The F Elements: Nikolas Kaltsoyannis and Peter Scott
The F Elements: Nikolas Kaltsoyannis and Peter Scott
S. E. Thomas Organic synthesis: The roles ofboron P. D. Bailey and K. M. Morgan Organonitrogen
and silicon chemistry
D. T. Davies Aromatic heterocyclic chemistry C. E. Wayne and R. P. Wayne Photochemistry
P.R. Jenkins Organometallic reagents in synthesis C. P. Lawrence, A. Rodger, and R. G. Compton
M. Sainsbury Aromatic chemistry Foundations of physical chemistry
L. M. Harwood Polar rearrangements R. G. Compton and G. H. W. Sanders
J. H. Jones Amino acid and peptide synthesis Electrode potentials
C. J. Moody and G. H. Whitham Reactive P. B. Whalley Two-phase flow and heat transfer
intermediates L. M. Harwood and T. D. W. Claridge
G. M. Hornby and J. M. Peach Foundations of Introduction to organic spectroscopy
organic chemistry C. E. Housecroft Metal-metal bonded carbonyl!
R. Henderson The mechanisms of reactions at tran- dimers and clusters
sition metal sites H. Maskill Mechanisms of organic reactions
H. M. Cartwright Applications of artificial P.C. Wilkins and R. G. Wilkins /norganic
intelligence in chemistry chemistry in biology
M. Bochmann Organometallics 1; Complexes with J. H. Jones Core carbonyl chemistry
transition metal—carbon o-bonds N. J. B. Green Quantum mechanics 1; Foundations
M. Bochmann Organometallics 2: Complexes with I. S. Metcalfe Chemical reaction engineering:
transition metal—carbon m-bonds A first course
C. E. Housecroft Cluster molecules of the p-block R.H.S. Winterton Heat transfer
elements N.C. Norman Periodicity and the s- and p-block
M. J. Winter Chemical bonding elements
R.S. Ward Bifunctional compounds R. W. Cattrall Chemical sensors
S. K. Scott Oscillations, waves, and chaos in M. Bowker The basis and applications of
chemical kinetics heterogeneous catalysis
T. P. Softley Atomic spectra M. C. Grossel Alicyclic chemistry
J. Mann Chemical aspects of biosynthesis J.M. Brown Molecular spectroscopy
B. G. Cox Modern liquid phase kinetics G. J. Price Thermodynamics of chemical processes
A. Harrison Fractals in chemistry A. G. Howard Aquatic environmental chemistry
M. T. Weller /norganic materials chemistry A.O.S. Maczek Statistical thermodynamics
R. P. Wayne Chemical instrumentation G. A. Attard and C. J. Barnes Surfaces
D. E. Fenton Biocoordination chemistry W. Clegg Crystal structure determination
W.G. Richards and P. R. Scott Energy levels in M. Brouard Reaction dynamics
atoms and molecules _K. Brisdon /norganic spectroscopic methods
M. J. Winter d-Block chemistry . Proctor Stereoselectivity in organic synthesis
D. M. P. Mingos Essentials of inorganic chemistry 1 .M.A. Brett and A. M. O. Brett Electroanalysis
G. H. Grant and W. G. Richards Computational Ze
Gy _ J.B. Green Quantum mechanics 2: The tool kit
chemistry D. M. P. Mingos Essentials of inorganic chemistry 2
S.A. Lee and G. E. Robinson Process development: I. Fleming Pericyclic reactions
Fine chemicals from grams to kilograms
N.S. Lawrence, J. D. Wadhawan & R. G. Compton
C. L. Willis and M. R. Wills Organic synthesis Foundations of physical chemistry. Worked examples
P. J. Hore Nuclear magnetic resonance J. R. Chipperfield Non-aqueous solvents
G. H. Whitham Organosulfur chemistry T. J. Mason Sonochemistry
A.C. Fisher Electrode dynamics
J. McCleverty Chemistry of the first-row transition
G. D. Meakins Functional groups: Characteristics metals
and interconversions
E. C. Constable Coordination chemistry of
A.J. Kirby Stereoelectronic effects macrocyclic compounds
P. A. Cox Introduction to quantum theory and C. E. Housecroft The heavier d-hlock metals:
atomic structure Aspects of inorganic and coordination chemistry
—_—_e——_—__C_C
HA 0196 504675 9001
la Loe =a
STILLORGAN LIBRARY
Inv/06 : K109 Price E16.05
Title: os elements
e | Class: Ub: ae
witr i as ee
Nikolas Kaltsoyannis
Lecturer in Chemistry, University College London
Peter Scott
Senior Lecturer in Chemistry, University of Warwick
WITHDRAWN FROM
DUN LAOGHAIRE-RATHDOWN COUNTY
LIBRARY STOCK
OXFORD
UNIVERSITY PRESS
OXFORD
UNIVERSITY PRESS
Preface
There are two rows of elements propping up the periodic table about which most graduate
chemists know very little. Some may have come across the lanthanide contraction and may
be aware of its consequences for the chemistry of the d block. Others may have met the sand-
wich molecule uranocene in courses on organometallic chemistry. It is also possible that the
unique magnetic and spectroscopic properties of these elements were encountered at the end
of a physical chemistry lecture series. This piecemeal treatment is unfortunate as the lantha-
nides are finding increasing technological and catalytic applications, and the concern over
the growing quantities of actinide waste littering the globe is well founded. Our graduates
ought to know more about the f elements.
In this text we have drawn together key concepts in f element chemistry to provide a
student-friendly introduction to this fascinating group of elements. The traditional approach
adopted in the few existing texts on the f elements is to treat each element in turn. The very
nature of the lanthanides and actinides as families of elements with similar properties means
that this involves considerable repetition. We have chosen instead to concentrate on the
trends within the two series and the comparisons between them and the other elements of the
periodic table. We hope that we have provided a collection of concepts to be applied rather
than facts to be memorised. The material is targetted at third or fourth year undergraduates,
and will put to use many ofthe skills developed in earlier years and indeed in other Primers.
We would like to thank Andrea Sella for casting a critical eye over the entire manuscript,
and Geoff Cloke, Bob Denning, Peter Moore, David Parker, and John Watkin for helpful
comments on various sections. We also want to thank J and N for their uniquely special
contributions.
NK PS
London Warwick
January 1999
Contents
Definitions
and origins
Coordination chemistry
nn
Aa
NHN
B&B
&= Organometallics
Further reading
Index
1 Definitions and origins
Fig. 1.1 The modem periodic table. The f elements are situated in the two shaded rows, and are divided into the lanthanides (light
shading) and the actinides (dark shading). The ‘parent f elements, lanthanum and actinium, are also shaded appropriately.
Although the ‘parent’ elements of the f block, lanthanum and actinium, are
really members of group three, they are often included in discussions of f
element chemistry. We will not, therefore, confine ourselves to the shaded
rows in Fig. 1.1, but will include certain aspects of the chemistry of
lanthanum and actinium where appropriate. Comparisons will also be made
between the lanthanides and scandium and yttrium.
Throughout this book we will use the the general symbols Ln and An to
refer to the lanthanide and actinide elements respectively.
been identified bar the radioactive promethium, conclusive evidence for which
had to wait until 1947, when J.A. Marinsky, L.E. Glendenin, and D.C.
Coryell observed element 61 in the radioactive decay products of ?°U.
Shortly after Moseley’s work, N. Bohr realised that the 14 elements
between lanthanum and hafnium reflected the fact that the fourth atomic
primary quantum shell could accomodate 32 electrons, 14 more than the third.
These additional electrons are placed in the 4f orbitals, and the lanthanide
elements were recognised as forming a new family in the periodic table. For
obvious reasons the lanthanides are often referred to as the 4f series of
elements. That the actinides form a second f series - the Sf - was not realised
until the work of G.T. Seaborg during the Second World War. His suggestion
took thorium, protactinium, and uranium out of groups four, five, and six
respectively and into their rightful place at the start of a new family of
elements, many of which Seaborg was instrumental in synthesising.
Lanthanides
The term ‘rare earth’, which is often used to describe the lanthanides, is rather
misleading as many of the 4f elements are quite abundant. Cerium is the 26"
most abundant element on the Earth, neodymium is more abundant than gold
and even thulium (the least common lanthanide except for the radioactive
promethium) is more abundant in the Earth’s crust than iodine. Only two
lanthanide-containing minerals are important commercially; monazite, a
mixed lanthanide orthophosphate (LnPO,) and bastnaesite, a fluorocarbonate
(LnCO,F). The most common metals in both ores are (in order of decreasing
abundance) cerium, lanthanum, neodymium, and praseodymium, with
monazite also containing up to 10% ThO, as well as smaller quantities of the
later lanthanides. Monazite deposits occur in many countries, including India,
Brazil, Sri Lanka, South Africa, Australia, and Malaysia, while the principal
sources of bastnaesite are China and the Sierra Nevada mountains in the
western USA.
Monazite is typically processed by dissolution in ca. 70% NaOH solution
for several hours followed by addition of hot water to generate a slurry of
crude hydrous oxides. The slurry is then added to boiling HCl until a pH of
3.5 is reached, at which point crude hydrous ThO, is precipitated.
Stoichiometric amounts of BaCl, and Ln,SO, (three and one equivalents
respectively) are then added to the remaining solution of impure LnCl, to
precipitate BaSO, and RaSO,, leaving a solution of mixed lanthanide
trichlorides. This is also the end product of bastnaesite processing, although a
somewhat different route is employed. Large-scale separation of the individual
lanthanide chlorides is achieved using solvent extraction, typically with a
complexing agent such as tributylphosphate - ("BuO),PO - in an inert diluent
such as kerosene. Separation relies on the increased solubility of Ln(II) with
increasing atomic mass. Alternatively, high purity, small-scale separation
4 Definitions and origins
Lanthanides Actinides
Element Symbol Atomic Electronic Element Symbol Atomic — Electronic
number configuration number configuration
Cerium Ce 58 [Xe]4f!5d!6s? Thorium Th 90 [Rn]6d?7s?
Praseodymium Pr 59 [Xe]4f%6s" Protactinium Pa 91 [Rn]5f°6d'7s"
Neodymium Nd 60 [Xe]4f*6s" Uranium U 92 [Rn]5f°6d'7s?
