You are on page 1of 124

This E Book is not for Sale

Complied as a study Aid for CSJM University


students , who are unable to buy expensive textbooks

ADVANCED INORGANIC CHEMISTRY


FOR POST GRADUATE STUDENTS
PART ONE
ISBN:978-1-365-43616-1
---------------------------------------------------------------
COMPLIED BY

Dr. Miss. Jyotsna Lal


Associate Prof Chemistry Department
Christ Church College Kanpur . U.P . INDIA

CONTENTS
PART 1
1] CLUSTERS

2] POLY ACIDS

3] METAL PI COMPLEXES

PART 2
4] METAL LIGAND BONDING

5]REACTION MECHANISM OF TRANSITION METAL COMPLEXES

6] METAL LIGAND EQUILIBRA IN SOLUTION

7] ELECTRONIC SPECTRA AND MAGNETIC PROPERTIES OF


TRANSITION METAL COMPLEXES

8] STEREOCHEMISTRY AND BONDING OF MAIN GROUP COMPOUNDS

REFERENCES ARE AT THE END , SO THAT


THESE FACTS CAN BE VERIFIED
MAY ALSO LEAD TO FURTHER READING

CHAPTER 1: CLUSTERS

This Ebook is not for sale


Complied as a study Aid for CSJM Kanpur University
students Who are unable to buy expensive textbooks.

CHAPTER 1
The topic Clusters encompases all giant molecules
be it cagelike Boranes, Carboranes,
Polycentred Carbonyls, Nitrosyls , Phosphines ,
Di Oxygen and Di Nitrogen Complexes
Water molecules , Large Gas molecules
Poly Acids and Hetero Poly Acids
Carbon structures
Interesting Reading for a young budding Chemist
The structures in modern inorganic chemistry
are unique and so very different .
As a Teacher one is challanged to help the students
gain proper understanding
TOPIC 1:1 DEFINATION

In chemistry, a cluster is an ensemble of bound atoms or


molecules that is intermediate in size between a molecule
and a bulk solid. Clusters exist of diverse stoichiometries
and nuclearities. For example, carbon and boron atoms form
fullerene and borane clusters, respectively.
Transition metals and main group elements form especially
robust clusters.Clusters can also consist solely of a certain
kind of molecules, such as water clusters.

The phrase cluster was coined by F.A. Cotton in the early


1960s to refer to compounds containing metal–metal bonds.
In another definition a cluster compound contains a group of
two or more metal atoms where direct and substantial metal
bonding is present.The prefixed terms "nuclear" and
"metallic" are used and imply different meanings. For
example, polynuclear refers to a cluster with more than one
metal atom, regardless of the elemental identities.
Heteronuclear refers to a cluster with at least two different
metal elements.

The main cluster types are "naked" clusters (without


stabilizing ligands) and those with ligands. For transition
metal clusters, typical stabilizing ligands include carbon
monoxide, halides, isocyanides, alkenes, and hydrides. For
main group elements, typical clusters are stabilized by
hydride ligands.

Transition metal clusters are frequently composed of


refractory metal atoms. In general metal centers with
extended d-orbitals form stable clusters because of
favorable overlap of valence orbitals.
Thus, metals with a low oxidation state for the later metals
and mid-oxidation states for the early metals tend to form
stable clusters.
Polynuclear metal carbonyls are generally found in late
transition metals with low formal oxidation states.
The polyhedral skeletal electron pair theory or Wade's
electron counting rules predict trends in the stability and
structures of many metal clusters. Jemmis mno rules have
provided additional insight into the relative stability of metal
clusters.

1.a] Atomic clusters


1.b]Molecular clusters
1.c]Transition metal carbonyl clusters
1.d] Transition metal organic carbon (organometallic)
clusters
1.e]Transition metal halide clusters
1.f] Boron hydrides
1. g]Fe-S clusters in biology
1.h] Zintl clusters
1.i]Metalloid clusters
1.j] Intermetalloid clusters
1.k] Gas-phase clusters and fullerenes
1.n] Extended metal atom chains

History and classification

The development of cluster chemistry occurred


contemporaneously along several independent lines, which
are roughly classified in the following sections. The first
synthetic metal cluster was probably calomel, which was
known in India already in the 12th century. The existence of
a mercury to mercury bond in this compound was
established in the beginning of the 20th century.
Atomic clusters

Atomic clusters can be either pure, formed from a single


atomic species, or mixed, formed from a mixed atomic
species. Classifications criteria:
by predominant bond nature: metallic, covalent, ionic;
by atomic count: "micro" < 20, "small" < 100 < "large"
by electric / magnetic properties

For the majority of atomic species there are clusters of


certain atomic counts (so-called "magic numbers") that have
preponderent representation in the mass spectra, an
indication of their greater stability with respect to dissociation
when compared to their neighboring atomic counts.

Molecular clusters
Atomic and molecular clusters are aggregates of 5-105
atomic or molecular units. They are classified according to
the forces holding them together:

Van der Waals clusters - attraction between induced electric


dipoles and repulsion between electron cores of closed
electronic configurations

Metallic clusters - long range valence electron sharing (over


many successive adjacent atoms) and partially directional

Ionic clusters - valence electrons are almost entirely


transferred among closest neighbors to yield 2 net, equal but
opposite, electric charge distributions that mutually attract.
Ralf Ludwig (2001). "Water: From Clusters to the Bulk". Angew. Chem.
Int. Ed. 40 (10): 1808–1827. Hypothetical (H2O)100 icosahedral water
cluster and the underlying structure.

FIGURE -SHOWS A WATER CLUSTER


The role of cluster formation in the precipitation of liquid
mixtures and in the condensation, adsorption to surface or
solidification phase transitions has long been investigated
from a theoretical standpoint.

Cluster system properties — stem both from their size and


composition (which contributes to the binding force types)
that determine:

the number of dimensions of their phase space


the ranges of accessible positions and velocities of their
atomic components

A gradual transition occurs between the properties of the


molecular species and those of the corresponding bulk mix.
And yet the clusters exhibit physical and chemical properties
specific only to their configuration space (in turn strongly
atom-count-dependent) and not specific to their bulk
counterparts.

Cluster systems are metastable with respect to at least one


of the following evolution classes:

Atom elimination or adsorption at cluster surface as a cause


for their disassociation or growth
configuration switches among a set of stable structures
(a.k.a. an "isomer class") accessible to all clusters of a same
atom count and a same relative component composition.

Many of their properties are due to the fact that a large


fraction of their component atoms is found at their surface.
With increasing size, the relative number of atoms at the
cluster surface will scale approximately as N−1/3. One has
to reach beyond a variable threshold of 9-27 component
molecules (depending on the strength of the inter-molecular
forces) to find global minimum configurations that hold at
least one interior molecule. At the other end of the scale a
cluster of about 105 atoms will expose only about 10% of the
atoms at its surface, a still significant percentage in
comparison to the bulk solid.
In chemistry a water cluster is a discrete hydrogen bonded
assembly or cluster of molecules of water.These clusters
have been found experimentally or predicted in silico in
various forms of water; in ice, in crystal lattices and in bulk
liquid water, the simplest one being the water dimer (H2O)2 .
Ongoing academic research is important because the
realization that water manifests itself as clusters rather than
an isotropic collection may help explain many anomalous
water characteristics such as its highly unusual density
temperature dependence. Water clusters are also implicated
in the stabilization of certain supramolecular structures. So
little is understood about water clusters in bulk water that it is
considered one of the unsolved problems in chemistry.
Even larger clusters are predicted: the fullerene-like cluster
(H2O)28 is called the water buckyball and even for a 280
water molecule monster icosahedral network (with each
water molecule coordinate to 4 others) there is found a local
energy minimum. The 280 molecule icosahedral structure,
which is 3 nm in diameter, consists of icosahedral shells with
280, 100 and 20 molecules (the 100 molecule structure is
shown the fig

According to the so-called in silico method quantum cluster


equilibrium (QCE) theory of liquids W8[clarification needed]
clusters dominate the liquid water bulk phase followed by
W5 and W6 clusters. In order to facilitate a water triple point
the presence of a W24 cluster is invoked. In another model
bulk water is built up from a mixture of hexamer and
pentamer rings containing cavities capable of enclosing
small solutes. In yet another model an equilibrium exists
between a cubic water octamer and two cyclic tetramers.
However, in spite of much model-making, no model yet has
reproduced the experimentally-observed density maximum
Transition metal organic carbon (organometallic) clusters

There are also organometallic clusters. These clusters


contain metal-metal bonds as well as at least a single
organic moiety directly bonded to a metal atom (an organic
ligand). One of the first examples of this type of clusters is a
tricobalt nonacarbonyl cluster (its structure was determined
by Dahl). This cluster contains a methyl carbyne ligand
(CCH3) bonded to three cobalt atoms:
[Co3(CCH3)(CO)9]. There are also more derivatives of the
tricobalt methyl carbyne cluster where other organic ligands
than methyl carbyne are introduced into the cluster structure
(mainly by the work of Seyferth et al.).
The above-mentioned cluster serves as an example of an
overall zero-charged (neutral) cluster. In addition, cationic
(positively charged) rather than neutral organometallic
trimolybdenumor tritungsten clusters are also known. The
first representative of these ionic organometallic clusters is
[Mo3(CCH3)2(O2CCH3)6(H2O)3]2+ (first prepared and
characterized by Bino). The former tricobalt clusters are
soluble in non-polar solvents, whereas the latter
trimolybdenum and tritungsten ones are soluble in polar
solvents including water.
Transition metal halide clusters

Linus Pauling showed that "MoCl2" consisted of Mo6


octahedra. F. Albert Cotton established that "ReCl3" in fact
features subunits of the cluster Re3Cl9, which could be
converted to a host of adducts without breaking the Re-Re
bonds. Because this compound is diamagnetic and not
paramagnetic the rhenium bonds are double bonds and not
single bonds. In the solid state further bridging occurs
between neighbours and when this compound is dissolved in
hydrochloric acid a Re3Cl123− complex forms. An example
of a tetranuclear complex is hexadecamethoxytetratungsten
W4(OCH3)12 with tungsten single bonds and molybdenum
chloride ((Mo6Cl8)Cl4) is a hexanuclear molybdenum
compound and an example of an octahedral cluster. A
related group of clusters with the general formula MxMo6X8
such as PbMo6S8 form a Chevrel phase, which exhibit
superconductivity at low temperatures. The eclipsed
structure of potassium octachlorodirhenate(III), K2Re2Cl8
was explained by invoking Quadruple bonding. This
discovery led to a broad range of derivatives including di-
tungsten tetra(hpp), the current (2007) record holder low
ionization energy.
Structure of Mo6Cl142−.
Boron hydrides

Contemporaneously with the development of metal cluster


compounds, numerous boron hydrides were discovered by
Alfred Stock and his successors who popularized the use of
vacuum-lines for the manipulation of these often volatile, air-
sensitive materials. Clusters of boron are boranes such as
pentaborane and decaborane. Composite clusters
containing CH and BH vertices are carboranes.
B12H122− is a famous boron hydride cluster

Fe-S clusters in biology

In the 1970s, ferredoxin was demonstrated to contain Fe4S4


clusters and later nitrogenase was shown to contain a
distinctive MoFe7S9 active site.[9] The Fe-S clusters mainly
serve as redox cofactors, but some have a catalytic function.
In the area of bioinorganic chemistry, a variety of Fe-S
clusters have also been identified that have CO as ligands.

Zintl clusters

Zintl compounds feature naked anionic clusters that are


generated by reduction of heavy main group p elements,
mostly metals or semimetals, with alkali metals, often as a
solution in anhydrous liquid ammonia or
ethylenediamine.[10] Examples of Zintl anions are [Bi3]3−,
[Sn9]4−, [Pb9]4−, and [Sb7]3−.[10] Although these species
are called "naked clusters," they are usually strongly
associated with alkali metal cations. Some examples have
been isolated using cryptate complexes of the alkali metal
cation, e.g., [Pb10]2− anion, which features a capped square
antiprismatic shape.[11] According to Wade's rules (2n+2)
the number of cluster electrons is 22 and therefore a closo
cluster. The compound is prepared from oxidation of K4Pb9
[12] by Au+ in PPh3AuCl (by reaction of tetrachloroauric acid
and triphenylphosphine) in ethylene diamine with 2.2.2-crypt.
This type of cluster was already known as is the endohedral
Ni@Pb102− (the cage contains one nickel atom). The
icosahedral tin cluster Sn122− or stannaspherene anion is
another closed shell structure observed (but not isolated)
with photoelectron spectroscopy.[13][14] With an internal
diameter of 6.1 Angstrom it is of comparable size to fullerene
and should be capable of containing small atoms in the
same manner as endohedral fullerenes and indeed exists a
Sn12 cluster that contains an Ir atom: [Ir@Sn12]3−.
Metalloid clusters

Elementoid clusters are ligand-stabilized clusters of metal


atoms that possess more direct element-element than
element-ligand contacts. Examples of structurally
characterized clusters feature ligand stabilized cores of Al77,
Ga84, and Pd145.
Intermetalloid clusters

These clusters consist of at least two different (semi)metallic


elements and possess more direct metal-metal than metal-
ligand contacts. The suffix "oid" designate that such clusters
possess at a molecular scale, atom arrangements that
appear in bulk intermetallic compounds with high
coordination numbers of the atoms such as for example in
Laves phase and Hume-Rothery phases.
Ligand-free intermetalloid clusters include also endohedrally
filled Zintl clusters.
A synonym for ligand-stabilized intermetalloid clusters is
"molecular alloy". The clusters appear as discrete units in
intermetallic compounds separated from each other by
electropositive atoms such as [Sn@Cu12@Sn20]12−,
as soluble ions [As@Ni12@As20]3−
or as ligand-stabilized molecules such as
[Mo(ZnCH3)9(ZnCp*)3].

Gas-phase clusters and fullerenes

Unstable clusters can also be observed in the gas-phase by


means of mass spectrometry even though they may be
thermodynamically unstable and aggregate easily upon
condensation.

Such naked clusters, i.e. those that are not stabilized by


ligands, are often produced by laser induced evaporation - or
ablation - of a bulk metal or metal-containing compound.
Typically, this approach produces a broad distribution of size
distributions. Their electronic structures can be interrogated
by techniques such as photoelectron spectroscopy, while
infrared multiphoton dissociation spectroscopy is more
probing the clusters geometry.
Their properties (Reactivity, Ionization potential, HOMO–
LUMO-gap)
often show a pronounced size dependence.

