional isomerism 43 ferrimagnetism 203 inert complexes 9, 52, 73, 318
donor-acceptor 1 ferrocene 212, 227, 292, 294 infrared spectroscopy 270
double groups 249 ferromagnetic linking 206 inner-sphere mechanism 323
droplet model 376 ferromagnetism 203 insertion reactions 66
dynamic exchange in NMR 319 first-order Zeeman effect 196, 467 intermediate field complexes 143
dynamic Jahn-Teller effect 170, 200 flash photolysis 392 intermediate neglect of differential overlap
flexidentate 26 (INDO) 223
E fluorescence 342 intraligand charge-transfer spectra 178,
edge bonds 34 7 fluxional isomerism 4 7 181
Eigen-Wilkins mechanism 331 fluxional molecules 21, 329, 338 ion replacements 239, 384
electric dipole allowedness 156, 270, 459 flyover complex 19 ionic potential 81
electric field gradient 287 formal oxidation states 115 ionic radius 81, 91, 239, 303
electric quadrupole moment 285 free-electron models 362 ionization isomerism 45
electrochemical methods 296 free energy 87, 311 irreducible representation 443
electron correlation 107, 118,220,228, 377 fullerenes 215, 362 Irving-Williams order 81, 311
electron deficiency 349 isoleptic 27
electron density difference maps 117 G isolobal principle 234, 345, 357
electron paramagnetic resonance (EPR) g tensor 289 isomer shifts 287
288, 397, 400, 402 general set off orbitals 241, 244
electron repulsion 107, 118, 128, 130, 138, geometrical isomerism 44, 209
143, 220, 223, 236, 245 gerade 96, 108, 131
electron spin resonance (ESR) 288 graph theory 361 Jahn-Teller effect 164, 166,293,311
electron-transfer reactions 317, 320 group (symmetry) 440 J ahn-Teller splitting 190
electron transfer 1 group constants (in protonation jellium model 376
electronegativity 82 equilibria) 404 Job's method 78
Index 1 489