Promethium Pm 61 [Xe]4f6s" Neptunium Np 93 [Rn]5f*6d'7s"
Samarium Sm 62 [Xe]4f%6s" Plutonium Pu 94 [Rn]5f°7s"
Europium Eu 63 [Xe]4f’6s" Americium Am 95 [Rn]5f’7s*
Gadolinium Gd 64 [Xe]4f’5d'6s" Curium Cm 96 [Rn]5f’6d'7s"
Terbium Tb 65 [Xe]4f6s" Berkelium Bk 97 AIST
Dysprosium Dy 66 [Xe]4f'°6s’ Californium Cf 98 {Rn]5f'°7
Holmium Ho 67 [Xe]4f!'6s? Einsteinium Es 99 [Rn]5sf'7¢
Erbium Er 68 [Xe]4f!*6s" Fermium Fm 100 [Rn]5f'?7s?
Thulium Tm 69 [Xe]4f'6s? Mendelevium Md 101 (Rn]5f'*7s"
Ytterbium Yb 70 [Xe]4f!*6s? Nobelium No 102 (Rn]5f'*7s?
Lutetium Lu as Xe]4fl5d'6s*__ | Lawrencium Ly 103 [Rn]5f'*6d'7s"
f orbitals
The 4f and 5forbitals play a central role in determining the physicochemical
properties of the lanthanides and actinides respectively. Unfortunately there is
no unique way of representing f orbitals. Figure 2.1 presents one of the most
common sets of f orbitals - the cubic set - which is appropriate for molecules
in which the x, y, and z axes are symmetry related, e.g. those belonging to
the O, or 7, point groups.
Thef elements 7
£,3 (= fy222-3:2-a2))- £,3 (= faex2-ay2-a22)) ANA £3 (= fyiay2-a22.a,2)) are similar and are oriented
along the x and y axes respectively.
xyz
8 Properties of the atoms and ions
Lanthanides
The chemistry of the lanthanides is dominated by the +3 oxidation state. One
of the principal reasons for this is illustrated by the ionization energy data in
Table 2.2, which reveal that in all cases the fourth ionization energy, /, [the
energy associated with the process Ln** (g) — Ln* (g) + e (g)] is greater than
the sum of the first three ionization energies. The extra energy required to
remove the fourth electron is so great that in most cases it cannot be
recovered through chemical bond formation, and thus the +4 oxidation state is
largely inaccessible.
Element 1, iD ic Label, iE
Ce 527 1047 1949 3523 3547
Pr 523 1018 2086 3627 3761
Nd 530 1035 2130 3695 3899
Pm 524 1052 2150 3726 3970
Sm 543 1068 2260 3871 3990
Eu 547 1085 2404 4036 4110
Gd 593 1167 1990 3750 4250
Tb 565 1112 2114 3791 3839
Dy 572 1126 2200 3898 4501
Ho 581 1139 2204 3924 4150
Er 589 1151 2194 3934 4115
Tm 597 1163 2285 4045 4119
Yb 603 1176 2415 4194 4220
Lu 524 1340 2022 3886 4360
Table 2.2 lonization energies of the lanthanides (kJ mot’).
The f elements 9
/kJ
|,
mol"
8
Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er
Fig. 2.2 The variation of the third ionization energy (/,) of the lanthanides.
The variation in the third ionization energy, /;, of the lanthanides is shown
in Fig. 2.2. The most striking feature of this plot is the very high values for
europium and ytterbium and the very low values for the elements immediately
following them, gadolinium and lutetium. This may be explained by
consideration of the electronic configuration of the Ln’* ions that are being
ionized to form the corresponding Ln**. Yb** has the [Xe]4f'* configuration.
Lu?*, however, has an additional electron in the 5d orbitals, which are less
stable than the 4f and therefore easier to ionize. The situation is less clear for
Gd**, because while there is little doubt that the electronic configuration of
Eu** is [Xe]4f’, it is not certain if that of Gd** is [Xe]4f’5d' or [Xe]4f*. If it is
the former, the low ionization energy may be rationalised in the same way as
10 Properties of the atoms and ions
for Lu’*. The [Xe]4f* configuration, however, may also be expected to ionize
more easily than the [Xe]4f’. In the latter, all seven f electrons occupy
different f orbitals with the same spin. In [Xe]4f*, however, one f orbital must
contain two paired electrons, and the increased repulsion between this pair
will destabilize the ion and make it easier to ionize. Furthermore, there is no
loss of exchange energy on ionization of the [Xe]4f* configuration, as the
electron that is removed has opposite spin to the other seven f electrons. By
contrast, there is a significant exchange energy loss on ionization of the
[Xe]4f’ configuration, and hence more energy is required to remove an electron
from it.
Given the data in Table 2.2 and Fig. 2.2 it should come as no surprise that
the lanthanides with the most extensive divalent chemistry are europium,
ytterbium, and to a lesser extent, samarium and thulium (i.e. those with large
I,). We shall come across many examples of divalent lanthanide compounds
throughout this book.
Actinides
The actinide elements display a much greater range of oxidation states than
the lanthanides, particularly in the early part of the series. Figure 2.3 shows
the oxidation states adopted by the actinides, from which it may be seen that
some of the lighter actinides resemble the transition metals in their range of
possible oxidation states, while the later actinides are more like the
lanthanides in favouring trivalency. This change of character as the series is
crossed is a general theme in actinide chemistry, one which we will return to
many times in this book.
+7 Oo. ©
+6 e oo oOo oO
2
S45 e o eo oO
6
S 44 0 OF —-O* Oh * « Os — -6— O
6
E +3 Oo oF OF 08F 0O06oe e060 0 06e 6 6 oO @®
LL
Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
Fig. 2.3 The oxidation states adopted by the actinide elements in their compounds. The most stable oxidation state in aqueous solution
is represented by the black circles. Open circles indicate other oxidation states adopted and squares indicate that the oxidation state is
found only in solids.
The f elements 11
Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
xs ‘ > z
A
a 20 8 +—e_ ee «©
160 Metallic Ln
(pm)
Radius
Ce Pr Nd Pm sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Fig. 2.5 Variation of metal radius and +3 ionic radius for the lanthanide elements.
The reduction in the lanthanide metal and Ln** radii with increasing atomic
number is often referred to as the lanthanide contraction, and the
corresponding effect in the 5f series as the actinide contraction. These
contiactions arise from the poor ability of f electrons to screen the other
valence electrons from the nuclear charge. Although the additional unit of
nuclear charge on moving one element to the right is exactly balanced by the
opposite charge of an extra f electron (such that the total charge on the atom
or ion does not change), the poor screening of the nucleus by the additional f
electron means that the effective nuclear charge experienced by all of the
valence electrons increases slightly, and the atom/ion contracts.
The poor screening ability of f electrons is primarily a consequence of their
high angular nodality. However, relativistic effects (see Chapter Three,
Section Two) also play a part. Calculations reveal that 4f and 5f electrons are
expanded and destabilised with respect to ficticious atoms in which the effects
of relativity are not included, and are even poorer at screening the nuclear
charge than would be anticipated purely on the basis of their angular nodality.
The f elements 13
190
a ye ee
he ~—— Metallic An
(pm)
Radius
Th Pa U Np Pu Am Cm Bk Cf Es
Fig. 2.6 Variation of metal radius and +3, +4, and +5 ionic radius for the actinide elements.
The effects of the actinide contraction on the size and chemistry of the 6d
series 1s difficult to ascertain as the transactinide elements may only be made
in very small quantities and decay radioactively with very short half-lives.
Nevertheless, calculations suggest that the actinide contraction may have an
even greater effect than the 4f equivalent. For example, element 111 (which
lies under gold) is calculated to have a radius even smaller than copper!
Calculations also reveal that the role of relativity in the actinide contraction is
appreciably greater than in the lanthanide contraction.
All of the lanthanide metals except europium and ytterbium may be
considered to consist of Ln** ions with three electrons per atom devoted to
metallic bonding. Europium and ytterbium, however, are best regarded as Ln’*
(which are larger than Ln**) with only two electrons per atom involved in
metallic bonding. This accounts for the significantly greater metallic radii of
europium and ytterbium with respect to the other lanthanides and is discussed
in more detail in Chapter Four, Section Three. A similar effect 1s believed to
be the cause of the trend in the metallic radius of the actinide elements (Fig.
2.6). As shown in Fig. 2.3, the most stable actinide oxidation state increases
14 Properties of the atoms and ions
Nuclear fission
Both the military and peaceful uses of actinide-based nuclear energy have as
their basis the same physical process, nuclear fission. Nuclear fission occurs
when a large nucleus splits into two smaller ones, a process which also
releases one or more neutrons which may collide with further nuclei causing
them to split and generate yet more neutrons. This is shown schematically for
the fission of a **°U nucleus in Fig. 2.7.
wy Ya
=e = — =e
O*
t Ry \\
Fig. 2.7 Schematic representation of the fission of a *°U nucleus, showing one of its many fission patterns. This process releases
2.1x 10" kJ mol’. Over 200 different isotopes of 35 different elements have been found in the fission products of 25U.
The f elements 15
Table 2.3 Half-lifes and radioactive decay mechanisms of the most stable isotopes of the
actinides.
In 1939 Albert Einstein wrote to This was the principle behind the first nuclear weapons developed at Los
President Roosevelt of the USA Alamos in the USA and used to such devastating effect at Hiroshima and
emphasising the potential military
Nagasaki in Japan at the end of the Second World War. For example, the
applications of nuclear fission and
raising the possibility that the Nazis ‘Little Boy’ bomb that was dropped on Hiroshima consisted of two subcritical
could develop an atomic bomb. masses of *°U (i.e. two amounts each less than the critical mass) which were
Roosevelt decided that it was vital that slammed together using conventional chemical explosives to create a single
the USA investigate the viability of supercritical mass. The uncontrolled chain reaction rapidly led to a nuclear
such weapons, and in late 1941 the
explosion equivalent to the detonation of 20 000 tons of TNT.
go-ahead was given for the ‘Manhattan
Project’, the goal of which was to build
a bomb based on the fission process. Nuclear power generation
The fissile properties of *°U are put to altogether more peaceful use as the
power source in nuclear power stations. **U is, for all practical purposes, the
only naturally occuring fissile nucieus. However, while natural abundance
uranium is capable of sustaining a fission chain reaction, the low
concentration of **°U (0.72%) is such that by the time the effects of the fuel
cladding and the other reactor materials are taken into account, it is
advantageous to increase the proportion of **°U in the nuclear fuel. Hence
nuclear fuel is typically composed of UO, enriched to 2-3% **U.