Examples of such clusters are certain aluminium clusters as


superatoms and certain gold clusters. Certain metal clusters
are considered to exhibit metal aromaticity. In some cases,
the results of laser ablation experiments are translated to
isolated compounds, and the premier cases are the clusters
of carbon called the fullerenes, notably clusters with the
formula C60, C70, and C84. The fullerene sphere can be
filled with small molecules, forming Endohedral fullerenes.
Extended metal atom chains

Extended metal atom chain complexes (EMAC) are a novel


topic in academic research. An EMAC is composed of linear
chains of metal atoms stabilized with ligands. EMACs are
known based on nickel (with 9 atoms),
chromium and cobalt (7 atoms) and ruthenium (5 atoms). In
theory it should be possible to obtain infinite one-
dimensional molecules and research is oriented towards this
goal. In one study [21] an EMAC was obtained that
consisted of 9 chromium atoms in a linear array with 4
ligands (based on an oligo pyridine) wrapped around it. In it
the chromium chain contains 4 quadruple bonds.
Metal clusters in catalysis

Although few metal carbonyl clusters are catalytically useful,


naturally occurring Iron-sulfur proteins catalyse a variety of
transformations such as the stereo-specific isomerization of
citrate to isocitrate via cis-aconitate, as required by the
tricarboxylic acid cycle. Nitrogen is reduced to ammonia at
an Fe-Mo-S cluster at the heart of the enzyme nitrogenase.
CO is oxidized to CO2 by the Fe-Ni-S cluster carbon
monoxide dehydrogenase. Hydrogenases rely on Fe2 and
NiFe clusters.
Isoprenoid biosynthesis, at least in certain organisms,
requires Fe-S clusters.
Catalysis by metal carbonyl clusters

Metal carbonyl cluster compounds have been evaluated as


catalysts for a wide range of reactions, especially for
conversions of carbon monoxide.No industrial applications
exist however. The clusters Ru3(CO)12 and Ir4(CO)12
catalyze the Water gas shift reaction, also catalyzed by iron
oxide, and Rh6(CO)16 catalyzes the conversion of carbon
monoxide into hydrocarbons, reminiscent of the Fischer-
Tropsch process, also catalyzed by simple iron compounds.

Some define cluster catalysis to include clusters that have


only one active site on one metal atom. The definition can be
further relaxed to include clusters that remain intact during at
least one reaction step, and can be fragmented in all others.
METAL Pi COMPLEXES

1:1 CARBORANES

A carborane is a cluster composed of boron, carbon and


hydrogen atoms. Like many of the related boranes, these
clusters are polyhedra and are similarly classified as closo-,
nido-, arachno-, hypho-, etc. based on whether they
represent a complete (closo-) polyhedron, or a polyhedron
that is missing one (nido-), two (arachno-), or more vertices.
Interesting examples of carboranes are the extremely stable
icosahedral closo-carboranes.These boron-rich clusters
exhibit unique organomimetic properties with chemical
reactivity matching classical organic molecules, yet
structurally similar to metal-based inorganic and
organometallic species

A prominent example is the charge-neutral C2B10H12 or o-


carborane with the prefix o derived from ortho, which has
been explored for use in a wide range of applications from
heat-resistant polymers to medical applications
Boron clusters with more than 12 vertices showing a very
rich and diverse chemistry were confined to the 13- and 14-
vertex metallacarboranes until 2003. Very recently,
significant progress in the syntheses of 13- and 14-vertex
carboranes has been made, leading to the preparation of 15-
vertex metallacarboranes. These studies open up new
possibilities for the development of polyhedral clusters of
extraordinary size.
The most heavily studied carborane is C2B10H12, m. p. 320
°C. It is often prepared from the reaction of acetylene with
decaborane. A variation on this method entails the use of
dimethyl acetylenedicarboxylate to give C2B10H10(CO2C
H3)2, which can be degraded to the C2B10H12.
Ortho dicarborane is the kinetic product from the addition of
acetylenes to decarborane precursors. Upon heating at 420
°C, the ortho dicarborane rearranges to the meta isomer.
Upon further heating, one obtains para-carborane. Like
arenes, carboranes also undergo electrophilic aromatic
substitution.
This anion forms sandwich compounds, referred to as
bis(dicarbollides), with many metal ions and some exist in
otherwise unusual oxidation states. The dianion is a nido
cluster prepared by degradation of the parent dicarborane:[6]

B10C2H12 + 3 CH3OH + KOH → KB9C2H12 +


B(OCH3)3 + H2O + H2

Bis(dicarbollides) often exhibit properties very different from


their metallocene surrogates. For example, Ni-based
bis(dicarbollide) cluster can be observed for the rare Ni(IV)
oxidation state of nickel. Some notable examples of potential
applications of these complexes include catalysis,[7] ion-
exchange materials for radioactive waste management,
biologically active protease inhibitors, and chemically inert
redox shuttles for dye-sensitized solar cells (DSSCs)

Carboryne, or 1,2-dehydro-o-carborane, is an unstable


derivative of ortho-carborane with the formula B10C2H10.
The hydrogen atoms on the C2 unit in the parent o-
carborane are missing. The compound resembles and is
isolobal with benzyne. A carboryne compound was first
generated in 1990 starting from o-carborane. The hydrogen
atoms connected to carbon are removed by n-butyllithium in
tetrahydrofuran and the resulting lithium dianion is reacted
with bromine at 0 °C to form the bromo monoanion.

Carborynes react with alkynes to benzocarboranes in an


adaptation of the above described procedure. O-carborane
is deprotonated with n-butyllithium as before and then
reacted with dichloro-di(triphenylphosphino) nickel to a nickel
coordinated carboryne. This compound reacts with 3-hexyne
in an alkyne trimerization to the benzocarborane.

Carboranes have been explored as a source of boron in


boron neutron capture therapy.They have also been
examined in structural studies in crystallography.
Carboranes have been used to make solid superacids. Solid
superacid catalysts alleviate the need to dispose of spent
acids, thus providing a significant environmental advantage
over dissolved acids.The carborane superacid
H(CHB11Cl11)is one million times stronger than sulfuric
acid. The reason for this high acidity is that the acid anion
CHB11Cl11− is very stable and substituted with
electronegative substituents. H(CHB11Cl11) is the only acid
known to protonate C60 fullerene without decomposing it.
Additionally, it is the only known anion capable of forming a
stable, isolable salt with protonated benzene, C6H7+. In
coordination chemistry carboranes can be used as unique
bulky ligand scaffolds. It was recently demonstrated, that the
same carboranyl moiety can act either as strongly electron-
withdrawing or electron-donating substituent, depending on
the positional attachment of the cluster to the heteroatom
Coordination chemistry of carborane anions
Investigating the reactions of the neutral nidocarborane
fragment C2B9H13 and its anions C2B9H12- and
C2B9H112- with metal amides, alkyls and chlorides.
Scientists are seeking to explore the synthesis and reactivity
of metallacarboranes of the early transition metals in their
highest oxidation states, with an emphasis on how the
carborane ligand influences the chemistry of the metal
fragment,exploring the reactions of carborane reagents with
some other metal containing reagents, and have made some
interesting observations, e.g., in zinc chemistry
Structures of salts of C2B9H12- and C2B9H112-

The 11 vertex carbon and boron hydrides C2B9H13 and


their mono- and di-anions C2B9H12- and C2B9H112-
obtained by deprotonation, were first described by
Hawthorne and co-workers in the 1960s. These are the most
common precursors used by our group and a great many
others to synthesise metallacarboranes and a wide range of
heterocarboranes. The structure of all three isomers of
C2B9H112- consists of an icosahedron with one vertex
vacant, and with an exo-hydrogen atom on each of the 11
vertices; the isomers differ in the location of the carbon
atoms. The mono-anions are a tiny bit more challenging
structurally, and there is a facinating history of the structure
proposed for the most widely used 7,8- C2B9H12- anion.
The location of the endo- or 12th hydrogen atom appears to
depend on the identity of the counter-ion, and we find that
the structure of carborane anion as the proton-sponge salt
differs from that reported by other workers for other counter-
ions.

The structure of (PSH+)(7,8-C2B9H12-) determined by


neutron diffraction at natural isotopic abundance.
The 10B isotope of boron is a potent absorber of neutrons,
and previous neutron diffraction studies of borane clusters
have utilised expensive 10B-depleted materials, but the
intense beam of the ISIS spallation source made this study
possible.

The molecular structure of (PSH+)(7,8-C2B9H12-)


determined by neutron diffraction (PS = proton sponge, 1,8-
bis(dimethylamino)naphthalene); Inorganic Chemistry, 2001,
40, 173.

Zinc metallacarboranes
The reaction of ZnMe2 with [nido-7,8-C2B9H12][NMe3H] in
a 1:1 ratio, gives an unusual macropolyhedral
metallacarborane, which is a novel dimer of the expected
[closo-NMe3-Zn-C2B9H11] product. The structure of the
product bears a passing resemblance to a pair of ear-muffs
or headphones.
Structures of salts of C2B9H12- and C2B9H112-
The 11 vertex carbon and boron hydrides C2B9H13 and
their mono- and di-anions C2B9H12- and C2B9H112-
obtained by deprotonation, were first described by
Hawthorne and co-workers in the 1960s. These are the most
common precursors used by our group and a great many
others to synthesise metallacarboranes and a wide range of
heterocarboranes. The structure of all three isomers of
C2B9H112- consists of an icosahedron with one vertex
vacant, and with an exo-hydrogen atom on each of the 11
vertices; the isomers differ in the location of the carbon
atoms. The mono-anions are a tiny bit more challenging
structurally, and there is a facinating history of the structure
proposed for the most widely used 7,8- C2B9H12- anion.
The location of the endo- or 12th hydrogen atom appears to
depend on the identity of the counter-ion, and the structure
of carborane anion as the proton-sponge salt differs from
that reported by other workers for other counter-ions.
The structure of (PSH+)(7,8-C2B9H12-) determined by
neutron diffraction at natural isotopic abundance.
The 10B isotope of boron is a potent absorber of neutrons,
and previous neutron diffraction studies of borane clusters
have utilised expensive 10B-depleted materials, but the
intense beam of the ISIS spallation source made this study
possible.

The molecular structure of (PSH+)(7,8-C2B9H12-)


determined by neutron diffraction (PS = proton sponge, 1,8-
bis(dimethylamino)naphthalene);
Reference Inorganic Chemistry, 2001, 40, 173.

unprecedented electron deficient bridging between zinc


atoms by boron atoms of nido-carborane anions:
preparation, crystal and molecular structure of the dimer
[(nido-C2B9H11)ZnNMe3]2;
Reference J. Chem. Soc., Chem. Commun., 1998, 1713.

Group 5 (niobium and tantalum) metallacarboranes


The carborane starting material nido-7,8-C2B9H13 reacts
with metal amides, e.g., M(NMe2)5, to liberate two
equivalents of amine and coordinate the dicarbollide dianion
to the resulting metal fragment, giving (C2B9H11)M(NMe2)3
(M = Nb or Ta). Most of the exploratory work here has been
carried out using exclusively tantalum, only a select few
examples of niobium have been studied - the chemistry is
essentially identical.
Group 6 (tungsten) metallacarboranes
The metallacarborane chemistry of tungsten is in fact well
established, and a large contribution to the chemical
literature has already been made in this area. Most of the
known group 6 metallacarboranes are from the group
[ Reference Prof. F. G. A. Stone (at Bristol and then Baylor,
Texas), and contain molybdenum or tungsten in low formal
oxidation states, typically M(II), with p-acceptor spectator
ligands, such as carbonyl. By contrast we have been
exploring the synthesis and reactivity of tungsten
metallacarboranes supported by imido (RN) ligands (with
which (C2B9H11 is related, both being s2,p4 ligands).
The work has involved exploring the balance between the p-
donor abilities of the imido (RN), amido (RHN) and
carborane (C2B9H11) ligands.
In common with work on zinc, niobium and tantaum
metallacarboranes, preferred means of coordinating a metal
and carborane fragment is alkane or amine elimination
between an acidic carborane precursor (usually
Me3NH+nido-7,8-C2B9H12-) and a metal alkyl or amide.
A readily accessible tungsten precursor is the bis(imido)
bis(amide), W(NtBu)2(NHtBu)2, prepared from WCl6 and
tBuNH2 in high yield by a one-pot reaction, and this provides
an entry into the chemistry [This work is being performed in
collaboration with Dr. Mark Fox.]
METAL Pi COMPLEXES

SECTION 1: 2 METAL CARBONYL


Metal carbonyls are coordination complexes of transition
metals with carbon monoxide ligands. Metal carbonyls are
useful in organic synthesis and as catalysts or catalyst
precursors in homogeneous catalysis, such as
hydroformylation and Reppe chemistry. In the Mond
process, nickel carbonyl is used to produce pure nickel. In
organometallic chemistry, metal carbonyls serve as
precursors for the preparation of other organometalic
complexes.
Metal carbonyls are coordination complexes of transition
metals with carbon monoxide ligandsThe carbon monoxide
ligand may be bound terminally to a single metal atom or
bridging to two or more metal atoms. These complexes may
be homoleptic, that is containing only CO ligands, such as
nickel carbonyl (Ni(CO)4), but more commonly metal
carbonyls are heteroleptic and contain a mixture of
ligands.binary carbonyls, i.e. species of the formula
[Mx(CO)n]z,

Nomenclature and terminology

The nomenclature of the metal carbonyls depends on the


charge of the complex, the number and type of central
atoms, and the number and type of ligands and their binding
modes. They occur as neutral complexes, as positively
charged metal carbonyl cations or as negatively charged
metal carbonylates. The carbon monoxide ligand may be
bound terminally to a single metal atom or bridging to two or
more metal atoms. These complexes may be homoleptic,
that is containing only CO ligands, such as nickel carbonyl
(Ni(CO)4), but more commonly metal carbonyls are
heteroleptic and contain a mixture of ligands.

Mononuclear metal carbonyls contain only one metal atom


as the central atom. Except Vanadium hexacarbonyl only
metals with even order number such as chromium, iron,
nickel and their homologs build neutral mononuclear
complexes. Polynuclear metal carbonyls are formed from
metals with odd order numbers and contain a metal-metal
bond. Complexes with different metals, but only one type of
ligand will be referred to as isoleptic.
The number of carbon monoxide ligands in a metal carbonyl
complex is described by a Greek numeral, followed by the
word carbonyl. Carbon monoxide has different binding
modes in metal carbonyls. They differ in the hapticity and the
bridging mode. The hapticity describes the number of carbon
monoxide atoms, which are directly bonded to the central
atom. The denomination shall be made by the letter ηn,
which is prefixed to the name of the complex. The
superscript n indicates the number of bounded atoms. In
monohapto coordination, such as in terminally bonded
carbon monoxide, the hapticity is 1 and it is usually not
separately designated. If carbon monoxide is bound via the
carbon atom and via the oxygen to the metal, it will be
referred to as dihapto coordinated η2

The carbonyl ligand engages in a range of bonding modes in


metal carbonyl dimers and clusters. In the most common
bridging mode, the CO ligand bridges a pair of metals. This
bonding mode is observed in the commonly available metal
carbonyls: Co2(CO)8, Fe2(CO)9, Fe3(CO)12, and
Co4(CO)12.In certain higher nuclearity clusters, CO bridges
between three or even four metals. These ligands are
denoted μ3-CO and μ4-CO. Less common are bonding
modes in which both C and O bond to the metal, e.g. μ3-η2.
CObondingmodes2.png

Structure and bonding


The HOMO of CO is a σ MO
Energy level scheme of the σ and π orbitals of carbon
monoxide
The LUMO of CO is a π* antibonding MO

Carbon monoxide bonds to transition metals using


"synergistic π* back-bonding." The bonding has
three components, giving rise to a partial triple bond. A
sigma bond arises from overlap of nonbonding sp-hybridized
electron pair on carbon with a blend of d-, s-, and p-orbitals
on the metal. A pair of π bonds arises from overlap of filled
d-orbitals on the metal with a pair of π-antibonding orbitals
projecting from the carbon of the CO. The latter kind of
binding requires that the metal have d-electrons, and that the
metal is in a relatively low oxidation state (<+2) which makes
the back donation process favorable. As electrons from the
metal fill the π-antibonding orbital of CO, they weaken the
carbon-oxygen bond compared with free carbon monoxide,
while the metal-carbon bond is strengthened. Because of the
multiple bond character of the M-CO linkage, the distance
between the metal and carbon is relatively short, often < 1.8
Â, about 0.2 Â shorter than a metal-alkyl bond. Several
canonical forms can be drawn to describe the approximate
metal carbonyl bonding modes.
Resonance structures of a metal carbonyl, from left to right
the contributions of the right-hand-side canonical forms
increase as the back bonding power of M to CO increases.

Infrared spectroscopy is a sensitive probe for the presence


of bridging carbonyl ligands. For compounds with doubly
bridging CO ligands, denoted μ2-CO or often just μ-CO,
νCO, νCO is usually shifted by 100–200 cm−1 to lower
energy compared to the signatures of terminal CO, i.e. in the
region 1800 cm−1. Bands for face capping (μ3) CO ligands
appear at even lower energies. Typical values for rhodium
cluster carbonyls are:[5] In addition to symmetrical bridging
modes, CO can be found bridge unsymmetrically or through
donation from a metal d orbital to the π* orbital of
CO.[6][7][8] The increased π-bonding due to back-donation
from multiple metal centers results in further weakening of
the C-O bond.

Physical characteristics
Most mononuclear carbonyl complexes are colorless or pale
yellow volatile liquids or solids that are flammable and
toxic.Vanadium hexacarbonyl, a uniquely stable 17-electron
metal carbonyl, is a blue-black solid.Di- and polymetallic
carbonyls tend to be more deeply colored. Triiron
dodecacarbonyl (Fe3(CO)12) forms deep green crystals.
The crystalline metal carbonyls often are sublimable in
vacuum, although this process is often accompanied by
degradation. Metal carbonyls are soluble in nonpolar and
polar organic solvents such as benzene, diethyl ether,
acetone, glacial acetic acid and carbon tetrachloride. Some
salts of cationic and anionic metal carbonyls are soluble in
water or lower alcohols.
Analytical characterization
Isomers of dicobalt octacarbonyl

Important analytical techniques for the characterization of


metal carbonyls are infrared spectroscopy and 13C NMR
spectroscopy. These two techniques provide structural
information on two very different time scales. Infrared active
vibrational modes, such as CO-stretching vibrations are
often fast compared to intramolecular processes, whereas
NMR transitions occur at lower frequencies and thus sample
structures on a time scale that, it turns out, is comparable to
the rate of intramolecular ligand exchange processes. NMR
data provides information on "time-averaged structures,"
whereas IR is an instant "snapshot."
Illustrative of the differing time scales, investigation of
dicobalt octacarbonyl (Co2(CO)8) by means of infrared
spectroscopy provides 13 νCO bands, far more than
expected for a single compound. This complexity reflects the
presence of isomers with and without bridging CO-ligands.
The 13C-NMR spectrum of the same substance exhibits only
a single signal at a chemical shift of 204 ppm. This simplicity
indicates that the isomers quickly interconvert.
The Berry pseudorotation mechanism for iron pentacarbonyl

Iron pentacarbonyl exhibits only a single 13C-NMR signal


owing to rapid exchange of the axial and equatorial CO
ligands by Berry pseudorotation.
Infrared spectra
The number of IR-active vibrational modes of several
prototypical metal carbonyl complexes.

The most important technique for characterizing metal


carbonyls is infra-red spectroscopy. The C-O vibration,
typically denoted νCO, occurs at 2143 cm−1 for CO gas. The
energies of the νCO band for the metal carbonyls correlates
with the strength of the carbon-oxygen bond, and inversely
correlated with the strength of the π-backbonding between
the metal and the carbon. The π basicity of the metal center
depends on a lot of factors; in the isoelectronic series (Ti to
Fe) at the bottom of this section, the hexacarbonyls show
decreasing π-backbonding as one increases (makes more
positive) the charge on the metal. π-Basic ligands increase
π-electron density at the metal, and improved backbonding
reduces νCO. The Tolman electronic parameter uses the
Ni(CO)3 fragment to order ligands by their π-donating
abilities.

The number of vibrational modes of a metal carbonyl


complex can be determined by group theory. Only vibrational
modes that transform as the electric dipole operator will have
non-zero direct products and are observed. The number of
observable IR transitions (but not the energies) can thus be
predicted.For example, the CO ligands of octahedral
complexes, e.g. Cr(CO)6, transform as a1g, eg, and t1u, but
only the t1u mode (anti-symmetric stretch of the apical
carbonyl ligands) is IR-allowed. Thus, only a single νCO
band is observed in the IR spectra of the octahedral metal
hexacarbonyls. Spectra for complexes of lower symmetry
are more complex. For example, the IR spectrum of
Fe2(CO)9 displays CO bands at 2082, 2019, 1829 cm−1.
The number of IR-observable vibrational modes for some
metal carbonyls are shown in the table. Exhaustive
tabulations are available.
Compound νCO (cm−1) 13C NMR shift
CO 2143 181
Ti(CO)6−2 1748
V(CO)6−1 1859
Cr(CO)6 2000 212
Mn(CO)6+ 2100
Fe(CO)62+ 2204
Fe(CO)5 2022, 2000 209
carbonyl νCO, µ1 (cm−1) νCO, µ2 (cm−1) νCO, µ3
(cm−1)
Rh2(CO)8 2060, 2084 1846, 1862
Rh4(CO)12 2044, 2070, 2074 1886
Rh6(CO)16 2045, 2075 1819
Nuclear magnetic resonance spectroscopy

The traditional method for the study of metal carbonyls is the


13C NMR spectroscopy. To improve the sensitivity of this
technique, complexes are often enriched 13CO. Typical
range for the chemical shift for terminally bound ligands is
150 to 220 ppm. Bridging ligands absorb between 230 to
280 ppm. The 13C signals shift toward higher fields with an
increasing atomic number of the central metal.

The nuclear magnetic resonance spectroscopy can be used


for experimental determination of the complex dynamics.
The activation energy of ligand exchanges processes can be
determined by the temperature dependence of the line
broadening.
Mass spectrometry

Mass spectrometry provides information about the structure


and composition of the complexes. Spectra for metal
polycarbonyls are often easily
interpretable, because the dominant fragmentation process
is the
loss of carbonyl ligands (m/z = 28).

M(CO)n+ → M(CO)n-1+ + CO

Electron impact ionization is the most common technique for


characterizing the neutral metal carbonyls. Neutral metal
carbonyls can be converted to charged species by
derivatization, which enables the use of electrospray
ionization, instrumentation for which is often widely available.
For example, treatment of a metal carbonyl with alkoxide
generates an anionic metallaformate that is amenable to
analysis by ESI-MS:

LnM(CO) + RO− → [LnM-C(=O)OR]−

Some metal carbonyls react with azide to give isocyanato


complexes with release of nitrogen. By adjusting the cone
voltage and/or temperature, the degree of fragmentation can
be controlled. The molar mass of the parent complex can be
determined, as well as information about structural
rearrangements involving loss of carbonyl ligands under ESI-
MS conditions.
Occurrence in nature
A heme unit of human carboxyhemoglobin, showing the
carbonyl ligand at the apical position, trans to the histidine
residue.
In the investigation of the infrared spectrum of the Galactic
Center monoxide vibrations of iron carbonyls in interstellar
dust clouds were detected.
Iron carbonyl clusters were also observed in Jiange H5
chondrites identified by infrared spectroscopy. Four infrared
stretching frequencies were found for the terminal and
bridging carbon monoxide ligands.

In the oxygen-rich atmosphere of earth metal carbonyls are


subject to oxidation to the metal oxides. It is discussed
whether in the reducing hydrothermal environments of the
pre-biotic prehistory such complexes were formed and could
have been available as catalysts for the synthesis of critical
biochemical compounds such as pyruvic acid.
Traces of the carbonyls of iron, nickel, and tungsten were
found in the gaseous emanations from the sewage sludge of
municipal treatment plants.

The hydrogenase enzymes contain CO bound to iron.


Apparently the CO stabilizes low oxidation states, which
facilitates the binding of hydrogen. The enzymes carbon
monoxide dehydrogenase and acetyl coA synthase also are
involved in bio-processing of CO.
Carbon monoxide containing complexes are invoked for the
toxicity of CO and signaling.
Synthesis

The synthesis of metal carbonyls is subject of intense


organometallic research. Since the work of Mond and then
Hieber, many procedures have been developed for the
preparation of mononuclear metal carbonyls as well as
homo-and hetero-metallic carbonyl clusters.
Direct reaction of metal with carbon monoxide

Nickel tetracarbonyl and iron pentacarbonyl can be prepared


according to the following equations by reaction of finely
divided metal with carbon monoxide:[29]

Ni + 4 CO → Ni(CO)4 (1 bar, 55 °C)


Fe + 5 CO → Fe(CO)5 (100 bar, 175 °C)

Nickel carbonyl is formed with carbon monoxide already at


80 °C and atmospheric pressure, finely divided iron reacts at
temperatures between 150 and 200 °C and a carbon
monoxide pressure of 50 to 200 bar.[30] Other metal
carbonyls are prepared by less direct methods.
Reduction of metal salts and oxides

Some metal carbonyls are prepared by the reduction of


metal halides in the presence of high pressure of carbon
monoxide. A variety of reducing agents are employed,
including copper, aluminum, hydrogen, as well as metal
alkyls such as triethylaluminum. Illustrative is the formation
of chromium hexacarbonyl from anhydrous chromium(III)
chloride in benzene with aluminum as a reducing agent, and
aluminum chloride as the catalyst:[29]

CrCl3 + Al + 6 CO → Cr(CO)6 + AlCl3


The use of metal alkyls, e.g. triethylaluminium and
diethylzinc as the reducing agent leads to the oxidative
coupling of the alkyl radical to the dimer:

WCl6 + 6 CO + 2 Al(C2H5)3 → W(CO)6 + 2 AlCl3 + 3


C4H10

Tungsten, molybdenum, manganese, and rhodium salts may


be reduced with lithium aluminum hydride. Vanadium
hexacarbonyl is prepared with sodium as a reducing agent in
chelating solvents such as diglyme.

VCl3 + 4 Na + 6 CO 2 diglyme → Na(diglyme)2[V(CO)6] +


3 NaCl
[V(CO)6]− + H+ → H[V(CO)6] → 1/2 H2 + V(CO)6

In aqueous phase nickel or cobalt salts can be reduced, for


example, by sodium dithionite. In the presence of carbon
monoxide, cobalt salts are quantitatively converted to the
tetracarbonylcobalt(-1) anion:

Co2+ + 1.5 S2O42− + 6 OH− + 4 CO → Co(CO)4− + 3


SO32− + 3 H2O

Some metal carbonyls are prepared using CO as the


reducing agent. In this way, Hieber and Fuchs first prepared
dirhenium decacarbonyl from the oxide:

Re2O7 + 17 CO → Re2(CO)10 + 7 CO2

If metal oxides are used carbon dioxide is formed as a


reaction product. In the reduction of metal chlorides with
carbon monoxide phosgene is formed, as in the preparation
of osmium carbonyl chloride from the chloride salts.
Carbon monoxide is also suitable for the reduction of
sulfides, where carbonyl sulfide is the byproduct.
Photolysis and thermolysis

Photolysis or thermolysis of mononuclear carbonyls


generates bi- and multimetallic carbonyls such as diiron
nonacarbonyl (Fe2(CO)9).
On further heating, the products decompose eventually into
the metal and carbon monoxide.

2 Fe(CO)5 → Fe2(CO)9 + CO

The thermal decomposition of triosmium dodecacarbonyl


(Os3(CO)12) provides higher-nuclear osmium carbonyl
clusters such as Os4(CO)13, Os6(CO)18 up to Os8(CO)23.

Mixed ligand carbonyls of ruthenium, osmium, rhodium, and


iridium are often generated by abstraction of CO from
solvents such as dimethylformamide (DMF) and 2-
methoxyethanol. Typical is the synthesis of IrCl(CO)(PPh3)2
from the reaction of iridium(III) chloride and
triphenylphosphine in boiling DMF solution.
Salt metathesis

Salt metathesis reaction of for example KCo(CO)4 with


[Ru(CO)3Cl2]2 leads selectively to mixed-metal carbonyls
such as RuCo2(CO)11.