Figure 2.8 is a schematic diagram of a typical nuclear reactor. Aside from
the fuel rods (zirconium or stainless steel tubes containing UO, pellets), the
other principal features are the circulating cooling fluid, the control rods, and
the moderator. The kinetic energy of the fission products is dissipated by
collisions with surrounding atoms, releasing huge amounts of heat (ca. 10°
times that produced by burning an equivalent mass of coal). In a reactor, this
Control
rods
heat is absorbed by the cooling fluid (usually water or heavy water, D,O)
which is subsequently used to drive steam turbines, thereby generating
electricity. The control rods are good neutron absorbers - usually boron steel
or boron nitride - and are used to regulate the flux of neutrons and thus
Fuel prevent the reactor from overheating. Although the **°U concentration is not
rods large enough for a nuclear reactor to explode like an atomic bomb,
overheating can cause sufficient damage for radioactive materials to be released
into the environment, as happened at Chernobyl in the former Soviet Union
in 1986,
The neutrons produced by **U fission are highly energetic (ca. 2 x 10° kJ
mol’), which is problematic in that they are not very good at causing fission
in other ~*°U nuclei. Fission is much more efficient with less energetic
Circulating neutrons, ideally ‘thermal’ neutrons with energies of ca. 2 kJ mol’.
cooling fluid/
Moderators are therefore used in nuclear reactors to slow down the *°U
moderator
neutrons. The best moderators are light nuclei such as '*C or 7H (D), and thus
in heavy water reactors the cooling fluid may also act as the moderator.
Fig. 2.8 Schematic diagram of a
nuclear reactor,
Separation of the isotopes of uranium
One of the greatest difficulties faced by the scientists working on the
Manhattan Project was to obtain enough **U to build a nuclear device. The
first attempts to separate *°U from **U were made by Ernest Lawrence at the
University of California in Berkeley. Lawrence’s approach was to take UCI,
and pass it through a cyclotron-like device in which ionized *°UCI, followed
a slightly different path from the heavier *“U species. Unfortunately the
The f elements 17
process did not work at all well, and only about | g of **°U was produced by
Lawrence.
The technique that was developed as an alternative relies on gaseous
diffusion. This approach, which has been extensively used in the production
of **°U enriched fuel for nuclear reactors, passes gaseous UF, through porous
metallic membranes (usually composed of nickel or aluminium) at 70-80 °C.
Graham’s law states that the rate of diffusion is inversely proportional to the
square root of the relative molecular mass, and thus repeated passes through
the membranes results in the mixture of “*UF, and **°UF, becoming richer in
the lighter molecules. Up to 3000 passes are made in practice, leading to a
90% concentration of *°UF,. Unfortunately gaseous diffusion plants are very
large and expensive to run. All of the materials that come into contact with
the UF, must be fluorine resistant, and the metal membranes must be
manufactured to high tolerances. Furthermore, the pumping of the UF,
through the plant is energetically very demanding.
Another method of isotope separation involving UF, is the gas centrifuge.
If UF, is spun in a gas centrifuge the heavier ***UF, will concentrate toward
the walls and the UF, in the axial position. The higher the rotation speed
and the lower the temperature the better the separation.
Lasers have also been used to separate *°U from ***U. One approach is
based on the fact that **°U has a slightly different ionization energy from **U,
and hence the irradiation of uranium vapour with a laser whose wavelength ts
tuned to the ionization energy of **°U will produce exclusively *°U*, which
can then be collected at a negatively charged electrode.
aethtpies
ool
“ppt grat U},5 = 22 min) (2.1)
233Pa—733U +B (t,,. = 27days)
SOS errs B
“(t,, = 24 min)
3Np
233U-373+B (2.2)
233Np—754Put+B (t,,, = 2.36days)
The vast majority of the uranium in conventional reactor fuel is **U.
Hence some *°’Pu is produced in all operating reactors (via the processes
shown in Egn 2.2) although the neutron yield from the 2-3% *°U is
insufficient to produce significant quantities of plutonium. However, if more
neutrons were available the production of **’Pu from ***U would increase to
the point at which it would become greater than the consumption of *°U.
This may be achieved in practice by removing the moderators in conventional
reactors and enriching the fuel to a greater extent than usual, to produce the
18 Properties of the atoms and ions
so-called ‘fast breeder’ reactors. The advantage of such reactors is that both the
>33UJ and **8U are used to produce heat, allowing up to 60 times the energy to
be extracted from naturally occuring uranium. However, not only does the
design and construction of fast breeder reactors pose many difficult technical
problems, but there are serious political objections to the use of a technology
which generates fissile >*°Pu, as it may be used to make nuclear weapons (the
‘Fat Man’ bomb dropped on Nagasaki was a **’Pu device). Thus fast breeder
reactors remain at the prototype stage. 4
Irradiated fuel
elements
‘Cooling off’ period (up to 100 days). Rods are immersed in ponds of water to allow the intensely
radioactive but short-lived decay products such | as '3'| (t,.=8.04 days) to lose most of their activity
Aqueous solution of
Solution treated with TBP in kerosene fission products and
transplutonium elements
[UO,(NOs)a(TBP)2] and
[Pu(NO,) ,(TBP),] in kerosene
Fe(NHzSO3)o, hydrazine, or hydroxlamine nitrates | reduce Pu(IV) to Pu(Ill) but do not affect U(VI)
UOx(NO3)2.6H20 Precipitate of
Pu(Cp O4)2.6H2O
400 cal
300 °C
[ra]
UO;
H,/700 °C
UO,
Fig. 2.9 Flow scheme showing the principal steps in nuclear fuel reprocessing.
20 Properties of the atoms and ions
2.5 Exercises
1. | Radioactive decay follows first order kinetics, i.e. ifN is the number of
atoms of a particular isotope at time rt, N, the number at time r=0, and
ty is the half-life, then
0.693
N = No exp(- yt
fy
How long does it take for a sample of **°Pu to decay to 10% of its
activity? (Note that the activity of a radioactive sample - the rate at
which it decays - depends only on the size of the sample).
tO How many (a) radial and (b) angular nodes do the following atomic
orbitals possess?
St 75 Gd 6p 7p
4. Nuclear fission is not the only nuclear process on which the generation
of electricity may be based. Research continues into the feasibility of
using nuclear fusion in electricity production. What are the principal
features of nuclear fusion, and for which elements is it an exothermic
process?
3 Relativity, electronic
spectroscopy, and magnetism
ral
In atomic units, the average radial velocity, <v,,y>, of the electrons in the
ls shell of an atom is approximately Z, where Z is the atomic number. For
uranium, for which Z = 92, <v,,,>/c is given by Eqn 3.2, where the velocity
of light is also expressed in atomic units.
a ere ee fe
ai — = —— = 0.67 (3:2)
e 137
The average relativistic mass increase of the Is electron in uranium is
therefore given by Eqn 3.3 where m, is the rest mass of the electron. As the
22 Relativity, electronic spectroscopy, and magnetism
expression for the Bohr radius has a 1/m dependence, this mass increase
produces a marked contraction of the Is electron, and a concomitant energetic
stabilisation.
me = 135m, (3.3)
a A= 0:67"
Russell-Saunders coupling
Russell-Saunders coupling considers that the individual orbital angular
momenta of all the electrons combine into a total atomic orbital angular
momentum with quantum number L. Similarly, the individual electronic spin
angular momenta combine to yield a total spin angular momentum for the
whole atom with quantum number S. The total atomic angular momentum is
given by the coupling of L and S and is described by the quantum number J.
Atomic or ionic energy levels are characterised by a term symbol, of
general form
Cen
J
The value of S is then used to obtain the spin multiplicity, (2S+1), of the
term. Finally the permitted values of J are determined according to Eqn 3.4,
and are known as the levels of the term.
PES, LRA (Leo! (3.4)
The most important term of an atom or ion is the ground term (the most
stable term). Fortunately there is a simple way to determine atomic/ionic
ground term symbols, using the three rules due to Hund.
Rule 1: The ground term always has the largest value of S. This rule is
known as the rule of maximum multiplicity.
Rule 2: If two terms have the same multiplicity, the one with the highest
value of L lies lowest in energy.
Rule 3: For electronic subshells that are less than half full, the level with the
lowest value of J lies lowest in energy. For greater than half-filled
subshells, the level with the highest value of J lies lowest in energy.
let us determine the ground term arising from the [Xe]4f? configuration of
Dy**. The best way to do this is to use Hund’s rules and the
‘electrons-in-boxes’ approach, in which the individual f orbitals are
represented by boxes and the electrons by arrows (up-and down arrows are used
to represent up spin (m, = +1/2) and down spin (m, = -1/2) electrons
respectively). The arrangement of the nine 4f electrons of Dy** which satisfies
Hund’s first two rules is shown in Fig. 3.1.
Fig. 3.1 The most stable arrangement of the nine 4f electrons of Dy**.
This arrangement has the maximum number of unpaired electrons, five, and
hence an S value of 5/2 (the sum of all of the individual electron m, values)
Note that we need consider only the
and (2S+1) = 6. The value of L in the above arrangement is five (the sum of
partly filled 4f subshell to obtain the
ground level. This is because all filled the individual electron m, values), and hence the ground term is °H. J can take
electron shells and subshells have no integer values from 5+5/2, 5+5/2-1......, |5-5/2| and because Dy* has a
net angular momentum and hence do greater than half filled 4f shell, the level with the highest J value (15/2) lies
not contribute to the term symbol.
lowest in energy. Thus the full symbol for the ground level of Dy** is °H,<,.
The ground levels of all of the Ln** ions are given in Table 3.1.
L.,/Bohr Magnetons
Ln* Electronic Ground Colour Calculated Observed
Configuration _ level (Eqn 3.5)
Ce* — [Xe]4f' "Bos Colourless 2.54 ya
oesNe:
Pr* [Xe]4f 3H, Green 3.58 3.4-3.6
Nd* — [Xe]4f? 4, Lilac 3.62 Dea
Pm* [Xe]4f* y, Pink 2.68 :
Sm* — [Xe]4f? 23 Yellow 0.85 (Aa7
Eu* [Xe]4f° usr Pale pink 0 33255
Gd* — [Xe]4f’ a Colourless 7.94 7.9-8.0
Th [Xe]4f* USP Pale pink 9.72 9.5-9.8
Dy* [Xel4f? Hg, Yellow 10.65 10.4-10.6
Ho* =‘ [Xe]4f"° *t. Yellow 10.60 10.4-10.7
Et [Xe]4f" “Hisp Rose-pink 9.58 9.4-9.6
Tm* = [Xe]4f”” 7H Pale green 7.56 That 5
Yb* =‘ [Xe]4f "sys Colourless 4.54 4,3-4.9
Lu*__ [Xej4f* Sa Colourless _0 0
Table 3.1 Spectroscopic and magnetic properties of Ln* ions in hydrated salts.
J-j coupiing
In this coupling scheme, the individual electronic orbital and spin angular
momenta combine to give a total angular momentum for each electron,
The f elements 25
denoted j. The j values then couple to produce the total atomic angular
momentum J.
Spin-orbit coupling in actinide atoms and ions and their compounds is
much greater than for the lanthanides, to the point that the Russell-Saunders
coupling scheme is much less valid. It would be both elegant and convenient
if we could treat actinide spin-orbit coupling using the j—j scheme, but
unfortunately a purely j—j based approach does not work either. This is partly
because even in actinide systems spin-orbit coupling does not dominate over
interelectronic repulsions, and partly because the 5f orbitals are much more
sensitive to atomic/ionic environment than are the lanthanide 4f orbitals
(particularly the 5f orbitals of the early actinides). Experimental studies of
actinide spin-orbit coupling, for example electronic spectroscopy and
magnetic measurements, are consequently more difficult to interpret than
analogous lanthanide data. The spectroscopic and magnetic properties of the
lanthanides and actinides are discussed in more detail in the following
Sections.
spectrum, the colours of Ln** compounds are typically less intense than those
of the transition metals.
The colours of the Ln** ions in hydrated salts are given in Table 3.1. The
lack of 4f orbital/ligand interaction means that the f — f transition energies
for a given Ln* change little between compounds, and hence the colours of
Ln* are often characteristic. A further consequence of the small interaction of
the Ln** 4f atomic orbitals with the surrounding ligands is that f — f
transition energies in Ln** compounds are well defined, leading to much
sharper bands in their electronic absorption spectra than are observed for
transition metal compounds. A typical spectrum, that of aqueous Pris
shown in Fig. 3.2, and the energies of the levels of free Pr** relative to the
3H, ground level are given in Fig. 3.3.
25000
3 P5
3 P,
20000
24 20 16 10 5 0
emi x 107
Fig. 3.2 Electronic absorption spectrum of aqueous Pr* [modified from Fig. 11.11 (b) of
15000
‘Physical Inorganic Chemistry’ by S.F.A. Kettle].
°F; ground level, but the f > d transition has been reported at only 22 000
cm! in Ce**-doped Y,A1,O,,. Even more remarkably, the peak at 17 650 cm’!
in the spectrum of [Ce{n°-C,H;(SiMe;),},] has also been assigned to the
f — d transition. The large differences in the energy of this transition for
different Ce** environments reflects the greater radial extension of the Sd
atomic orbitals with respect to the 4f, and their greater interaction with the
surrounding ligands.
Note that Ln** ions are often highly coloured. This arises because the 4f
orbitals in Ln’* are destabilised with respect to those in Ln**, and hence lie
closer in energy to the 5d orbitals. This change in orbital energy separation
causes the f — d transitions to shift from the ultraviolet into the visible
region of the spectrum.
Excited
singlet
Excited
triplet
Excited
Ln3* level
Lns+
fluorescence
This process is particularly favoured for Tb** and Eu**, which have excited
levels at slightly lower energy than the excited triplets of typical ligands. The
main emissions for Tb** are between the °D, and ’F, (n = 6-0) levels, and
generate green light, while for Eu** the °D, — ’F, (n = 4-0) transitions emit
red light. This fluorescence is employed in colour television sets, the screens
of which are made up of a large number of tiny clusters, each containing three
phosphor dots. The three dots in each cluster emit red, green, or blue light
respectively. The red phosphors are typically Eu** in Y,0,S or Eu**:Y,0,,
while one of the choices for green emission is Tb**:La,O,S. The best blue
emitter is Ag,Al:ZnS, which has no Ln** component. The television set has
three separate cathodes, one for each colour. A metallic mask behind the
screen has tiny holes in it which allow only electrons from the green cathode
The f elements 29
to hit the green phosphor dots in each cluster, and similarly for the other two
primary colours.
There is currently a great deal of research into replacements for the cathode
ray tube in television sets. One promising prospect for the generation of
colour in flat panel displays are tris(pyrazolyl)borate compounds of cerium,
europium, and terbium (Fig. 3.5), which have the potential for bright, highly
efficient, durable, low power light emission with a narrow bandwidth over the
full spectral range.
Fig. 3.5 Tris(pyrazolyl)borate complexes of cerium, europium, and terbium are promising
phosphors for flat panel displays.
etc
[Xe]4f8
4145/2 0.74
I Alaa 0.52
“lit 0.28
“loro 0
3.7 Magnetism
Notice the similarity between Eqn 3.5 and the spin-only formula which
For first row transition metal
works so well for first row transition metal compounds. In these compounds
compounds the effective magnetic
the orbital contribution to {,, is quenched by the interaction of the d orbitals moment is well approximated using
with the surrounding ligands, so that only S$ is required (and hence only the the spin-only formula
number of unpaired electrons). The replacement of § by J in Eqn 3.5 is Hey = ajn(n + 2)
necessary because the surrounding ligands fail to quench the orbital where n is the number of unpaired
contribution to {l,, in Ln** compounds, owing to the 4f orbitals being so electrons.
contracted.
The calculated (Eqn 3.5) and observed u,, values for all of the Ln’* are
given in Table 3.1. It may be seen that there is good agreement between the
calculated and experimental values in all cases except for Sm** and Eu**. The
discrepancies for the latter arise because both ions have excited levels (°H,,
for Sm** and ’F, and ’F, for Eu**) which are sufficiently close to the ground
level to be thermally accessible. If allowance is made for this (by assuming a
Boltzmann population distribution over the energy levels) then calculated and
experimental [,, values once again agree.
The high J of the ground levels of the later Ln** result in very high [,
values (Table 3.1). As a consequence, placing salts of these ions in strong
magnetic fields leads to a slight warming because the stabilisation energy
given out as the salts are attracted into the field manifests itself as heat. This
effect is exploited for obtaining very low temperatures. Gadolinium and
dysprosium salts are typically used, and are cooled with liquid helium in the
presence of a strong magnetic field. When the system is at liquid helium
temperature the magnetic field is removed, which causes further cooling as the
salts lose their magnetic orientation, a process known as _ adiabatic
demagnetisation.
The strongest known permanent magnetic material is an alloy of
neodymium, iron, and boron, of chemical formula Nd,Fe,,B. The unit cell of
this material contains 68 atoms, with six distinct iron sites, two different
neodymium sites, and one boron site. The iron and neodymium sites each
have their own magnetic moments, and these align in the same direction to
produce a bulk magnetism more than 50 times that of steel. Compounds with
the general formula Ln,Fe,,B have been identified for all of the lanthanides
except promethium and europium, although none has the same permanent
magnetic strength as the neodymium alloy.
One of the most widely studied actinide ions is U* ({(Rn]5f’). The magnetic
data obtained for U** compounds are usually interpreted by considering only
In spite of the reduced applicability of
the Russell-Saunders coupling
the 7H, ground level but unlike the Ln** ions, the effects of the surrounding
scheme to actinide electronic ligands must also be taken into account. The *H, ground level is spilt by the
structure, term symbols based upon surrounding ligands into several new energy levels, and the magnitude of the
this approach are often used as the splitting is comparable to thermal excitation energies. The interaction of the
starting point for discussions of the
5f atomic orbitals with the surrounding ligands therefore creates a range of
spectroscopic and magnetic
properties of actinide compounds.
thermally accessible excited levels from the free ion ground level.
There have been several magnetic studies of [NEt,],[U(NCS)g], which
above 30 K contains U** in a site of cubic symmetry surrounded by eight
nitrogen atoms. Magnetic data have been used to show that below 30 K, the
geometry distorts to Dy, around the U**. [U(n?-C,H;);R] (R = BHy, BF,, OR,
F, Cl, Br, I) have also received a good deal of attention, and have been divided
into two categories on the basis of their magnetic behaviour: (a) molecules
with small dipole moments and a small range of TIP and (b) molecules with
larger dipole moments and a more extended range of TIP. The differences
between the two types of behaviour are attributed to an increasing trigonal
distortion for molecules in category (b).
The magnetic properties of the actinidocenes - [An(n*-C,H,).] - are
discussed in Chapter Six, Section Six.
3.8 Exercises
1. | Use Hund’s rules to determine the Russell-Saunders ground levels of
3. Why are aqueous solutions of Ce** colourless but those of Ti** purple?
4.1 Halides
LnF3
HF(g)
Ln2O03
700 °C
LnF,; (Ln = La—Pm) adopt the ‘tysonite’ (LaF,) structure (Fig. 4.1) in
which the Ln** is coordinated by nine F in a tricapped trigonal prismatic
arrangement, with a further two F at a slightly greater distance. Beyond Fig. 4.1 The LaF, (‘tysonite’)
promethium, all of the LnF, have the YF; structure which features eight close structure. The shaded circles
represent the two more distant fluorine
Ln3*-F contacts at ca. 2.3 A and one longer Ln**-F' distance (ca. 2.6 A),
atoms.
with the F in an approximately trigonal prismatic arrangement around the
Ln*. This reduction in the primary coordination number from nine to eight is
a result of the decreasing size of the Ln**.
LnF, have several technological uses, including thin film coatings of
optical elements, high pressure anti-wear lubricants, and host lattices for
phosphors and scintillators.
34 Solid state compounds
Cel
HCI(g)
LnCl3.xH,O
105-350°C
LnCl,
LnCl3.xH,O
reflux
LnCl, (Ln = La—Gd) adopt the nine coordinate UCI, structure, a tricapped
trigonal prismatic arrangement which is like the structure of LaF, but with
the two’ more distant F removed. This structure is also adopted by
{Ln(H,0),]** of the early lanthanides. TbCl, has the eight coordinate PuBr,
structure (which may be regarded as the UCI, structure with one of the
capping Cl removed) and all of the other lanthanide trichlorides the six
coordinate AICI, structure. Thus the trend toward lower coordination number
with decreasing Ln** radius, noted above for LnF;, is also apparent in the
trichlorides. The size of the anion is also important in determining the
coordination number of a given Ln**, with a trend toward decreasing
coordination number with increasing anion radius.
Routes to LnBr, and Lnl, include:
HBr(g) HI/H>
n> nis Lnclj. —————> Ll,
400-600 °C heat
There are three different structural types among the LnBr,. LaBr,, CeBr,,
and PrBr; ali have the UCI, structure, while the tribromides of neodymium-—
europium adopt the PuBr, structure. The remaining LnBr, have the six
The f elements 35
coordinate FeCl, structure. LnI, have either the PuBr, (Ln = La—Pm) or FeCl,
structure.