KCo(CO)4 + [Ru(CO)3Cl2]2 → 2 RuCo2(CO)11 + 4 KCl

Metal carbonyl cations and carbonylates

The synthesis of ionic carbonyl complexes is possible by


oxidation or reduction of the neutral complexes. Anionic
metal carbonylates can be obtained for example by
reduction of dinuclear complexes with sodium. A familiar
example is the sodium salt of iron tetracarbonylate
(Na2Fe(CO)4,Collman's reagent), which is used in organic
synthesis.

The cationic hexacarbonyl salts of manganese, technetium


and rhenium can be prepared from the carbonyl halides
under carbon monoxide pressure by reaction with a Lewis
acid.

Mn(CO)5Cl + AlCl3 + CO → Mn(CO)6+AlCl4−

The use of strong acids succeeded in preparing gold


carbonyl cations such as [Au(CO)2]+, which is used as a
catalyst for the carbonylation of olefins.
The cationic platinum carbonyl complex [Pt(CO)4]+ can be
prepared by working in so-called super acids such as
antimony pentafluoride.
Reactions

Metal carbonyls are important precursors for the synthesis of


other organometalic complexes. The main reactions are the
substitution of carbon monoxide by other ligands, the
oxidation or reduction reactions of the metal center and
reactions of carbon monoxide ligand.

CO substitution

The substitution of CO ligands can be induced thermally or


photochemically by donor ligands. The range of ligands is
large, and includes phosphines, cyanide (CN−), nitrogen
donors, and even ethers, especially chelating ones. Olefins,
especially diolefins, are effective ligands that afford
synthetically useful derivatives. Substitution of 18-electron
complexes generally follows a dissociative mechanism,
involving 16-electron intermediates.

Substitution proceeds via a dissociative mechanism:

M(CO)n → M(CO)n-1 + CO
M(CO) n-1 + L → M(CO)n-1L

The dissociation energy is 105 kJ mol−1 for nickel carbonyl


and 155 kJ mol−1 for chromium hexacarbonyl.[1]

Substitution in 17-electron complexes, which are rare,


proceeds via associative mechanisms with a 19-electron
intermediates.

M(CO)n + L → M(CO)nL
M(CO)nL → M(CO)n-1L + CO

The rate of substitution in 18-electron complexes is


sometimes catalysed by catalytic amounts of oxidants, via
electron-transfer.
Reduction

Metal carbonyls react with reducing agents such as metallic


sodium or sodium amalgam to give carbonylmetalate anions:

Mn2(CO)10 + 2 Na → 2 Na[Mn(CO)5]

For iron pentacarbonyl, one obtains the tetracarbonylferrate


with loss of CO:

Fe(CO)5 + 2 Na → Na2[Fe(CO)4] + CO

Mercury can insert into the metal-metal bonds of some


polynuclear metal carbonyls:
Co2(CO)8 + Hg → (CO)4Co-Hg-Co(CO)4

Nucleophilic attack at CO

The CO ligand is often susceptible to attack by nucleophiles.


For example, trimethylamine oxide and bistrimethylsilylamide
convert CO ligands to CO2 and CN−, respectively. In the
"Hieber base reaction", hydroxide ion attacks the CO ligand
to give a metallacarboxylic acid, followed by the release of
carbon dioxide and the formation of metal hydrides or
carbonylmetalates. A well-known example of this
nucleophilic addition reaction is the conversion of iron
pentacarbonyl to hydridorion tetracarbonyl anion:

Fe(CO)5 + NaOH → Na[Fe(CO)4CO2H]


Na[Fe(CO)4COOH] + NaOH → Na[HFe(CO)4] + NaHCO3

Protonation of the hydrido anion gives the neutral iron


tetracarbonyl hydride:

Na[HFe(CO)4] + H+ → H2Fe(CO)4 + Na+

Organolithium reagents add with metal carbonyls to


acylmetal carbonyl anions. O-alkylation of these anions, e.g.
with Meerwein salts, affords Fischer carbenes.
Synthesis of Fischer carbenes
With electrophiles

Despite being in low formal oxidation states, metal carbonyls


are relatively unreactive toward many electrophiles. For
example, they resist attack by alkylating agents, mild acids,
mild oxidizing agents. Most metal carbonyls do undergo
halogenation. Iron pentacarbonyl, for example, forms ferrous
carbonyl halides:
Fe(CO)5 + X2 → Fe(CO)4X2 + CO

Metal-metal bonds are cleaved by halogens:

Mn2(CO)10 + Cl2 → 2 Mn(CO)5Cl

Compounds

Most metal carbonyl complexes contain a mixture of ligands.


Examples include the historically important
IrCl(CO)(P(C6H5)3)2 and the anti-knock agent
(CH3C5H4)Mn(CO)3. The parent compounds for many of
these mixed ligand complexes are the binary carbonyls, i.e.
species of the formula [Mx(CO)n]z, many of which are
commercially available. The formula of many metal
carbonyls can be inferred from the 18 electron rule.
Charge-neutral binary metal carbonyls

Group 4 elements with 4 valence electrons are rare, but


substituted derivatives of Ti(CO)7 are known.
Group 5 elements with 5 valence electrons, again are
subject to steric effects that prevent the formation of M-M
bonded species such as V2(CO)12, which is unknown. The
17 VE V(CO)6 is however well known.
Group 6 elements with 6 valence electrons form metal
carbonyls Cr(CO)6, Mo(CO)6, and W(CO)6 (6 + 6x2 = 18
electrons). Group 6 elements (as well as group 7) are well
also well known for exhibiting the cis effect (the labilization of
CO in the cis position) in organometallic synthesis.
Group 7 elements with 7 valence electrons form metal
carbonyl dimers Mn2(CO)10, Tc2(CO)10, and Re2(CO)10 (7
+ 1 + 5x2 = 18 electrons).
Group 8 elements with 8 valence electrons form metal
carbonyls Fe(CO)5, Ru(CO)5 and Os(CO)5 (8 + 5x2 = 18
electrons). The heavier two members are unstable, tending
to decarbonylate to give Ru3(CO)12, and Os3(CO)12. The
two other principal iron carbonyls are Fe3(CO)12 and
Fe2(CO)9.
Group 9 elements with 9 valence electrons and are
expected to form metal carbonyl dimers M2(CO)8. In fact the
cobalt derivative of this octacarbonyl is the only stable
member, but all three tetramers are well known: Co4(CO)12,
Rh4(CO)12, Rh6(CO)16, and Ir4(CO)12 (9 + 3 + 3x2 = 18
electrons). Co2(CO)8 unlike the majority of the other 18 VE
transition metal carbonyls is sensitive to oxygen.
Group 10 elements with 10 valence electrons form metal
carbonyls Ni(CO)4 (10 + 4x2 = 18 electrons). Curiously
Pd(CO)4 and Pt(CO)4 are not stable.

Anionic binary metal carbonyls

Group 4 elements as dianions resemble neutral group 6


derivatives: [Ti(CO)6]2−.[38]
Group 5 elements as monoanions resemble again neutral
group 6 derivatives: [V(CO)6]−.
Group 7 elements as monoanions resemble neutral group
8 derivatives: [M(CO)5]− (M = Mn, Tc, Re).
Group 8 elements as dianaions resemble neutral group 10
derivatives: [M(CO)4]2− (M = Fe, Ru, Os). Condensed
derivatives are also known.
Group 9 elements as monoanions resemble neutral group
10 metal carbonyl. [Co(CO)4]− is the best studied member.

Large anionic clusters of Ni, Pd, and Pt are also well known.
Cationic binary metal carbonyls

Group 7 elements as monocations resemble neutral group


6 derivative [M(CO)6]+ (M = Mn, Tc, Re).
Group 8 elements as dications also resemble neutral
group 6 derivatives [M(CO)6]2+ (M = Fe, Ru, Os).[39]

Metal carbonyl hydrides


Metal Carbonyl hydride pKa
HCo(CO)4 "strong"
HCo(CO)3(P(OPh)3) 5.0
HCo(CO)3(PPh3) 7.0
HMn(CO)5 7.1
H2Fe(CO)4 4.4, 14
[HCo(dmgH)2PBu3] 10.5

Metal carbonyls are relatively distinctive in forming


complexes with negative oxidation states. Examples include
the anions discussed above. These anions can be
protonated to give the corresponding metal carbonyl
hydrides. The neutral metal carbonyl hydrides are often
volatile and can be quite acidic.[40]
Applications
Spheres of nickel manufactured by the Mond process
Metallurgical uses

Metal carbonyls are used in several industrial processes.


Perhaps the earliest application was the extraction and
purification of nickel via nickel tetracarbonyl by the Mond
process (see also carbonyl metallurgy).

By a similar process carbonyl iron, a highly pure metal


powder, is prepared by thermal decomposition of iron
pentacarbonyl. Carbonyl iron is used inter alia for the
preparation of inductors, pigments, as dietary
supplements,[41] in the production of radar-absorbing
materials in the stealth technology,[42] and in Thermal
spraying.
Catalysis
Metal carbonyls are used in a number of industrially
important carbonylation reactions. In the oxo process, an
olefin, dihydrogen, and carbon monoxide react together with
a catalyst (e.g. dicobalt octacarbonyl to give aldehydes.
Illustrative is the production of butyraldehyde:

H2 + CO + CH3CH=CH2 → CH3CH2CH2CHO

Butyraldehyde is converted on an industrial scale to 2-


Ethylhexanol, a precursor to PVC plasticizers, by aldol
condensation, followed by hydrogenation of the resulting
hydroxyaldehyde. The "oxo aldehydes" resulting from
hydroformylation are used for large-scale synthesis of fatty
alcohols, which are precursors to detergents. The
hydroformylation is a reaction with high atom economy,
especially if the reaction proceeds with high regioselectivity.

Hydroformulation mechanism

Another important reaction catalyzed by metal carbonyls is


the hydrocarboxylation. The example below is for the
synthesis of acrylic acid and acrylic acid esters:

Hydrocarboxylation of acetylene

Hydrocarboxylation of acetylene with an alcohol

Also the cyclization of acetylene to cyclooctatetraene uses


metal carbonyl catalysts:

Cyclooctatetraene from acetylene production


In the Monsanto and Cativa processes, acetic acid is
produced from methanol, carbon monoxide, and water using
hydrogen iodide as well as rhodium and iridium carbonyl
catalysts, respectively. Related carbonylation reactions
afford acetic anhydride.
CO-releasing molecules (CO-RMs)

Carbon monoxide-releasing molecules are metal carbonyl


complexes that are being developed as potential drugs to
release CO. At low concentrations, CO functions as a
vasodilatory and an anti-inflammatory agent. CO-RMs have
been conceived as a pharmacological strategic approach to
carry and deliver controlled amounts of CO to tissues and
organs.
Related compounds

Many ligands are known to form homoleptic and mixed


ligand complexes that are analogous to the metal carbonyls.

The development of metal carbonyl clusters such as


Ni(CO)4 and Fe(CO)5 led quickly to the isolation of
Fe2(CO)9 and Fe3(CO)12. Rundle and Dahl discovered that
Mn2(CO)10 featured an "unsupported" Mn-Mn bond, thereby
verifying the ability of metals to bond to one another in
molecules. In the 1970s, Paolo Chini demonstrated that very
large clusters could be prepared from the platinum metals,
one example being [Rh13(CO)24H3]2-. This area of cluster
chemistry has benefited from single-crystal X-ray diffraction.

Structure of Rh4(CO)12.
Nitrosyl complexes

Toxicology
Metal carbonyls are toxic by skin contact, inhalation or
ingestion, in part because of their ability to carbonylate
hemoglobin to give carboxyhemoglobin, which prevents the
binding of O2.

The toxicity of metal carbonyls is due to toxicity of carbon


monoxide, the metal, and because of the volatility and
instability of the complexes. Exposure occurs by inhalation,
or for liquid metal carbonyls by ingestion or due to the good
fat solubility by skin resorption. Most clinical experience were
gained from toxicological poisoning with nickel carbonyl and
iron pentacarbonyl. Nickel carbonyl is considered as one of
the strongest inhalation poisons.
Inhalation of nickel carbonyl causes acute non-specific
symptoms similar to a carbon monoxide poisoning as
nausea, cough, headaches, fever and dizziness. After some
time, severe pulmonary symptoms such as cough,
tachycardia cyanosis or problems in the gastrointestinal tract
occur. In addition to pathological alterations of the lung, such
as by metallation of the alveoli, damages are observed in the
brain, liver, kidneys, adrenal glands and the spleen. A metal
carbonyl poisoning often requires a long-lasting
recovery.[49]

Chronic exposure by inhalation of low concentrations of


nickel carbonyl can cause neurological symptoms such as
insomnia, headaches, dizziness and memory loss. Nickel
carbonyl is considered carcinogenic, but it can take 20 to 30
years from the start of exposure to the clinical manifestation
of cancer.[50]

Metal carbonyl clusters have several properties that suggest


that they may prove as useful catalysts. The absence of
large bulk phases leads to a high surface-to-volume ratio,
which is advantageous in any catalyst application as this
maximizes the reaction rate per unit amount of catalyst
material, which also minimizes cost. Although surface metal
sites in heterogeneous catalysts are coordinatively
unsaturated, most synthetic clusters are not. In general, as
the number of atoms in a metal particle decrease, their
coordination number decreases, and significantly so in
particles having less than 100 atoms.This is illustrated by the
figure at right, which shows dispersion (ratio of
undercoordinated surface atoms to total atoms) versus
number of metal atoms per particle for ideal icosahedral
metal clusters.
Stereodynamics of clusters

Metal clusters are sometimes characterized by a high


degree of fluxionality of surface ligands and adsorbates
associated with a low energy barrier to rearrangement of
these species on the surface.The rearrangement of ligands
on a cluster exterior is indirectly related to the diffusion of
adsorbates on solid metal surfaces. Interconversion ligands
between terminal, double-, and triply bridging sites is often
facile. It has further been found that metal atoms themselves
can easily migrate in or break their bonds with the cluster
structure.