F Gil Br I
ic ~ - - MoSi,/8
PG - - _ MoS1,/8
Nd ~ PbC1,/7+2 PbCL,/7+2 SrBr,/7,8
Pm CaF,/8 ~ - -
Sm CaF,/8 PbCI,/7+2 PbClI,/7+2, Eul,/7
SrBr,/7,8
Eu - PbCI,/7+2 SrBr,/ 7,8 Eul,/7
Gd - = - MoS1,/8
Tb - - - —
Dy - SrBr,/7,8 SrI,/7 CdCl,/6
Ho ~ - - —
Er - - - -
Tm - orLJ7 SrL/7 CdL/6
Ao} CaF,/8 SrL/7 Sri, CdlL,/6
CaCl,/6
Ga - - - -
Table 4.1 Dihalides of the lanthanides. The structural type and metal coordination number
are indicated.
which consist of trans edge sharing Ln, octahedra arranged in single chains
with the Cl ions capping the faces of the octahedra. Further reduction can lead
to compounds of the general formula LnXH, (X = Cl, Br; Ln = Sc, Y, Gd,
Lu), which again feature edge sharing Ln, octahedra (now arranged in double
metal layers) but which also have interstitial hydrogen atoms.
iz Fo C.c
4. > uF (6 a ee es
250-400 °C Bt 4
BCl, lp
UF; ——. UC U Uly
-107 °C 500 °C/20 kPa
HBr 900 °C
UF; ————— UF, Urn See URE,
CCl, Br,
al tee
160 °C/20 atm 500 °C
(Fig. 4.1), but BkF, and CfF, have the eight coordinate YF, structure. Note
that this change in structural type is also seen for LnF, but occurs three
elements further along the actinide series on account of the greater size of the
actinide ions. A similar reduction in coordination number occurs in AnBr,,
with AcBr; to NpBr, adopting the nine coordinate UCI, structure and PuBr, to
BkBr; (plus another NpBr, phase) having the eight coordinate PuBr, structure.
Actinide hexahalides are confined to UF,, UCI,, NpF,, and PuF,. All four
are strongly oxidising and highly moisture sensitive. The pentahalides are
more numerous; all four PaX,; are known as well as UX, (X = F, Cl, Br) and
NpF,. All AnX, are very moisture sensitive. All four tetrahalides are known
for thorium—neptunium, although only the flouride is known for plutonium—
californium. Eight coordination is common, often in dodecahedral geometry
(e.g. ThCl, ThBr,, and UCI,). The UX, family provides another illustration of
decreasing coordination number with increasing anion radius; UF, and UCI,
feature eight coordinate uranium, UBr, seven coordination (pentagonal
bipyramid) and UI, six coordination via edge sharing UI, octahedra. AnX, (X
= Cl, Br, I) are hygroscopic, and are readily soluble in polar solvents such as
(CH;),CO and CH;CN.
AnF, (An = U-Es) are high melting and insoluble in water, by contrast to
the other trihalides of these elements which are hygroscopic and water
soluble. Many may be crystallised out of water solutions as the eight
coordinate hexahydrates [AnX,(H,O),]*. The only trihalides of thorium and
protactinium are the triiodides; ThI, is poorly characterised, but Pal, is well
established. It may be made by heating Pal, at 360 °C in vacuo, and has the
eight coordinate PuBr; structure.
AnX, (An = Am, Cf, Es; X = Cl, Br, I) and ThI, are known. The latter has
metallic properties and has two forms, @ (black) and B (gold). By analogy
with the [Xe]4f’ configuration of Eu’, it might be anticipated that americium
would have a divalent chemistry, and the existence of black AmX, supports
this expectation.
4.2 Oxides
All of the Ln,O, are strongly basic. They are insoluble in water but
dissolve readily in aqueous acids to produce solutions that contain
{Ln(H,0),]* (provided the pH is kept below 5). Ln,O, absorb both CO, and
H,O from the atmosphere.
LnO (Ln = Nd, Sm, Eu, Yb) may be prepared (Eqn 4.1) by reduction of
Ln,O, with elemental lanthanide at high temperature and pressure (high
pressure is not required for europium)
Ln + Ln,O; — 3LnO (4.1)
All four LnO have the NaCl structure. However, while EuO and YbO are
insulating or semiconducting, NdO and SmO exhibit metallic conductivity.
The mechanism for this process is believed to be the same as that discussed
for LnS in Section 4.3.
LnO, (Ln = Ce, Pr, Tb) adopt the fluorite structure in the limiting dioxide
stoichiometry, although a range of non-stoichiometric phases exists between
Ln,O, and LnO,. CeO, (‘ceria’) is white when pure but is usually pale yellow
on account of the non-stoichiometric phases. The non-stoichiometry of ceria
is exploited in the catalytic converters used to remove pollutants from motor
vehicle exhaust emissions. These converters contain several catalytically
active components, including platinum and other metais from groups nine and
ten, as well as ceria. Ceria has several functions in the catalytic converter:
e Promotion of the water-gas shift reaction
CO + H,O > CO, + H, (4.2)
Oxide superconductors
Prior to the mid 1980s the highest recorded value of 7, (the temperature at
which a material becomes superconducting) was ca. 23 K for Nb,Ge. Then in Superconductivity was discovered in
1911 by the Dutch physicist H.
late 1986 Bednorz and Miiller reported their discovery of a new ceramic
Kamerlingh Onnes. He cooled
compound, La, ,.Ba,CuO,, which has a T. of 30 K, and in March of the mercury to below 4.2 K and observed
following year Wu and Chu reported a T, of 92 K for YBa,Cu,O,; (0 < 8 < that it lost all resistance to the flow of
1). This latter result is especially significant as a material that is an electrical current.
superconducting at liquid N, temperatures (between 63 and 77 K) is much
more likely to find a practical use than one with very low T..
@ =Cu
O-0
= O (removed to create
O-deficient phases up
to YBa,Cu,0¢)
under 23.5 GPa of hydrostatic pressure) and all feature puckered planes of
CuO, (as does YBa,Cu,0,) which are believed to provide a path for
superconduction. The exact mechanism of superconductivity remains
uncertain, although it is widely believed that in a-superconducting material
pairs of electrons (known as Cooper pairs) move through the solid, the first
electron distorting the lattice in such a way that the second can follow it very
easily.
Unfortunately the commercial exploitation of superconductivity in, for
example, magnetic devices, power transmission, and communications, is
hampered by the ceramic nature of the high 7. compounds. The innate
brittleness of ceramics prevents their extrusion into wires, and it now seems
that the most promising way forward is the deposition of thin films of
superconductor onto metal oxide or silver surfaces to form flexible tapes.
fed) +6
s
D
=
@ U;0;
2 +5 Pa,Os ®
o
x AmO>;
o +4 e NpO, @ @ ®@ @ 8kO,
E
+3 P @ @€£5,0,
Cm, O3 Cfo Oz
Th Pa U Np Pu Am Cm Bk Cf ES
Chapter Two, Section Four) has the fluorite structure (as do all of the actinide
dioxides) and this is preserved on the addition of extra oxygen until U,O, is
formed, at which point the structure is related to fluorite but with interstitial
oxygen. The next well characterised phase is U,O,, which is the end product
of heating any uranium oxide in air at temperatures above 650 °C, and which
contains pentagonal bipyramidal {UO,} units. UO,, the only anhydrous
actinide trioxide, has several crystalline forms, most of which contain uranyl]
(UO,”) groups linked by bridging oxygen. One form (y-UO,) has a structure
based on edge and corner sharing UO, octahedra. A variety of preparative
routes exists to UO;, which may be reduced by H, to UO).
400 °C
UO,.2H,O
air, 500 °C Hp
0, U0, —— U0,
300-600 °C
Oz
UO>(NO3)5.6H,O
400-600 °C
NpO, is the usual end product when neptunium compounds are burnt in air
at elevated temperatures. Reaction of Np(I[V) hydroxide with O, gives
NpO,.H,O, which when heated to 300 °C under vacuum produces the highest
oxide of neptunium, the non-stoichiometric Np,O;. Beyond neptunium the
highest oxide of each element is the dioxide, which is formed by heating of
the oxalate or hydroxide in air or O,. The sesquioxides become increasingly
stable beyond plutonium, and have the same structures as the analogous
lanthanide compounds. The C-type is the most stable form of each An,O,,
but some elements also form the A- and B-type structures (e.g. Am,O,
exhibits all three structural types).
electron (known as the Fermi level) can be easily promoted into low-lying
empty energy levels. As a result they are relatively free to move through the
solid, giving rise to high electrical conductivity. By contrast, electrons in a
solid in which all the bands are full cannot undergo this process, and the solid
will be an electrical insulator.
YDODOOOD OOO
Fig. 4.4 Ina line of half-filled, weakly interacting s atomic orbitals, electron-electron repulsion localises the electrons one per atom
(above), in order to avoid having to pair two electrons in the same atom (below).
DOODOBOOO
We may quantify these ideas as follows. The energy required to remove an
electron from an atom is the ionization energy, J. If we now place the electron
in an atom which already has a half-filled s orbital we get back the electron
affinity of that atom, A,. The energy required to move the electron is therefore
given by Eqn 4.3, where U is called the Mott-Hubbard splitting, or Hubbard
U, and may be interpreted as the repulsion energy between two electrons in
the same atom.
The f elements 43
U=I-A (4.3)
In order for metallic conductivity to occur, the overlap of the s orbitals must
be sufficiently large that the band width, W, is greater than the value of U. If
U > W, electron-electron repulsion prevails and the solid is insulating.
Elemental lanthanides
We have previously seen that the 4f orbitals of lanthanide atoms and ions are
radially contracted and interact to a small extent with the surrounding ligands
in lanthanide compounds. A further consequence of this contraction is that the
overlap of the 4f atomic orbitals on neighbouring atoms in the elemental
lanthanides is very small, and hence 4f band widths are also small (typically <
0.1 eV). The metallic nature of the elemental lanthanides comes from the 5d
and 6s atomic orbitals, which are much more diffuse and overlap to a much
greater extent. That the 4f orbitals cannot be responsible for the metallic
properties of the lanthanide elements is confirmed by Fig. 4.5, which plots
the value of the Hubbard U for the lanthanide 4f orbitals. These data have
been experimentally determined using solid state X-ray photoelectron
spectroscopy, and clearly show that U is very much greater than 0.1 eV for
the 4f orbitals of all of the lanthanides.
Ce Pk Nd Pm sm Eu Gd Tb Dy Ho Er Tm
Fig. 4.5 Hubbard U values for the 4f atomic orbitals of the lanthanides.