Although metal carbonyl clusters are rarely used, they have


been subjected to many studies aimed at demonstrating
their reactivity. Some of these examples include the
following
Reaction Core Metals Catalyst Reference
Alkene hydroformylation Mo-Rh Mo2RhCp3(CO)5 [31]
Alkene hydroformylation Rh Rh4(CO)10+x(PPh3)2-x
(x=0,2) [32]
CO hydrogenation Ru-Os H2RuOs3(CO)13 [31]
CO hydrogenation Ru-Co RuCo2(CO)11 [31]
CO hydrogenation Ir Ir4(CO)12 [30][32][33]
CO hydrogenation Fe Fe3(CO)12 [34]
Alkene hydrogenation Os-Ni H3Os3NiCp(CO)9 [31]
Alkene hydrogenation Ni Ni2+xCp2+x(CO)2 (x=0,1)
[32]
Alkyne hydrogenation Os-Ni Os3Ni3Cp3(CO)9 [31]
Hydrogenation of aromatics Ni Ni2+xCp2+x(CO)2
(x=0,1) [32]
Acetaldehyde hydrogenation Ni Ni4(Me3CNC)7 [32]
Alkene isomerization V-Cr VCrCp3(CO)3 [31]
Hydrocarbon isomerization Fe-Pt
Fe2Pt(CO)6(NO)2(Me3CNC)2 [31]
Butane hydrogenolysis Rh-Ir Rh3+xIr3-x(CO)16
(x=0,1,2) [31]
Methanol hydrocarbonylation Ru-Co Ru2Co2(CO)13
[31]
Hydrodesulfurization Mo-Fe Mo2Fe2S2Cp2(CO)8
[31]
CO and CO2 methanation Ru-Co HRuCo3(CO)12
[31]
Ammonia synthesis Ru-Ni H3Ru3NiCp(CO)9 [31]
Fischer-Tropsch catalysis

Species that are typical ligands for a metal cluster represent


obvious reactant-catalyst combinations.For example,
hydrogenation of CO (Fischer-Tropsch synthesis) can be
catalyzed using several metal clusters, as shown in the table
above. It has been proposed that coordination of CO to
multiple metal sites weakens the triple-bond enough to allow
hydrogenation.[30] As in the industrially significant
heterogeneous process, Fischer-Tropsch synthesis by
clusters yields alkanes, alkenes, and various oxygenates.
The selectivity is heavily influenced by the particular cluster
used. For example, Ir4(CO)12 produces methanol, whereas
Ru2Rh(CO)12 produces ethylene glycol.[30] Selectivity is
determined by several factors, including steric and electronic
effects. Steric effects are the most important consideration in
many cases, however electronic effects dominate in
hydrogenation reactions where one adsorbate (hydrogen) is
relatively small.[25]

The cyclo oligomerization of thio ethanes illustrates the


influence of steric effects on selectivity.[25] The selectivity
ratio S3(CH2)9/S6(CH2)18 is 6.0 for Os4(CO)12 and 1.5 for
W(CO)6, rationalized by greater steric effects in the Os
cluster, leading to a preference for the smaller ring product.
In some cases, a metal cluster must be "activated" for
catalysis by substitution of one or more ligands, such as
acetonitrile.[35] For example, Os3(CO)12 will have one
active site after thermolysis and the dissociation of a single
carbonyl group. Os3(CO)10(CH3CN)2 will have two active
sites.

Metal nitrosyls, featuring NO as a ligand are numerous,


although homoleptic derivatives are not. Relative to CO, NO
is a stronger acceptor and isocyanides are better donors.
Well known nitrosyl carbonyls include CoNO(CO)3 and
Fe(NO)2(CO)2.
Thiocarbonyls complexes

Complexes containing CS are known but are


uncommon.The rarity of such complexes is attributable in
part to the fact that the obvious source material, carbon
monosulfide, is unstable. Thus, the synthesis of thiocarbonyl
complexes requires more elaborate routes, such as the
reaction of disodium tetracarbonylferrate with thiophosgene:

Na2Fe(CO)4 + CSCl2 → Fe(CO)4CS + 2 NaCl

Complexes of CSe and CTe are very rare.

All metal carbonyls undergo substitution by


organophosphorus ligands. For example, the series
Fe(CO)5-x(PR3)x is well known for various phosphine
ligands for x = 1, 2, and 3. PF3 behaves similarly but is
remarkable because it readily forms homoleptic analogues of
the binary metal carbonyls. For example the volatile, stable
complexes Fe(PF3)5 and Co2(PF3)8 represent CO-free
analogues of Fe(CO)5 and Co2(CO)8 (unbridged isomer).
Isocyanide complexes
Isocyanides also form extensive families of complexes that
are related to the metal carbonyls. Typical isocyanide
ligands are methyl and t-butyl isocyanides (Me3CNC). A
special case is CF3NC, an unstable molecule that forms
stable complexes whose behavior closely parallels that of
the metal carbonyls.
METAL PI COMPLEXES

SECTION 1:3 NITROSYLS

Most complexes containing the NO ligand can be viewed as


derivatives of the nitrosyl cation, NO+. The nitrosyl cation is
isoelectronic with carbon monoxide, thus the bonding
between a nitrosyl ligand and a metal follows the same
principles as the bonding in carbonyl complexes. The nitrosyl
cation serves as a two-electron donor to the metal and
accepts electrons from the metal via back-bonding. The
compounds Co(NO)(CO)3 and Ni(CO)4 illustrate the analogy
between NO+ and CO. Similarly, two NO groups are
isoelectronic with three CO groups.
This trend is illustrated by the isoelectronic pair
Fe(CO)2(NO)2 and [Ni(CO)4].These complexes are
isoelectronic and, incidentally, both obey the 18-electron
rule. The formal description of nitric oxide as NO+ does not
match certain measureable and calculated properties.
In an alternative description, nitric oxide serves as a 3-
electron donor, and the metal-nitrogen interaction is a triple
bond.
linear and bent M-NO bonds
Linear vs bent nitrosyl ligands

The M-N-O unit in nitrosyl complexes is usually linear, or no


more than 15° from linear. In some complexes, however,
especially when back-bonding is less important, the M-N-O
angle can strongly deviate from 180°. In such cases, the NO
ligand is sometimes described as the anion, NO−.
Prototypes for such compounds are the organic nitroso
compounds, such as nitrosobenzene. A complex with a bent
NO ligand is trans-[Co(en)2(NO)Cl]+. Trends in structure and
bonding are usually analyzed using the Enemark-Feltham
approach.[3] In their framework, the factor that determines
the bent vs linear NO ligands in octahedral complexes is the
sum of electrons of pi-symmetry. Complexes with "pi-
electrons" in excess of 6 tend to have bent NO ligands.
Thus, [Co(en)2(NO)Cl]+, with seven electrons of pi-
symmetry (six in t2g orbitals and one on NO), adopts a bent
NO ligand, whereas [Fe(CN)5(NO)]3−, with six electrons of
pi-symmetry, adopts a linear nitrosyl.
Linear and bent NO ligands can be distinguished using
infrared spectroscopy. Linear M-N-O groups absorb in the
range 1650–1900 cm−1, whereas bent nitrosyls absorb in
the range 1525–1690 cm−1. The differing vibrational
frequencies reflect the differing N-O bond orders for linear
(triple bond) and bent NO (double bond).
Bridging nitrosyl ligands

Nitric oxide can also serve as a bridging ligand. In the


compound [Mn3(η5C5H5)3 (μ2-NO)3 (μ3-NO)], three NO
groups bridge two metal centres and one NO group bridge to
all three.[2]
Enemark-Feltham notation

The Enemark-Feltham notation is used to describe the


number of d-type electrons present in a complex. It is
deliberately ambiguous, because the true oxidation state of a
metal coordinated by non-innocent ligands is often unclear.
When written with this notation, the d-electron count is
always consistent no matter how the nitrosyl ligands are
treated (i.e. NO+, NO·, or NO−).

To illustrate its use, the {MNO} d-electron count of the


[Cr(CN)5NO]3− anion is shown. In this example, the cyanide
ligands are "innocent", i.e., they have a charge of −1 each,
−5 total. To balance the fragment's overall charge, the
charge on {CrNO} is thus +2 (−3 = −5 + 2). Using the neutral
electron counting scheme, Cr has 6 d electrons and NO· has
one electron for a total of 7. Two electrons are subtracted to
take into account that fragment's overall charge of +2, to
give 5. Written in the Enemark-Feltham notation, the d
electron count is {CrNO}5. The results are exactly the same
if the nitrosyl ligand were considered NO+ or NO−.[3]
Homoleptic nitrosyl complexes

Metal complexes containing only nitrosyl ligands are called


isoleptic nitrosyls. They are rare, the premier member being
Cr(NO)4. Even trinitrosyl complexes are uncommon,
whereas polycarbonyl complexes are routine.

The nitroprusside anion, [Fe(CN)5NO]2−, a mixed nitrosyl


cyano complex, has pharmaceutical applications as a slow
release agent for NO. The signalling function of NO is
effected via its complexation to haeme proteins, where it
binds in the bent geometry.
Nitric oxide can serve as a ligand in complexes. The
resulting complexes are called metal nitrosyls, and can bond
to a metal atom in two extreme modes: as NO+ and as NO−.
It is generally assumed that NO+ coordinates linearly, the
M−N−O angle being 180°, whereas NO− forms a bent
geometry, with an M−N−O angle of approximately 120°.
Roussin red and black salts

One of the earliest examples of a nitrosyl complex to be


synthesized is Roussin's red salt, which is a sodium salt of
the anion [Fe2(NO)4S2]2−. The structure of the anion can be
viewed as consisting of two tetrahedra sharing an edge.
Each iron atom is bonded linearly to two NO+ ligands and
shares two bridging sulfido ligands with the other iron atom.
Roussin's black salt has a more complex cluster structure.
The anion in this species has the formula [Fe4(NO)7S3]−. It
has C3v symmetry. It consists of a tetrahedron of iron atoms
with sulfide ions on three faces of the tetrahedron. Three iron
atoms are bonded to two nitrosyl groups. The iron atom on
the threefold symmetry axis has a single nitrosyl group which
also lies on that axis.
The anion in Roussin's Red Salt, [Fe2S2(NO)4]2−.
The anion in Roussin's black salt, [Fe4S3(NO)7]−.
The nitroprusside anion, [Fe(CN)5NO]2−, an octahedral
complex containing a "linear NO" ligand.
trans-[Co(en)2(NO)Cl]+, an octahedral complex containing a
"bent NO" ligand.

Preparation
Nitrosyl complexes can be prepared by many routes. Direct
formation from nitric oxide is common. The nitrosylation of
cobalt carbonyl is illustrative:

Co2(CO)8 + 2 NO → 2 CoNO(CO)3 + 2 CO

In such approaches one must guard against the tendency of


nitric oxide to be oxidized by air. Replacement of ligands by
the nitrosyl cation may be accomplished using nitrosyl
tetrafluoroborate, [NO][BF4]. Other indirect methods are
indirect with the NO group deriving from some other species,
often accompanied by oxidation and reduction reactions. A
classic example is provided by the brown ring test in which
the nitrate ion is the source of a nitric oxide ligand. Nitrosyl
chloride is also useful, being applicable to [Mo(NO)2Cl2]n.
Reactions

An important reaction is the acid/base equilibrium:

[LnMNO]2+ + 2OH− is in equilibrium with LnMNO2 + H2O

This equilibrium serves to confirm that the linear nitrosyl


ligand is, formally, NO+, with nitrogen in the oxidation state
+3

NO+ + 2 OH− is in equilibrium with NO2− + H2O

Since nitrogen is more electronegative than carbon, metal-


nitrosyl complexes tend to be more electrophilic than related
metal carbonyl complexes. Nucleophiles often add to the
nitrogen.[1] The nitrogen atom in bent metal nitrosyls is
basic, thus can be oxidized, alkylated, and protonated, e.g.:

(Ph3P)2(CO)ClOsNO + HCl → (Ph3P)2(CO)ClOsN(H)O


In rare cases, NO is cleaved by metal centers:

Cp2NbMe2 + NO → Cp2(Me)Nb(O)NMe
2 Cp2(Me)Nb(O)NMe → 2 Cp2Nb(O)Me + ½MeN=NMe

Applications

Metal-catalyzed reactions of NO are not commonly


synthetically useful. Nitric oxide is however an important
signalling molecule in nature and this fact is the basis of the
most important applications of metal nitrosyls. The
nitroprusside anion, [Fe(CN)5NO]2−, a mixed nitrosyl cyano
complex, has pharmaceutical applications as a slow release
agent for NO. The signalling function of NO is effected via its
complexation to haeme proteins, where it binds in the bent
geometry.

Nitric oxide can serve as a ligand in complexes. The


resulting complexes are called metal nitrosyls, and can bond
to a metal atom in two extreme modes: as NO+ and as NO−.
It is generally assumed that NO+ coordinates linearly, the
M−N−O angle being 180°, whereas NO− forms a bent
geometry, with an M−N−O angle of approximately 120°.
However, the results of many studies have shown that the
ionic descriptions of the NO ligand do not correlate with
metal-NO geometry. A more realistic description of electron-
counting in metal-nitrosyl chemistry is given by the Enemark-
Feltham notation.
Organonitroso compounds

Nitroso compounds can be prepared by the reduction of nitro


compounds or by the oxidation of hydroxylamines. A good
example is (CH3)3CNO, known formally as 2-methyl-2-
nitrosopropane, or t-BuNO, which is prepared by the
following sequence:[1]

(CH3)3CNH2 → (CH3)3CNO2
(CH3)3CNO2 → (CH3)3CNHOH
(CH3)3CNHOH → (CH3)3CNO

(CH3)3CNO is blue and exists in solution in equilibrium with


its dimer, which is colorless, m.p. 80–81 °C.

In the Fischer-Hepp rearrangement aromatic 4-nitroso-


anilines are prepared from the corresponding nitrosamines.
Another named reaction involving a nitroso compound is the
Barton reaction.

Organonitroso compounds serve as a ligands for transition


metals.
Nitrosation vs. nitrosylation

Nitrite can enter two kinds of reaction, depending on the


physico-chemical environment.
Nitrosylation is adding a nitrosyl ion NO− to a metal (e.g.
iron) or a thiol, leading to nitrosyl iron Fe-NO (e.g., in
nitrosylated heme = nitrosylheme) or S-nitrosothiols
(RSNOs).
Nitrosation is adding a nitrosonium ion NO+ to an amine -
NH2 leading to a nitrosamine. This conversion occurs at
acidic pH, particularly in the stomach, as shown in the
equation for the formation of N-phenylnitrosamine:

NO2− + H+ \overrightarrow{\leftarrow} HONO


HONO + H+ \overrightarrow{\leftarrow} H2O + NO+
C6H5NH2 + NO+ → C6H5N(H)NO + H+

Many primary alkyl N-nitroso compounds, such as


CH3N(H)NO, tend to be unstable with respect to hydrolysis
to the alcohol. Those derived from secondary amines (e.g.,
(CH3)2NNO derived from dimethylamine) are more robust. It
is these N-nitrosamines that are carcinogens in rodents.

In food Nitrosyl-heme

In foodstuffs and in the gastro-intestinal tract, nitrosation and


nitrosylation do not have the same consequences on
consumer health.

In cured meat: Meat processed by curing contains nitrite and


has a pH of 5 approximately, where almost all nitrite is
present as NO2− (99%). Cured meat is also added with
sodium ascorbate (or erythorbate or Vitamin C). As
demonstrated by S. Mirvish, ascorbate inhibits nitrosation of
amines to nitrosamine, because ascorbate reacts with NO2−
to form NO.
Ascorbate and pH 5 thus favor nitrosylation of heme iron,
forming nitrosyl-heme, a red pigment when included inside
myoglobin, and a pink pigment when it has been released by
cooking. It participates to the "bacon flavor" of cured meat:
nitrosyl-heme is thus considered a benefit for the meat
industry and for consumers.

In the stomach: secreted Hydrogen Chloride makes an


acidic environment
(pH=2) and ingested nitrite (with food or saliva) leads to
nitrosation of amines, that yields nitrosamines (potential
carcinogens). Nitrosation is low if amine concentration is low
(e.g., low-protein diet, no fermented food) or if Vitamin C
concentration is high (e.g., high fruit diet). Then S-
nitrosothiols are formed, that are stable at pH 2.

In the colon: neutral pH does not favor nitrosation. No


nitrosamine is formed in stools, even after addition of a
secondary amine or nitrite. Neutral pH favors NO− release
from S-nitrosothiols, and nitrosylation of iron. The previously
called NOC (N-nitroso compounds) measured by Bingham's
team in stools from red meat-fed volunteers were, according
to Bingham and Kuhnle, largely non-N-nitroso ATNC
(Apparent Total Nitroso Compounds), e.g., S-nitrosothiols
and nitrosyl iron (as nitrosyl heme).
METAL PI COMPLEXES

DI -OXYGEN COMPLEXES
Dioxygen complexes are coordination compounds that
contain O2 as a ligand. The study of these compounds is
inspired by oxygen-carrying proteins such as myoglobin,
hemoglobin, hemerythrin, and hemocyanin.[3] Several
transition metals form complexes with O2, and many of
these complexes form reversibly.[4] The binding of O2 is the
first step in many important phenomena, such as cellular
respiration, corrosion, and industrial chemistry. The first
synthetic oxygen complex was demonstra

Mononuclear complexes of O2
TMdioxygenCmpx.pngted in 1938 with cobalt(II) complex
reversibly bound O2Complexes of η1-O2 ligands

O2 binds to a single metal center either “end-on” (η1-) or


“side-on” (η2-). The bonding and structures of these
compounds are usually evaluated by single-crystal X-ray
crystallography, focusing both on the overall geometry as
well as the O---O distances, which reveals its bond order.
O2 adducts derived from cobalt(II) and iron(II) porphyrin
complexes and related anionic ligands exhibit this bonding
mode. Myoglobin and hemoglobin are famous examples,
and many synthetic analogues have been described that
behave similarly. Binding of O2 is usually described as
proceeding via electron transfer from the metal(II) center to
give superoxide (O2−) complexes of metal(III) centers.
Complexes of η2-O2 ligands

η2- bonding is the most common motif seen in coordination


chemistry of dioxygen. Such complexes can generated by
treating low-valent metal complexes with gaseous oxygen.
For example, Vaska's complex reversibly binds O2 (Ph =
C6H5):

IrCl(CO)(PPh3)2 + O2 \overrightarrow{\leftarrow}
IrCl(CO)(PPh3)2O2

The conversion is described as a 2 e− redox process: Ir(I)


converts to Ir(III) as dioxygen converts to peroxide. Since O2
has a triplet ground state and Vaska's complex is a singlet,
the reaction is slower than when singlet oxygen is used.[6]

Complexes containing η2-O2 ligands are fairly common, but


most are generated using hydrogen peroxide, not O2.
Chromate ([CrO4)]2−) can for example be converted to the
tetraperoxide [Cr(O2)4]2−. The reaction of hydrogen
peroxide with aqueous titanium(IV) gives a brightly colored
peroxy complex that is a useful test for titanium as well as
hydrogen peroxide.[7]
Binuclear complexes of O2

O2 can bind to one metal of a bimetallic unit via the same


modes discussed above for mononuclear complexes. A well-
known example in nature is hemerythrin, which features a
diiron carboxylate that binds O2 at one Fe center. Dinuclear
complexes can also cooperate in the binding, although the
initial attack of O2 probably occurs at a single metal. These
binding modes include μ2-η2,η2-, μ2-η1,η1-, and μ2-η1,η2-.
Depending on the degree of electron-transfer from the
dimetal unit, these O2 ligands can again be described as
peroxo or superoxo. In nature, such dinuclear dioxygen
complexes often feature copper.[8]

Relationship to other oxygenic ligands and applications

Dioxygen complexes are the precursors to other families of


oxygenic ligands. Metal oxo compounds arise from the
cleavage of the O-O bond after complexation. Hydroperoxo
complexes are generated in the course of the reduction of
dioxygen by metals. The reduction of O2 by metal catalysts
is a key half-reaction in fuel cells.
Metal-catalyzed oxidations with O2 proceed via the
intermediacy of dioxygen complexes, although the actual
oxidants are often oxo derivatives. The reversible binding of
O2 to metal complexes has been used as a means to purify
oxygen from air, but cryogenic distillation of liquid air
remains the dominant technology.

2nd Generation Grubbs Catalyst, based on a saturated N-


heterocyclic carbene (1,3-bis(2,4,6-
trimethylphenyl)dihydroimidazole):

In the Hoveyda–Grubbs Catalysts, the benzylidene ligands


have a chelating ortho-isopropoxy group attached to the
benzene rings. The ortho-isopropoxybenzylidene moiety is
sometimes referred to as a Hoveyda chelate. The chelating
oxygen atom replaces a phosphine ligand, which in the case
of the 2nd generation catalyst, gives a completely
phosphine-free structure.
SECTION 1:5 DI-NITROGEN COMPLEXES

Metal dinitrogen complexes are a coordination compounds


that contain the dinitrogen (N2) as a ligand. In the area of
coordination chemistry, the atomic and diatomic forms of
nitrogen are distinguished, although otherwise "nitrogen"
refers to N2.

Metal complexes of N2 have been studied since 1965 when


the first complex was reported by Allen and Senoff. This
complex, [Ru(NH3)5(N2)]2+ was synthesised from hydrazine
hydrate and ruthenium trichloride and consists of a 16e−
[Ru(NH3)5]2+ centre attached to one end of N2.[1][2]
Interest in such complexes arises because N2 comprises the
majority of the atmosphere and because many useful
compounds contain nitrogen atoms. Biological nitrogen
fixation probably occurs via the binding of N2 to a metal
center in the enzyme nitrogenase, followed by a series of
steps that involve electron transfer and protonation. The
hydrogenation of N2 is only weakly exothermic, hence the
industrial hydrogenation of nitrogen via the Haber-Bosch
Process employs high pressures and high
temperatures.Bonding modes
In terms of its bonding to transition metals, N2 is related to
CO and acetylene as all three species have triple bonds. A
variety of bonding modes have been characterized.
End-on

As a ligand, N2 usually binds to metals as an "end-on"


ligand, as illustrated by Allen and Senoff's complex. Such
complexes are usually analogous to related CO derivatives.
A good example of this relationship are the complexes
IrCl(CO)(PPh3)2 and IrCl(N2)(PPh3)2.[4] Few complexes
contain more than one N2 ligand, and no example features
three (in contrast metal hexacarbonyls are common). The
dinitrogen ligand in W(N2)2(Ph2CH2CH2PPh2)2 can be
reduced to produce ammonia.
Bridging, end-on

N2 also serves as a bridging ligand, as illustrated by


{[Ru(NH3)5]2(μ-N2)}4+.
A study in 2006 of iron-dinitrogen complexes showed that
the N–N bond is significantly weakened upon complexation
with iron atoms with a low coordination number. The
complex involved bidentate chelating ligands attached to the
iron atoms in the Fe–N–N–Fe core, in which N
2 acts as a bridging ligand between the iron atoms.
Increasing the coordination number of iron by modifying the
chelating ligands and adding another ligand per iron atom
showed an increase in the strength of the N–N bond in the
resulting complex. It is thus suspected that Fe in a low-
coordination environment is a key factor to the fixation of
nitrogen by the nitrogenase enzyme, since its Fe–Mo
cofactor also features Fe with low coordination numbers.
Side-on, bridging

In a second mode of bridging, bimetallic complexes are


known wherein the N-N vector is perpendicular to the M-M
vector. One example is [(η5-C5Me4H)2Zr]2(μ2,η2,η2-N2).
METAL PI COMPLEXES : PHOSPHINES

Phosphines are used as ligands for many metal complexes.


Perhaps the most popular phosphine ligand used is
triphenylphosphine, a shelf-stable solid that undergoes
oxidation in air relatively slowly. Unlike most metal ammine
complexes, metal phosphine complexes tend to be lipophilic,
displaying good solubility in organic solvents.
They also are compatible with metals in multiple oxidation
states. Because of these two features, metal phosphine
complexes are useful in homogeneous catalysis.
Prominent examples of metal phosphine complexes include
Wilkinson's catalyst (Rh(PPh3)3Cl), Grubbs' catalyst, and
tetrakis(triphenylphosphine)palladium(0).

Phosphines are L-type ligands. They are both σ-donors and


π-acceptors. Phosphine ligands' π-acidity is due to P-C σ*
anti-bonding orbitals. Arylphosphines are much stronger π-
acceptors than alkylphosphines, which are poor π-acceptors
or possibly π-donors. The phosphine with the strongest π-
acidity is trifluorophosphine (PF3); its π-acidity approaches
that of the carbonyl ligand.[
Phosphines primarily function as Lewis bases, interacting
with metals as σ donor ligands. They also can accept
electron density from metal into P–C σ* antibonding orbitals
that have π symmetry.[5

One of the first applications of phosphine ligands in catalysis


was the use of triphenylphosphine in “Reppe” chemistry
(1948), which included reactions of alkynes, carbon
monoxide, and alcohols
In his studies, Reppe discovered that this reaction more
efficiently produced acrylic esters using NiBr2(PPh3)2 as a
catalyst instead of NiBr2. Shell developed cobalt-based
catalysts modified with trialkylphosphine ligands for
hydroformylation (now a rhodium catalyst is more commonly
used for this process).

Grubbs' Catalysts are a series of transition metal carbene


complexes used as catalysts for olefin metathesis. They are
named after Robert H. Grubbs, the chemist who first
synthesized them. There are two generations of the catalyst,
as shown on the right.In contrast to other olefin metathesis
catalysts, Grubbs' catalysts tolerate other functional groups
in the alkene, are air-tolerant and are compatible with a wide
range of solvents.For these reasons, Grubbs' catalysts have
become popular in synthetic organic chemistry

This initial ruthenium catalyst was followed in 1995 by what


is now known as the first generation Grubbs catalyst. It is
easily synthesized from RuCl2(PPh3)3,
phenyldiazomethane, and tricyclohexylphosphine in a one-
pot synthesis.
The first generation Grubbs catalyst, while largely replaced
by the second generation catalyst in usage, was not only the
first catalyst to be developed other than those developed by
Richard R. Schrock (Schrock carbenes), but is also
important as a precursor to all other Grubbs-type catalysts.
The second generation catalyst has the same uses in
organic synthesis as the first generation catalyst, but
generally with higher activity. This catalyst is stable toward
moisture and air, thus is easier to handle in the lab.

Shortly before the discovery of the 2nd generation Grubbs'


catalyst, a very similar catalyst based on an unsaturated N-
heterocyclic carbene (1,3-bis(2,4,6-
trimethylphenyl)imidazole) was reported independently by
Nolan and Grubbs in March 1999, and by Fürstner in June of
the same year. Shortly thereafter, in August 1999, Grubbs
reported the 2nd generation 2nd Generation Grubbs
Catalyst, based on a saturated N-heterocyclic carbene (1,3-
bis(2,4,6-trimethylphenyl)dihydroimidazole):
In both the saturated and unsaturated cases a phosphine
ligand is replaced with an N-heterocyclic carbene (NHC),
which is characteristic of all 2nd generation type catalysts.

Both the 1st and 2nd generation catalysts are commercially


available, along with many derivatives of the 2nd generation
catalyst.
Hoveyda–Grubbs Catalyst

In the Hoveyda–Grubbs Catalysts, the benzylidene ligands


have a chelating ortho-isopropoxy group attached to the
benzene rings. The ortho-isopropoxybenzylidene moiety is
sometimes referred to as a Hoveyda chelate. The chelating
oxygen atom replaces a phosphine ligand, which in the case
of the 2nd generation catalyst, gives a completely
phosphine-free structure. The 1st generation Hoveyda–
Grubbs catalyst was reported in 1999 by the Hoveyda
group,[13] and in the following year, the 2nd generation
Hoveyda–Grubbs catalyst was described in nearly
simultaneous publications by the Blechert[14] and
Hoveyda[15] laboratories. Blechert's name is not commonly
included in the eponymous catalyst name. The Hoveyda–
Grubbs catalysts, while more expensive and slower to
initiate than the Grubbs catalyst from which they are derived,
are popular because of their improved stability.[3] Hoveyda–
Grubbs catalysts are easily formed from the corresponding
Grubbs catalyst by the addition of the chelating ligand and
the use of a phosphine scavenger like copper(I) chloride:[15]
Preparation of Hoveyda–Grubbs Catalyst from the 2nd
generation Grubbs Catalyst
The 2nd generation Hoveyda–Grubbs catalysts can also be
prepared from the 1st generation Hoveyda–Grubbs catalyst
by the addition of the NHC:
Preparation of Hoveyda–Grubbs Catalyst from the 1st
generation version

In one study a water-soluble Grubbs catalyst is prepared by


attaching a polyethylene glycol chain to the imidazolidine
group This catalyst is used in the ring-closing metathesis
reaction in water of a diene carrying an ammonium salt
group making it water-soluble as well.Ring closing
metathesis reaction in water with water-soluble

The initiation rate of the Grubbs' catalyst can be altered by


replacing the phosphine ligand with more labile pyridine
ligands. By using 3-bromopyridine the initiation rate is
increased more than a million fold:
The principle application of the fast-initiating catalysts is as
initiators for ring opening metathesis polymerisation
(ROMP). Because of their usefulness in ROMP these
catalysts are sometimes referred to as the 3rd generation
Grubbs' catalysts.The high ratio of the rate of initiation to the
rate of propagation makes these catalysts useful in living
polymerization, yielding polymers with low polydispersity.
On October 5, 2005, Robert H. Grubbs, Richard R. Schrock
and Yves Chauvin won the Nobel Prize in Chemistry in
recognition of their contributions to the development of this
widely used process.