4.4 Exercises
hie LnO (Ln = Nd, Sm, Eu, Yb) have the NaCl structure, while LnO, (Ln
= Ce, Pr, Tb) and all of the actinide dioxides have the CaF, structure.
Both of these structural types may be described by the filling of holes
in face centred cubic (fec) lattices. Which type of hole is filled in each
case, and which ions form the fee lattice? What is the coordination
number of the anions and cations in each structural type, and what is
the relationship between the ratio of the anion:cation coordination
number and the stoichiometry of the oxide?
to From the values of AH,° for LnCl,(s) given below, estimate values for
EuCl,(s) and YbCI,(s). Compare your estimates to the experimental
values of -936 and -960 kJ mol’ respectively. Rationalise the
differences between the estimated and actual values. Why is there a trend
toward decreasing AH,° for LnCl,(s) with increasing Ln atomic
number?
Fig. 5.1 Molecular structure of [Nd(NO,).(H,O),]*. The f element ions are large, hard acids
and form high coordination number complexes with hard bases.
Oxidation states
As we have seen in Chapter Two, Section Two, the early actinides have many
more stable oxidation states than do the lanthanides.
Lanthanides
The divalent lanthanides tend to reduce water to hydrogen (Eqn 5.1), although
Eu* is more stable in aqueous solution than are Sm** and Yb**. Also, since
all these ions are oxidised by molecular oxygen, the solutions must be
handled under an inert atmosphere. The tetravalent ions of neodymium,
dysprosium, praseodymium, terbium, and cerium oxidise water to oxygen
(Eqn 5.2), and of these only Ce** is sufficiently kinetically stable to form
48 Coordination chemistry
aqueous coordination compounds. Hence, the trivalent ions Ln* are by far the
most readily studied.
2Ln**(aq) + 2H*(aq) = 2Ln**(aq) + H,(g) (520)
Actinides
The oxidation states adopted by the actinides and the-relative stabilities of the
di-, tri-, and tetravalent states are represented in Figs. 2.3 and 2.4. In this
section we will illustrate the general trend of increasing stability of the lower
oxidation states across the series. It is worth noting that complexes of the
actinide ions in the same oxidation state usually have the same structures, for
example, the hexavalent ions all exist as AnO,”* in aqueous solution.
We have noted that the stability of the trivalent state for the actinides
increases as we move across the series. Th** is rapidly oxidised by water, and
the aqueous chemistry of this ion is exclusively of the tetravalent state. The
U* ion is, however, readily obtained by reduction of higher valent species and
is stable for days in the absence of air. Eventually it reduces water to
hydrogen, and the reaction in Eqn 5.3 is accelerated by the presence of acid.
The tetravalent ion U* is quite stable in the absence of oxygen.
U*(aq) + H,O(l) > U*(aq) + 1/2H,(g) + OH’ (aq) (5.3)
Moving further along the 5f series, aqueous solutions of Np** are stable in
air, and this ion is readily prepared by electrolytic reduction of acidic solutions
of Np**.. By the time we reach americium, the trivalent ion is the preferred
state. All the actinide elements after protactinium can exist as An** in aqueous
solution.
A corresponding trend is observed for the penta- and hexavalent actinyl
ions. Their stability falls across the series from uranium, for which UO,”*
predominates in the presence of air, to curium for which the actinyls are
unknown. The physical properties of these ions are dealt with separately in
Section 5.3.
The redox chemistry of plutonium in aqueous solution is fascinating in its
complexity. The element is capable of attaining five oxidation states in water,
from Pu(II) to Pu(VIJ), although the latter only exists in strong alkali.
The redox potentials (see above) separating the most common states are
roughly equal. Interconversion between oxidation states occurs through
disproportionation and conproportionation reactions, and the matter is further
complicated by the reducing effect of the a decay of the plutonium atoms.
Given these factors, it is not surprising that all four oxidation states Pu(IID) to
Pu(VI) commonly occur together in aqueous solution, and it is difficult to
prepare pure solutions of any one ion. The redox couples Pu(IV/III) and
Pu(VI/V) are reversible, and rapid reactions occur with one-electron oxidising
The f elements 49
and reducing agents. Those couples which would require the making or
breaking of bonds, e.g. the oxidation of Pu** to PuO,", are irreversible.
kx10’
dm*
mol
s"
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Ef Tm
Fig. 5.3 Rate constants for the formation of 1:1 Ln*/oxalate complexes at 25 °C.
The f elements 51
The rates are essentially constant at ca. 8 x 10’ mol dm® s"' in the region
La* to Eu* but then fall rapidly via Gd** and Tb** to another region Dy** to
Tm* where they are again constant at ca. | x 10’ mol dm® s’'. It is worth
noting that the ion midway between the two regions of constancy is Gd**, and
it is widely held that it is at this point along the lanthanide series that the
coordination number in aquo complexes changes from nine (early lanthanides)
to eight (late lanthanides). This reduction in the number of water molecules in
the primary coordination sphere would lead to a sudden increase in the
strength of the remaining Ln—OH, bonds. Hence, if the breaking of this bond
is rate determining, there will be a dramatic reduction in the rate of
substitution.
The stoichiometric mechanism in Fig. 5.4, consistent with these data, has The stoichiometric mechanism
describes the sequence of elementary
been proposed. It has the following steps.
steps by which the reaction takes
(i) The incoming ligand becomes associated with the aquo complex, place.
probably through hydrogen bonding with the highly. polarised water
ligands.
(ii) One of the water ligands begins to dissociate, perhaps under the
influence of the incoming group.
(1) The loss of water ligand (breaking the strong Ln—O bond) is rate
determining.
Since oxalate is a bidentate ligand, it might be expected to replace two
water ligands, and not just one as shown for simplicity in Fig. 5.4. A second
substitution reaction converting an 7'-oxalate to n° will follow step (iii).
(i)
[Ln(H»O)9]**(aq) + (oxalate)* (aq) ee ([Ln(H2O)g--7*- (oxalate)]*(aq)
(iii)
[Ln(H2O)g, (oxalate)]*(aq) + HO | (LO) (oxalate): (H2O)}*(aq)
Fig. 5.4 A stoichiometric mechanism for substitution of H,O ligands for a simple bidentate
ligand such as oxalate (C,0,*).
Perhaps the most closely related species to the actinyl ions in the d block
is the d? osmyl(VI) ion OsO,** which is also linear with Os—O double bonds
of ca. 1.75 A. Since the ionic radius of U(VI) is nearly 0.2 A greater than that
of Os(VJ) it is quite surprising that in the uranyl ion UO,"* the U—O distances
are as short as 1.7 to 1.8 A. This is taken as good evidence for the presence of
a significant degree of multiple (covalent) bonding in the actinyl ions (see
later).
A further key feature of the structure of coordination complexes of the
actinyl ions is that the auxiliary ligands are almost always accommodated in
the equatorial plane. Usually between four and six ligands are supported in
this way giving (a) octahedral, (b) pentagonal bipyramidal, and (c) hexagonal
bipyramidal structures (Fig. 5.6).
2 O 3 O
20
os H Oo en a ae -O== N
=Cl. ro
ca e FSaes |ee
ee i2 | es
-=
Ay gl eee een
Ci-* ~ ios
& = SET
yeas Wis. 70?
O O
| i O
|
(a) (b) (c)
Fig. 5.6 Coordination in the equatorial plane of uranyl (VI) ions. (a) The octahedral dianion of the salt Cs,[UO,Cl,], (b) pentagonal
bipyramidal trianion in K,[UO,F.}, and (c) hexagonal bipyramidal uranyl nitrate dihydrate [UO,(NO,),(H,O),).
The f elements 53
The latter class of structure is only formed when three bidentate ligands such
as carbonate, nitrate, or sulphate are present. In the cases with six equatorial
monodentate ligands, the equatorial plane is puckered.
Actinides in the environment, whether they be naturally-occurring or
arising from accidental contamination, are usually found in the form of
actinyl ions, except of course in the case of the later actinides where the
trivalent ions are more stable. A highly topical area of research is in the
coordination chemistry of actinyls with ligands such as carbonate and
bicarbonate which are found in relatively high concentrations in natural
groundwaters. Such systems can be quite complicated, and several different
species may exist in rapid equilibrium with one another. However, using a
wide range of techniques, it has been shown that the dominant species in the
uranyl(VI) carbonate system at pH = 6 is the trimer [(UO,),(CO,),]°, Fig.
ay i
O----C-----0
Fig. 5.8 Examples of possible An—O bonding combinations in actinyl ions. (a) 7,: out of
phase oxygen p/An d,, and (b) 7: in phase oxygen p/An f,,2.
The principal features of the ordering of the six bonding molecular orbitals
shown in Fig. 5.9 have been established for the uranyl(VI) ion using a variety
of spectroscopic and theoretical techniques. It can be seen that the = bonding
molecular orbitals are lower in energy than the o bonding molecular orbitals,
implying that o bonding is less important than 7 bonding. This is quite
unusual since we normally expect 6 overlap to be more efficient than 1
overlap where p orbitals are involved. One explanation for this may be that at
short U—O distances there is antibonding overlap between the o-oriented
O----C-----0
oxygen 2p, orbitals and the toroidal lobes of the uranium d and f orbitals.
Fig. 5.10 The o,, bonding interaction This is illustrated in Fig. 5.10.
in the uranyl(VI) ion between uranium
f,3 and oxygen p showing the Ps ae anti-bonding (o and m)
possibility of destructive overlap.
Fig. 5.9 A schematic molecular orbital energy level diagram for the uranyl(VI) ion UO,**.
The f elements 55
Of course, not all the f and d orbitals on the metal are required, and some
remain non-bonding, for example the 5f based molecular orbitals which are of
5 and @ symmetry. In the case of UO,**, we have zero electrons from U(VI)
and six p electrons from each O* making a total of 12.' This completely fills
the bonding molecular orbitals shown in Fig. 5.9 and accounts for the Oo 2+
The chelate effect may also be explained in the following way. When one end
of a bidentate ligand attaches itself to a metal centre, the effective
Some of the dramatic increase in K for
Eqn 5.10 can be attributed to charge concentration of the other end of the ligand is artificially high, hence affecting
neutralisation of Nd** by EDTA* the equilibrium position. For macrocyclic ligands such as
which would loosen the solvation tetraazacyclododecanetetracetic acid (DOTA*) this phenomenon is amplified
sphere around the ions and thus since the ligand is unlikely to be able to distort in such a way so as to
increase the entropy of the system.
remove a ligating atom from the coordination sphere.