Applications

Olefin metathesis is a reaction between two molecules


containing double bonds. The groups bonded to the carbon
atoms of the double bond are exchanged between
molecules, to produce two new molecules containing double
bonds with swapped groups. Whether a cis isomer or trans
isomer is formed in this type of reaction is determined by the
orientation the molecules assume when they coordinate to
the catalyst, as well as the sterics of the substituents on the
double bond of the newly forming molecule.
CHAPTER 3 POLYACIDS

STRUCTURE :J.F.Keggin ,Nature 1933,131,908

Phosphotungstate ion
A heteropoly acid is a class of acid made up of a particular
combination of hydrogen and oxygen with certain metals and non-
metals. This type of acid is a common re-usable acid catalyst in
chemical reactions

To qualify as a heteropoly acid, the compound must contain:

a metal such as tungsten, molybdenum or vanadium, termed the


addenda atom;
oxygen;
an element generally from the p-block of the periodic table, such as
silicon, phosphorus or arsenic, termed the hetero atom;
acidic hydrogen atoms.

The metal addenda atoms linked by oxygen atoms form a cluster with
the hetero-atom inside bonded via oxygen atoms. Examples with
more than one type of metal addenda atom in the cluster are well
known. The conjugate anion of a heteropoly acid is known as a
polyoxometalate.

Due to the possibilities of there being different combinations of


addenda atoms and different types of hetero atoms there are a lot of
heteropolyacids. Two of the better known groups of these are based
on the Keggin, HnXM12O40, and Dawson, HnX2M18O62, structures.

Some examples are:

H4Xn+M12O40, X = Si, Ge; M = Mo, W


H3Xn+M12O40, X = P, As; M = Mo, W
H6X2M18O62, X=P, As; M = Mo, W

The heteropolyacids are widely used as homogeneous and


heterogeneous catalysts,[2] particularly those based on the Keggin
structure as they can possess qualities such as good thermal
stability, high acidity and high oxidising ability. Some examples of
catalysis are:[3]

Homogeneous acid catalysis


hydrolysis of propene to give propan-2-ol by H3PMo12O40 and
H3PW12O40
Prins reaction by H3PW12O40
polymerisation of THF by H3PW12O40
Heterogeneous acid catalysis
dehydration of propan-2-ol to propene and methanol to
hydrocarbons by H3PW12O40
reformation of hexane to 2-methylpentane (isohexane) by
H3PW12O40 on SiO2
Homogeneous oxidation
cyclohexene + H2O2 to adipic acid by the mixed addenda
H3PMo6V6O40
ketone by O2 to acid and aldehyde by mixed addenda
H5PMo10V2O40

Heteropolyacids have long been used in analysis and histology and


are a component of many reagents e.g. the Folin-Ciocalteu reagent,
folins phenol reagent used in the Lowry protein assay and EPTA,
ethanolic phosphotungstic acid.
See also

Phosphotungstic acid

Phosphotungstic acid (PTA), tungstophosphoric acid (TPA), is a


heteropoly acid with the chemical formula H3PW12O40. It is normally
present as a hydrate. EPTA is the name of ethanolic phosphotungstic
acid, its alcohol solution used in biology. It has the appearance of
small, colorless-grayish or slightly yellow-green crystals, with melting
point 89 °C (24 H2O hydrate). It is odorless and soluble in water (200
g/100 ml). It is not especially toxic, but is a mild acidic irritant. The
compound is known by a variety of different names and acronyms
(see 'other names' section of infobox).

In these names the "12" or "dodeca" reflects the fact that the anion
contains 12 tungsten atoms. Some early workers who did not know
the structure, such as Hsien Wu,[2] called it phospho-24-tungstic
acid, formulating it as 3H2O.P2O5 24WO3.59H2O,
(P2W24O80H6).29H2O, which correctly identifies the atomic ratios of
P, W and O. This formula was still quoted in papers as late as 1970.
Phosphotungstic acid is used in histology as a component for staining
of cell specimens, often together with haematoxylin as PTAH. It binds
to fibrin, collagen, and fibres of connective tissues, and replaces the
anions of dyes from these materials, selectively decoloring them.

Phosphotungstic acid is electron dense, opaque for electrons. It is a


common negative stain for viruses, nerves, polysaccharides, and
other biological tissue materials for imaging by a transmission
electron microscope.
Gouzerh summarises the historical views on the structure of
phosphotungstic acid leading up to Keggin's determination of the
structure as:

H7[P(W2O7)6] proposed by Miolati and further developed by


Rosenheim
H3[PO4W12O18(OH)36] (Pauling)

The structure was determined by J.F Keggin first published in 1933[5]


and then in 1934[6] and is generally known as the Keggin structure.
The anion has full tetrahedral symmetry and comprises a cage of
twelve tungsten atoms linked by oxygen atoms with the phosphorus
atom at its centre. The picture on the right shows the octahedral
coordination of oxygen atoms around the tungsten atoms, and that
the surface of the anion has both bridging and terminal oxygen
atoms. Further investigation showed that the compound was a
hexahydrate not a pentahydrate as Keggin had proposed.
Preparation and chemical properties

Phosphotungstic acid can be prepared by the reaction of sodium


tungstate, Na2WO4.2H2O, with phosphoric acid, H3PO4, acidified
with hydrochloric acid, HCl.[2]

Phosphotungstic acid solutions decompose as the pH is increased. A


step-wise decomposition has been determined and the approximate
compositions at various pH values are as follows:[8]

pH principal components
1.0 [PW12O40]3−
2.2 [PW12O40]3−, [P2W21O71]6−, [PW11O39]7−
3.5 [PW12O40]3−, [P2W21O71]6−, [PW11O39]7−,
[P2W18O62]6−, [P2W19O67]10−
5.4 [P2W21O71]6−, [PW11O39]7−, [P2W18O62]6−
7.3 [PW9O34]9−
8.3 PO43−, WO42−

The species [PW11O39]7− is a lacunary, or defective Keggin ion.


The [P2W18O62]6− has a Dawson structure. At pH less than 8, the
presence of ethanol or acetone stabilises the anion, [PW12O40]3−,
reducing decomposition.[8]
Tungstophosphoric acid is thermally stable up to 400 °C, and is more
stable than the analogous silicotungstic acid, H4SiW12O40.[9]

Large quantities of polar molecules such as pyridine are absorbed


into the bulk phase and not simply on the surface. Solid state NMR
studies of ethanol absorbed in the bulk phase show that both
protonated dimers, ((C2H5OH)2H+) and monomers, (C2H5OH2+)
are present.

Phosphotungstic acid is less sensitive to reduction than


phosphomolybdic acid. Reduction with uric acid or iron(II) sulfate
produces a brown coloured compound. the related silicotungstic acid
when reduced forms a similar brown compound where one of the four
W3 units in the Keggin structure becomes a metal-metal bonded
cluster of three edge shared W(IV) octahedra.[10]

Phosphotungstic acid is the strongest of heteropolyacids. Its


conjugate base is the PW12O403− anion.[11] Its acidity in acetic acid
has been investigated and shows that the three protons dissociate
independently rather than sequentially, and the acid sites are of the
same strength.[12] One estimate of the acidity is that the solid has an
acidity stronger than H0 =−13.16,[9] which would qualify the
compound as a superacid. This acidic strength means that even at
low pH the acid is fully dissociated.
Uses
Catalyst

In common with the other heteropolyacids phosphotungstic acid is a


catalyst and its high acidity and thermal stability make it a catalyst of
choice according to some researchers.[13] It is in solution as a
homogeneous catalyst, and as a heterogeneous catalyst "supported"
on a substrate e.g. alumina, silica. Some acid catalysed reactions
include:

the homogeneous catalysis of the hydrolysis of propene to give 2-


propanol
the homogeneous catalysis of the Prins reaction
the heterogeneous catalysis of the dehydration of 2-propanol to
propene and methanol to hydrocarbons.
Dyeing and pigments

Phosphotungstic acid has been used to precipitate different types of


dyes as "lakes".[14] Examples are basic dyes and triphenylmethane
dyes, e.g. pararosaniline derivatives.[15]
Histology

Phosphotungstic acid is used in histology for staining specimens, as


a component of phosphotungstic acid haematoxylin, PTAH, and
“trichrome” reagents, and as a negative stain for imaging by a
transmission electron microscope.

Phosphotungstic acid haematoxylin (PTAH)

Mallory described the reagent now generally known as PTAH in


1897.[16] PTAH stains tissues either reddish brown or blue
depending on their type. This property of simultaneously staining two
different colours is different from other haematoxylin reagents e.g.
alum-haematoxylin. The role of phosphotungstic acid and the
mechanism of staining is not fully understood. Interestingly the active
component of haematoxylin is the oxidised form, haematin, although
this rarely acknowledged in the literature which refer to haematoxylin
staining. Phosphotungstic acid forms a lake with haematin.[17] The
make –up of the reagent is uncertain, examination of a year old
sample showed there to be three coloured components, blue, red and
yellow.[18] These were not identified. Some investigations of “model”
systems, reacting various compounds such as amino acids, purines,
pyrimidines and amines with PTAH show that they give rise to
different colours.[19]

Trichrome reagents
In these reagents two or three basic dyes are used with
phosphotungstic acid, in either a one step or multi-stage procedure.
These reagents colour different tissue types different colours. Again
the mechanism of staining is not fully understood. Some explanations
include the proposal that phosphotungstic acid acts as a mordant to
bind the dye to the tissue[20] or that alternatively it binds to tissue
blocking it to dye molecules.[21]
Negative staining
Adsorption onto tissue or the surface of viruses and its electron
density are the bases of phosphotungstic acids action as a negative
stain. This electron density arises from the presence of the 12
tungsten atoms which each have an atomic number of 74. The
mechanism of the adsorption onto tissue has been proposed as being
electrostatic rather than involving hydrogen bonding, as adsorption is
not affected by pH.[3]

Analysis

The potassium salt is only slightly soluble, unlike most other


phosphotungstate salts, and has been proposed as a method for the
gravimetric analysis of potassium.[22]
Precipitation of proteins

In a number of analytical procedures one of the roles of


phosphotungstic acid is to precipitate out proteins. It has been termed
a "universal" precipitant for polar proteins.[23] Further studies showed
that no precipitation occurred with α-amino groups but did occur with
guanidino, ε-amino and imidazole groups.[
Medicinal

Very little work appears to have been carried out in this area. One
example relates to liver necrosis in rats.[
Composite proton exchange membranes

The heteropoly acids, including phosphotungstic acid, are being


investigated as materials in composite proton exchange membranes,
such as Nafion. The interest lies in the potential of these composite
materials in the manufacture of fuel cells as they have improved
operating characteristics.
Phosphomolybdic acid
Silicotungstic acid
The first α-Keggin anion, ammonium phosphomolybdate
((NH4)3[PMo12O40]), was first reported by Berzelius in 1826. In
1892, Blomstrand proposed the structure of phosphomolybdic acid
and other poly-acids as a chain or ring configuration. Alfred Werner,
using the coordination compounds ideas of Copaux, attempted to
explain the structure of silicotungstic acid. He assumed a central
group, [SiO4]4− ion, enclosed by four [RW2O6]+, where R is a
unipositive ion. The [RW2O6]+ are linked to the central group by
primary valences. Two more R2W2O7 groups were linked to the
central group by secondary valences. This proposal accounted for the
characteristics of most poly-acids, but not all.

In 1928, Linus Pauling proposed a structure for α-Keggin anions


consisting of a tetrahedral central ion, [XO4]n−8, caged by twelve
WO6 octahedral. In this proposed structure, three of the oxygen on
each of the octahedral shared electrons with three neighboring
octahedral. As a result, 18 oxygen atoms were used as bridging
atoms between the metal atoms. The remaining oxygen atoms
bonded to a proton. This structure explained many characteristics
that were observed such as basicities of alkali metal salts and the
hydrated of some of the salts. However the structure could not
explain the structure of dehydrated acids.

J.F. Keggin with the use of X-ray diffraction experimentally


determined the structure of α-Keggin anions in 1934. The Keggin
structure accounts for both the hydrated and dehydrated α-Keggin
anions without a need for significant structural change. The Keggin
structure is the widely accepted structure for the α-Keggin anions.
The structure is composed of one heteroatom surrounded by four
oxygen to form a tetrahedron. The heteroatom is located centrally
and caged by 12 octahedral MO6-units linked to one another by the
neighboring oxygen atoms. There are a total of 24 bridging oxygen
atoms that link the 12 addenda atoms. The metal centres in the 12
octahedra are arranged on a sphere almost equidistant from each
other, in four M3O13 units, giving the complete structure an overall
tetrahedral symmetry. The bond length between atoms varies
depending on the heteroatom (X) and the addenda atoms (M). For
the 12–phosphotungstic acid, Keggin determined the bond length
between the heteroatom and each the four central oxygen atoms to
be 1.5 Å. The bond length form the central oxygen to the addenda
atoms is 2.43 Å. The bond length between the addenda atoms and
each of the bridging oxygen is 1.9 Å. The remaining 12 oxygen atoms
that are each double bonded to an addenda atom have a bond length
of 1.70 Å. The octahedra are therefore distorted.[3][4] This structure
allows the molecule to hydrate and dehydrate without significant
structural changes and the molecule is thermally stable in the solid
state for use in vapor phase reactions at high temperatures (400−500
°C).