La Ce Pr Nd Pm sm Eu Gd Tb Dy
Fig. 5.12 Equilibrium constants K for the formation of 1:1 Ln*/EDTA* complexes in aqueous solution at 25 °C.
element ions in the aqueous phase and those attached to the resin [Eqn 5.12,
(r) denotes a resin-bound species].
Ln**(r) + 3citrate(aq) + 3Na*(aq) = [Ln(citrate),](aq) + 3Na*(r) (5.12)
The order in which the ions are eluted (Fig. 5.13) reflects the balance
between the affinity of the ions for the stationary phase and the stability
constant K of the chelate complex. The ions with the most stable chelate
complexes (i.e. the smallest ions) are eluted first.
————
eluant
in
concentration
concentration
eluant
in
——3>
Fig. 5.13 The elution of lanthanide (upper) and actinide (lower) ions from an ion exchange
column. The heavier elements appear first.
Ion exchange 1s rapid and selective, and since the positions of elements in
the elution series can be predicted with some accuracy, the technique was of
key importance to the discovery of new (transcurium) elements. The first
synthesis and identification of berkelium, element 97, is described thus by
Seaborg.
...we were somewhat surprised to see the rather large gap between
97 and curium; we shouldn’t have been surprised because there is a
notably large gap between the elution peaks of the homologous
lanthanide elements terbium and gadolinium.
The f elements 59
5.1 5.2
The first use of lanthanide shift reagents was in the 'H NMR analysis of
molecules which under normal conditions display overlapping peaks in the
alkyl region of the spectrum, and which contain a ligating group (e.g.
aliphatic alcohols and amines). Under the influence of the unpaired electrons,
the resonances are spread out over a much greater chemical shift range, thus
simplifying the spectrum. Also, since the groups closest to the paramagnetic
centre are most strongly affected, it is possible to use the lanthanide induced
shift to estimate structural data. With the advent of high field NMR
spectrometers, these reagents are used less often. However, chiral lanthanide
shift reagents such as 5.2 are widely applied to the NMR analysis of
mixtures of enantiomers. Since one enantiomer is bound more strongly than
the other, it experiences on average a larger contact shift of its resonances.
Hence, signals for the two enantiomers are separated. Integration of the
spectrum obtained allows the determination of enantiomeric excess.
Bonding
In principle, the alkoxide ligand RO» can act as a one, two, or five electron
donor to a metal centre depending on the hybridisation of the oxygen atom as
(a) sp’, (b) sp’, or (c) sp.
Since the lanthanide and actinide elements tend to form ionic complexes and
have large numbers of orbitals with suitable symmetry for overlap with
ligand orbitals, it is most usually the case that linear or near-linear R-O-M
fragments are observed, and we may describe the alkoxide radical as a five
electron donor. Similarly, the fragments R,N—M are almost invariably planar
and the dialkylamide radical is a three electron donor. In general, however, the
The f elements 61
number of electrons donated has little effect on how many ligands the metal
centre can accommodate. This is largely determined by steric effects (i.e. the
size of the ligands).
5.3 (a)
Me——04,,,, | w'tO— Me
"1 eet
om |
So Me
fo 9
5.5
For a given alkyl group R, the dialkylamide ligand R,N- is far more
sterically demanding than the alkoxide ligand RO-, simply because there are
two alkyl groups in the former. As a consequence, it is more difficult to
design monomeric alkoxide complexes of the f elements. For example, the
trivalent cerium complex [{Ce(OCHBu',),},] 5.6 is a dimer despite having
two ‘Bu groups on each alkoxide ligand 5.7. For this reason, several classes
of exceptionally bulky aryloxide ligands have been used. One of the most
successful 5.8 is derived from an inexpensive substituted phenol which is
used as an antioxidant in polymers and foods. The complex [Ce(OAr),] 5.9 is
monomeric and sublimes readily.
Me a
a Me
-O Me
Me Me
5.7
The f elements 63
Me Me
5.8
OMe 2-
- 4LICN MeOr,,, | ss OMe Pb‘+
UCI, + 6 LiOMe————> (Lit), 2S ————_ [U(OMe)¢]
Meo |™ ome
5.5
OMe
5.10
Fig. 5.14 Synthesis of a hexavalent uranium alkoxide via a tetravalent ate complex.
64 Coordination chemistry
Protonolysis
The protonated or acid form of the desired ligand is allowed to react with a
pre-formed complex such as an alkyl, dialkylamide, or alkoxide. This
strategy, which circumvents the problems mentioned earlier of ate complex
formation, works for the f elements because the bonds are so polar, and thus
is not so useful in the more covalent transition metal systems. The driving
force of the reaction (Fig. 5.15) can be provided by (i) the relative strengths of
the metal-element bonds or the pK, of the ligands (ii) the production of a
volatile ‘acid’ or (iii) the chelate effect (Section 5.4).
OH
ul But But But
Me3Si. H
—..--——_
ile ge
O Bu
Me3Si
Bu!
DAR
N(CH,CH NHR), 4 PhaNH
<< 1/2 [{U(NEt,),})] ————_ [U(NPho)4] + 4 EtpNHT
(it) 5.3 (i) 5.4
Fig. 5.15 Synthesis of f element alkoxides and dialkylamides using acidic reagents.
oe Ce(IV) 2
: silent ke ee mt oe
is OAc
Sml>
Fig. 5.17 Sml, promoted reductive cleavage of C—X bonds in the o.-position of ketones.
66 Coordination chemistry
H O-=> Smi(IIl)
PhCHp>Li/CeCl, OH
(a) OF
CH>Ph
e PhCH,Cl oe
(b) ee oe
Cat. TmCl,
Fig. 5.19 Exploiting the Lewis acidity of lanthanides in (a) selective nucleophilic addition and
(b) Friedel-Crafts alkylation.
Heterogeneous catalysts
One major use of lanthanides in heterogeneous catalysis is the stabilisation of
zeolite structures used in petroleum cracking applications. Although the
precise function is somewhat unclear, the addition of small amounts of
lanthanide halides to the zeolite allows it to maintain the high degree of
acidity required, under the harsh conditions present in an industrial reactor.
Another important use is in the exhaust catalysts in cars. This was
discussed in Chapter Four, Section Two.
5.7 Exercises
1. Why is it that carbonate CO,” in natural groundwater can compete
effectively for coordination sites at actinyl centres even though the
excellent ligand H,O is in vast excess?
2. Using Fig. 5.12, predict the equilibrium constant for the formation of
the 1:1 Y**/EDTA* complex in aqueous solution.
Synthesis
In the context of the problems outlined in Section 6.1, the synthesis of f
element alkyls MR, presented a great challenge, but nevertheless this has
been achieved by two general methods.
First, by analogy with the aryloxides and dialkylamides described in
Chapter Five, Section Five, it is possible to synthesise homoleptic
complexes of the f elements such as [U{CH(SiMe,),},] 6.3 using the
extremely bulky bis(trimethylsilyl)methyl ligand.
i ne Me 3-
Vi
Me3Si. H U tte
Saree
SiMe, Me aa
‘lk = Me |
" En
Me
M : H
“SiMe, sn Me
a~\~
| Me
€3Si M
SiMe, Me
Second, in some instances where smaller alkyl groups are used, the metal
centre may accommodate more alkyl ligands than it requires in order to satisfy
its valency. For example, the trivalent lutetium centre in 6.4 has four
2,6-dimethylphenyl groups in a tetrahedral array and thus has an overall
negative charge. Such anionic ate complexes are relatively common in f
element chemistry (see Chapter Five, Section Five). A quite spectacular
example of this behaviour is given by complexes such as_ the
hexamethylerbiate trianion 6.5. Similar compounds are known for most of
the lanthanides and thorium.
oak B-elimination Py es
Ren R
fi eee
Ap 6.7
6.6
The f elements do not form strong bonds with 7 ligands such as alkenes
(Section 6.4) and so the intermediate complex 6.6, if formed at all, will
rapidly lose alkene leaving a hydride species 6.7 which may decompose
further.
70 Organometallics
A
often take part in chemical transformations, and thus they are an example of
what are known as ‘spectator ligands’. Cyclopentadienyls are capable of
forming both covalent and ionic complexes with between one and five carbon
Yo— OBu! atoms bonded to the metal centre. In the n° bonding mode they can be
a 6.8
considered to occupy three coordination sites. For example, we can describe
the trivalent ytterbium complex 6.8 as being seven coordinate. One
important advantage of these ligands is that their steric demand (or how
effective they are at protecting the metal centre) can be altered subtly and
methodically by introduction of alkyl group substituents on the ring. Early
chemistry was based around the simple C;H; ligand known as Cp, but today
the use of very bulky ligands is more common.
The f elements 71
Simple cyclopentadienyls
For the trivalent lanthanides, three general types of cyclopentadienyl
complexes are available, [LnCp,X;,,] where n = 1-3. These compounds are
made by reaction of the anhydrous lanthanide halides with sodium
cyclopentadienyl (Eqn 6.1). Ether solvents such as THF work best, but since
they are themselves powerful ligands to lanthanide and actinide centres, it is
often difficult to remove them subsequently.
LnCl, + n NaCp > [LnCp,Cl,.,] + n NaCl (6.1)
The compounds [LnCp,] react with FeCl, to give ferrocene, as would an
ionic cyclopentadieny! such as NaCp, and this is taken as an indication of the
polar nature of the Cp—Ln bond. This polarity also manifests itself in the fact
that quite rapid exchange of Cp ligands occurs between lanthanide centres.
Nevertheless, compounds containing the Cp,LnCl unit can be isolated as
dimers 6.9 or as Lewis base adducts such as 6.10.
6.11 6.12
The trivalent actinide complexes behave similarly to their lanthanide
analogues, and although the compounds [AnCp,] are well known, complexes
72 Organometallics
such as [AnCpCl,] are often poorly characterised. Many of these problems are
solved by the use of bulky substituents on the cyclopentadienyl ligands, as
we shall see later.
Most of the cyclopentadienyl chemistry of the actinides, however, involves
the metals in a tetravalent state, and complexes [AnCp,Cl,.,,] (n = 3, 4) can be
prepared as in Eqn 6.2. A combination of the high affinity of the actinide ions
for cyclopentadienyl rings and the relatively modest steric demand of the
unsubstituted Cp ligand makes it impossible to stop reaction in Eqn 6.2 at
the [AnCpCl,] and [AnCp,Cl,] (i.e. n = 1, 2) stages. The vast majority of
tetravalent cyclopentadienyls of the actinides are, as expected, of thorium and
uranium, although some protactinium and neptunium complexes are known.
AnCl, + n NaCp > [AnCp,Cl,.,] + n NaCl (6.2)
The fully cyclopentadienylated compound [U(n’-Cp),] 6.14 has been
shown to have the Cp ligands arranged at the vertices of a tetrahedron. This is
the only type of organometallic compound which has four Cp groups all
bonded in the n° fashion. In comparison, the analogous zirconium compound
has the structure [Zr(n°-Cp),(n'-Cp)].
Unlike the lanthanide cyclopentadienyls, tetravalent [UCp,Cl] does not
react with FeCl, to form ferrocene, and this has been taken as an indication of
a significant degree of covalency in the Cp—An(IV) bond. The chloride ligand
in [UCp,Cl] and in the analogous [ThCp,Cl] can be readily substituted
making these compounds among the most important synthons in_ the
organometallic chemistry of the actinides (Fig. 6.3). The three
cyclopentadienyl groups protect the metal centre quite effectively, but at the
same time allow chemistry to take place at the remaining coordination site.
AnCl, [AnCp,(OEt)]
W (i)
Si
[AnCp3(CH3)] [AnCp3(BH,)]
Fig. 6.3 Synthesis and some reactions of [AnCp,Cl]. Reagents (i) 3 TICp, (ii) NaOEt,
(iii) LiMe, and (iv) NaBH,,.
Substituted cyclopentadienyls
Replacement of C5H, (Cp) with bulkier analogues such as C,Me, (Cp*) and
C;H;(SiMe;), (Cp") leads to complexes with more convenient properties.
They are soluble, form crystals more readily, and do not suffer so much from
the aforementioned ligand redistribution reactions. In the same way that the
CH, ligand facilitated the exploration of the organometallic chemistry of the
d block, the use of bulky cyclopentadieny! ligands such as Cp* paved the way
for the f elements.
The f elements 73
6.18 6.17
CK CH LE CHs -CH 2
SX
Be SS L 0)ww H ee Lu— CH,
Wee Se
Fig. 6.5 Hydrocarbon activation via o bond metathesis mediated by a lutetium alkyl 6.19. The methyl group in 6.20 can be thought of
as acting as a strong base which deprotonates the alkane.
Samarium(II)
Lewis base free complexes such as [SmCp,] are expected to be strongly Lewis
acidic, as well as being powerful one electron reducing agents. Unfortunately,
the bis(cyclopentadienyl) complex 6.21 synthesised by the route in Fig. 6.6
incorporates at least two THF ligands into its coordination sphere, and these
are difficult to remove. Use of the bulkier Cp* ligand, however, allows the
removal of coordinated THF from the complex 6.22. The molecular structure
of 6.23 (Fig. 6.7) reveals that it has a bent structure, as do the analogous
europium and ytterbium compounds. The reasons for this are not fully
understood, and similar transition metal complexes (such as ferrocene 6.1)
adopt a parallel structure. As expected, the complex 6.23 is highly reactive,
Fig. 6.7 Molecular structure of as we shall see in Section 6.4.
[SmCp*,] 6.23.
2 NaCp
[SmCp2(THF)o] —«————_—_——-._. Sm
6.21 THF
2NaCp*| THF
C
wae ce LE eee
1 Sm"
4 Ss _e
6.23 6.22
Fig. 6.6 Synthesis of samarium(ll) cyclopentadienyls.
The f elements 75
Thorium(III)
The aqueous chemistry of thorium is confined to the +4 oxidation state since
the trivalent state reduces water rapidly. Even in organometallic chemistry,
only a few compounds containing trivalent thorium have been detected, and
these are exceptionally sensitive to oxygen and water. It was only by the use
of the sterically demanding Cp" ligand that the structure of such a compound
6.24 could be characterised by X-ray crystallography (Fig. 6.8).
SiMe;
wee SiM €3
Na/K alloy pies ak) ”
3 iiiCp Cl) —_{___ 2 +
toluene Me,Si ae
IM@3
6.24
SiMe,
In the same way that the Ti(II) centre in [TiCp;] has the electronic
configuration [Ar]3d' we might expect Th(III) in 6.24 to be [RnJ5f'.
Relativistic calculations on the model complex {[ThCp,], however, show that
the 6d level in this molecule is surprisingly stable and that the electronic
configuration is [Rn]6d'. This is consistent with experimental data in that the
electronic spectrum of the complex consists of a series of strong (i.e. Al = +1
allowed) d — f bands, and the molecule is an intense blue colour.
o donor m™ acceptor
P2205
chopo0 GSEs
Fig. 6.9 The main bonding interactions in metal carbonyl complexes.
It was pointed out in Chapter Two, Section Two that the greater range of
oxidation states exhibited by the early actinides indicates that the valence
electrons are less tightly bound than they are for the lanthanides and later
actinides, and that these electrons may take part in covalent bonding to
ligands. This phenomenon is illustrated by the synthesis under fairly standard
laboratory conditions of CO complexes of trivalent uranium surrounded by
bulky cyclopentadienyl spectator ligands, [UCp';(CO)] 6.25 and
[U(n>-C;Me,H),(CO)] 6.26. Infrared spectroscopy shows that the
wavenumbers of the carbonyl stretch v(CO) for these two compounds are
1976 cm and 1900 cm’ respectively. Since these values are rather lower than
that in free CO (2146 cm'') there must be a significant amount of U-CO n
backbonding. This conclusion is supported by molecular orbital calculations
on the model compound [UCp,(CO)}.
oS
Ss
Me3Si
> SiMe,
6.25 6.26
b- - f- - =Pt(PPh3)3
i Cl
a LY:
ra
Cl Cl
Bu!
< > Bu
Bul 6.31
C49,0"
Filled HOMO
in 18 electron = —-_ 419 9
complexes
= B29, 8
12
jelectrons o and
Fig. 6.11 A section of the molecular orbital energy level diagram for [Sc(y ®-C,H,Bu',).]
showing the orbitals around the HOMO.
The f elements 79
30
ro)
8
oO
Promotion
cm"
10°
energy
x
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Ef Tm Yb
Fig. 6.12 Variation in promotion energy [Xe]4f'6s* — [Xe]4f*'5d'6s? across the lanthanide series. This graph mirrors closely the
variation in third ionization energies (Fig. 2.2).
(c) @oy/f26.2-y2)
Fig. 6.13 Symmetry matched combinations of COT ring and actinide orbitals of 5
symmetry: the major source of bonding in actinidocenes. You might like to think of the metal
f,2,2) in (c) as comprising two d,2.,2 orbitals (a) of opposite phases stacked on top of one-
another. The d,, and f,,, orbitals in (b) and (d) are similarly related.
between the two rings. Hence these are gerade (subscript g) combinations and
have symmetry matches with the d orbitals shown.
655; 0
Se e:
- .
wos
beg os
=.
wee eee eee eee ee ee seven 5f orbitals
0) 0 U
Fig. 6.14 A molecular orbital scheme for D,,, symmetry [U(n®-COT),]: the ligand/metal orbital combinations giving rise to the &
bonding orbitals in the box are shown in Fig. 6.13.
The second most important interaction is between the ligand e,, orbitals
and the uranium 5f. These bonds are depicted in Fig. 6.13(c) and (d). In
addition to the two nodal planes within each COT ring, there is a nodal plane
between the two rings. These ungerade combinations have symmetry matches
with the f orbitals shown. It is important to note that these e,, symmetry 6
82 Organometallics
Other actinidocenes
Compounds analogous to uranocene are known for all the actinides which
|Pa_| 21VE |paramagnetic_| have a stable tetravalent state (i.e. thorium, protactinium, neptunium, and
plutonium). They all appear to have the same structure as uranocene, and their
magnetic properties (Fig. 6.15) suggest that they can be described by the
same bonding model. For example, thoracene would be expected to have the
electronic structure e,,"e,,° (Fig. 6.14) and therefore be diamagnetic, as would
Fig. 6.15 Valence electron (VE)
counts and magnetic properties of
plutonocene (e>,"e3,.).
the actinidocenes [An(n°-COT).,].
Lanthanide COT complexes
Of the lanthanides, only Ce forms a neutral sandwich of this type,
[Ce(n*-COT),]. Cerocene was upheld for many years as a rare example of a
tetravalent cerium organometallic, these species being generally unstable
because of the strong oxidising power of Ce(IV). Recently, however,
theoretical calculations and physical measurements have shown that this
compound is better described as having a trivalent metal Ce centre [Xe]4f!
complexed by two COT'®* ions (the e,, 5 bonding orbitals in Fig. 6.14
contain only three electrons). In ionic notation we might write
(Ce $COT.}"}:
Neutral cerocene is readily reduced, for example by potassium metal, to
give what is undoubtedly a trivalent cerium compound (Eqn 6.4). This type of
sandwich anion can be prepared for most of the lanthanides directly by the
route in Eqn 6.5.
6.7 Exercises
1. Why are agostic interactions difficult to detect in f element
organometallic complexes?
Further reading
Cotton, S.A. (1991). Lanthanides and actinides. MacMillan Education,
London, UK.
Cox, P.A. (1987). The electronic structure and chemistry of solids. Oxford
University Press, Oxford, UK.
Kaltsoyannis, N. (1997). Relativistic effects in inorganic and organometallic
chemistry, Journal of the Chemical Society, Dalton Transactions, 1—
ily
Katz, J.J., Seaborg, G.T. and Morss, L.R. (eds) (1986). The chemistry of the
actinide elements, (2nd edition). Chapman and Hall, London, UK.
Kettle, S.F.A. (1996). Physical inorganic chemistry: a_ coordination
chemistry approach. Spektrum, Oxford, UK.
Index
SERIES EDITORS
This series of short texts provides accessible accounts of a range of essential topics
in chemistry and chemical engineering. Written with the needs of the student in
mind, the Oxford Chemistry Primers offer just the right level of detail for
undergraduate study, and will be invaluable as a source of material commonly
presented in lecture courses yet not adequately covered in existing texts. All the
basic principles and facts in a particular area are presented in a clear and
straightforward style, to produce concise yet comprehensive accounts of topics
covered in both core and specialist courses.
The felements is a short accessible text which covers much of the fundamental
chemistry of the lanthanide and actinide elements, including coordination
chemistry, solid state compounds, organometallic chemistry, electronic
spectroscopy, and magnetism. Comparisons are made between the chemistry of
the lanthanides and actinides and that of the transition elements, which is generally
much more familiar to undergraduate chemistry students.
ISBN 0-19-850467-5
IO
———