Structure
Reference
Dodecatungstophosphoric acid hexahydrate,
(H5O2+)3(PW12O403−). The true structure of Keggin's
`pentahydrate' from single-crystal X-ray and neutron
diffraction data Brown G.M., Noe-Spirlet M.-R., Busing W.R.,
Levy H.A., Acta Crystallogr., 1977, B33, 1038
doi:10.1107/S0567740877005330
Isomerism

Including the original Keggin structure there are 5 isomers,


designated by the prefixes α-, β-,γ-, δ- and ε-. The original Keggin
structure is designated α- . These isomers are sometimes termed
Baker, Baker-Figgis or rotational isomers,[6] These involve different
rotational orientations of the Mo3O13 units, which lowers the
symmetry of the overall structure.
Lacunary Keggin structures

The term lacunary is applied to ions which have a fragment missing,


sometimes called defect structures. Examples are the (XM11O39)n−
and (XM9O34)n− formed by the removal from the Keggin structure of
sufficient Mo and O atoms to eliminate 1 or 3 adjacent MO6
octahedra. The Dawson structure, X2M18O62n−, is made up of two
Keggin lacunary fragment with 3 missing octahedra.
Dawson structure
Group 13 cations with the Keggin structure

The cluster cation (Al13O4(OH)24(H2O)12)7+ has the Keggin


structure with a tetrahedral Al atom in the centre of the cluster
coordinated to 4 oxygen atoms. The formula can be expressed as
(AlO4Al12(OH)24(H2O)12)7+.[7] This ion is generally called the Al13
ion. A Ga13 analogue is known[8] an unusual ionic compound with an
Al13 cation and a Keggin polyoxoanion has been characterised.[9]
The iron Keggin ion

Due to the similar aqueous chemistries of aluminum and iron, it has


been long thought that an analogous iron polycation should be
isolatable from water. Moreover, in 2007, the structure of ferrihydrite
was determined and shown to be built of iron Keggin ions.[10] This
further captured scientists’ imagination and drive to isolate the iron
Keggin ion. In 2015, the iron Keggin ion was isolated from water, but
as a polyanion with a -17 charge; and protecting chemistry was
required. Iron-bound water is very acidic; so it is difficult to capture
the intermediate Keggin ion form without bulky and nonprotic ligands
instead of the water that is found in the aluminum Keggin ion.
However, more important in this synthesis was the bismuth (Bi3+)
counterions that provided high positive charge to stabilize the high
negative charge of the heptadecavalent polyanion.
Chemical properties

The stability of the Keggin structure allows the metals in the anion to
be readily reduced. Depending on the solvent, acidity of the solution
and the charge on the α-Keggin anion, it can be reversibly reduced in
one- or multiple electron step.[12] For example silicotungstate anion
can be reduced to 20th state.[13] Some anions such as silicotungstic
acid are strong enough as an acid as sulfuric acid and can be used in
its place as an acid catalyst.
Preparation

In general α-Keggin anions are synthesized in acidic solutions. For


example, 12-Phosphotungstic acid is formed by condensing
phosphate ion with tungstate ions. The heteropolyacid that is formed
has the Keggin structure.[5]

PO43− + 12 WO42− + 27H+ → H3PW12O40 + 12H2O

Uses

α-Keggin anions have been used as catalyst in the following


reactions: hydration, polymerization and oxidation reaction as
catalysts
Japanese chemical companies have commercialized the use of the
compounds in hydration of propene, oxidation of methacrolein,
hydration isobutene, hydration of n-butene, and polymerization of
THF
Suppliers

12-Phosphotungstic acid the compound J.F. Keggin used to


determine the structure can be purchased commercially. Other
compounds that contain the α-Keggin anion such as silicotungstic
acid and phosphomolybdic acid are also commercially available at
Aldrich Chemicals, Fisher Chemicals, Alfa Aesar, VWR Chemical,
American Elements, etc.

Structure of the Molecule of 12-


Phosphotungstic Acid J. F.
Keggin, Nature 1933, 131, 908.
REFERENCES TO TEXT

CLUSTERS
Inorganic Chemistry Huheey, JE, 3rd ed. Harper and Row,
New York
Mingos, D. M. P.; Wales, D. J. (1990). Introduction to cluster
chemistry. Englewood Cliffs, N.J: Prentice Hall. ISBN
0134743059.
P. W. Sutton; L. F. Dahl (1967). "Molecular Structure of
Co3(CO)9CCH3. A Tricyclic Organocobalt Complex
Containing a Metal-Coordinated Triply Bridging Aliphatic
Carbon Atom". J. Am. Chem. Soc. 89: 261–268.
doi:10.1021/ja00978a016.
D. Seyferth; J. E. Hallgren; P. L. K. Hung (1973). "The
Preparation of Functional Alkylidynetricobalt Nonacarbonyl
Complexes from Dicobalt Octacarbonyl". J. Organomet.
Chem. 50: 265–275. doi:10.1016/S0022-328X(00)95113-1.
D. Seyferth (1976). "Chemistry of Carbon-Functional
Alkylidynetricobalt Nonacarbonyl Cluster Complexes". Adv.
Organomet. Chem. 14: 97–144. doi:10.1016/s0065-
3055(08)60650-4.
A. Bino; M. Ardon; I. Maor; M. Kaftory; Z. Dori (1976).
"[Mo3(OAc)6(CH3CH2O)2(H2O)3]2+ and Other New
Products of the Reaction between Molybdenum
Hexacarbonyl and Acetic Acid". J. Am. Chem. Soc. 98:
7093–7095. doi:10.1021/ja00438a067.
A. Bino; F. A. Cotton; Z. Dori (1981). "A New Aqueous
Chemistry of Organometallic, Trinuclear Cluster Compounds
of Molybdenum". J. Am. Chem. Soc. 103: 243–244.
doi:10.1021/ja00391a068.
F. A. Cotton; Z. Dori; M. Kapon; D. O. Marler; G. M. Reisner;
W. Schwotzer; M. Shaia (1985). "The First Alkylidyne-
Capped Tritungsten(IV) Cluster Compounds: Preparation,
Structure, and Properties of
[W3O(CCH3)(O2CCH3)6(H2O)3]Br2*2H2O". Inorg. Chem.
24: 4381–4384. doi:10.1021/ic00219a036.
"Metal Clusters in Chemistry" P. Braunstein, L. A. Oro, P. R.
Raithby, eds Wiley-VCH, Weinheim, 1999. ISBN 3-527-
29549-6.
S. Scharfe; F. Kraus; S. Stegmaier; A. Schier; T. F. Fässler
(2011). "Homoatomic Zintl Ions, Cage Compounds, and
Intermetalloid Clusters of Group 14 and Group 15 Elements".
Angewandte Chemie International Edition. 50: 3630–3670.
doi:10.1002/anie.201001630.
Zintl Ions: Principles and Recent Developments, Book
Series: Structure and Bonding. T. F. Fässler (Ed.), Volume
140, Springer, Heidelberg, 2011 doi:10.1007/978-3-642-
21181-2
A. Spiekermann; S. D. Hoffmann; T. F. Fässler (2006). "The
Zintl Ion [Pb10]2−: A Rare Example of a Homoatomic closo
Cluster". Angewandte Chemie International Edition. 45 (21):
3459–3462. doi:10.1002/anie.200503916. PMID 16622888.
itself made by heating elemental potassium and lead at
350°C
Tin particles are generated as K+Sn122− by laser
evaporation from solid tin containing 15% potassium and
isolated by mass spectrometer before analysis
Li-Feng Cui; Xin Huang; Lei-Ming Wang; Dmitry Yu.
Zubarev; Alexander I. Boldyrev; Jun Li; Lai-Sheng Wang
(2006). "Sn122−: Stannaspherene". J. Am. Chem. Soc. 128
(26): 8390–8391. doi:10.1021/ja062052f. PMID 16802791.
J.-Q. Wang; S. Stegmaier; B. Wahl; T. F. Fässler (2010).
"Step by Step Synthesis of the Endohedral Stannaspherene
[Ir@Sn12]3− via the Capped Cluster Anion [Sn9Ir(COD)]3−".
Chem. Eur. J. 16: 3532–3552.
doi:10.1002/chem.200902815.
A. Schnepf; H. Schnöckel (2002). "Metalloid aluminum and
gallium clusters: element modifications on the molecular
scale?". Angewandte Chemie International Edition. 114:
1793–1798. doi:10.1002/1521-
3773(20021004)41:19<3532::AID-ANIE3532>3.0.CO;2-4.
S. Stegmaier; T. F. Fässler (2011). "A Bronze Matryoshka –
The Discrete Intermetalloid Cluster [Sn@Cu12@Sn20]12− in
the Ternary Phases A12Cu12Sn21 (A = Na, K)". J. Am.
Chem. Soc. 133: 19758–19768. doi:10.1021/ja205934p.
T. F. Fässler; S. D. Hoffmann (2004). "Endohedral Zintl Ions:
Intermetalloid Clusters". Angewandte Chemie International
Edition. 116: 6400–6406. doi:10.1002/anie.200460427.
R. A. Fischer; et al. (2008). "Twelve One-Electron Ligands
Coordinating One Metal Center: Structure and Bonding of
[Mo(ZnCH3)9(ZnCp*)3]". Angewandte Chemie International
Edition. 47: 9150–9154. doi:10.1002/anie.200802811.
Fielicke A, Kirilyuk A, Ratsch A, Behler J, Scheffler M, von
Helden G, Meijer G (2004). "Structure determination of
isolated metal clusters via far-infrared spectroscopy". Phys.
Rev. Lett. 93 (2): 023401. Bibcode:2004PhRvL..93b3401F.
doi:10.1103/PhysRevLett.93.023401. PMID 15323913.
Rayyat H. Ismayilov; Wen-Zhen Wang; Rui-Ren Wang;
Chen-Yu Yeh; Gene-Hsiang Lee; Shie-Ming Peng (2007).
"Four quadruple metal–metal bonds lined up: linear
nonachromium(II) metal string complexes". Chem. Commun.
(11): 1121–1123. doi:10.1039/b614597c. PMID 17347712.
Bioorganometallics: Biomolecules, Labeling, Medicine;
Jaouen, G., Ed. Wiley-VCH: Weinheim, 2006.3-527-30990-
X.
Eric Oldfield "Targeting Isoprenoid Biosynthesis for Drug
Discovery: Bench to Bedside" Acc. Chem. Res., 2010, 43
(9), pp 1216–1226. doi:10.1021/ar100026v
Cluster Chemistry: Introduction to the Chemistry of
Transition Metal and Main Group Element Molecular
Clusters Guillermo Gonzalez-Moraga 1993 ISBN 0-387-
56470-5
Rosenberg, E; Laine, R (1998). Concepts and models for
characterizing homogeneous reactions catalyzed by
transition metal cluster complexes. New York: Wiley-VCH.
pp. 1–38. ISBN 0-471-23930-5.
Jos de Jongh, L (1999). Physical properties of metal cluster
compounds. Model systems for nanosized metal particles.
New York: Wiley-VCH. pp. 1434–1453. ISBN 3-527-29549-
6.
Martino, G (1979). "clusters: Models and precursors for
metallic catalysts". Growth and Properties of Metal Clusters.
Amsterdam: Elsevier Scientific Publishing Company. pp.
399–413. ISBN 0-444-41877-6.
Suss-Fink, G; Jahncke, M (1998). Synthesis of organic
compounds catalyzed by transition metal clusters. New York:
Wiley-VCH. pp. 167–248. ISBN 0-471-23930-5.
Calhorda, M; Braga, D; Grepioni, F (1999). Metal clusters -
The relationship between molecular and crystal structure.
New York: Wiley-VCH. pp. 1491–1508. ISBN 3-527-29549-
6.
<Douglas, Bodie; Darl McDaniel; John Alexander (1994).
Concepts and Models of Inorganic Chemistry (third ed.).
New York: John Wiley & Sons, Inc. pp. 816–887. ISBN 0-
471-62978-2.
Braunstein, P; Rose, J (1998). Heterometallic clusters for
heterogeneous catalysis. New York: Wiley-VCH. pp. 443–
508. ISBN 0-471-23930-5.
Smith, A; Basset, J (3 February 1977). "Transition metal
cluster complexes as catalysts. A review". Journal of
Molecular Catalysis. 2 (4): 229–241. doi:10.1016/0304-
5102(77)85011-6.
Ichikawa, M; Rao, L; Kimura, T; Fukuoka, A (17 January
1990). "Transition Heterogenized bimetallic clusters: their
structures and bifunctional catalysis". Journal of Molecular
Catalysis. 62 (1): 15–35. doi:.
Hugues, F; Bussiere, P; Basset, J; Commereuc, D; Chauvin,
Y; Bonneviot, L; Olivier, D (1981). "Catalysis by supported
clusters: Chemisorption, decomposition and catalytic
properties in Fischer-Tropsch synthesis of Fe3(CO)12,
[HFe3(CO)11]- (and Fe(CO)5) supported on highly divided
oxides". Studies in Surface Science and Catalysis. 7 (1):
418–431. doi:10.1016/S0167-2991(09)60288-3.
Lavigne, G; de Bonneval, B (1998). Activation of ruthenium
clusters for use in catalysis: Approaches and problems. New
York: Wiley-VCH. pp. 39–94. ISBN 0-471-23930-5.
Nakazawa, T; Igarashi, T; Tsuru, T; Kaji, Y (12 March 2009).
"Ab initio calculations of Fe–Ni clusters". Computational
Materials Science. 46 (2): 367–375.
doi:10.1016/j.commatsci.2009.03.012.
Ma, Q999; Xie, Z; Wang, J; Liu, Y; Li, Y (4 January 2007).
"Structures, binding energies and magnetic moments of
small iron clusters: A study based on all-electron DFT". Solid
State Communications. 142 (1-2): 114–119.
Bibcode:2007SSCom.142..114M.
doi:10.1016/j.ssc.2006.12.023.
Pacchioni, G; Kruger, S; Rosch, N (1999). Electronic
structure of naked, ligated, and supported transition metal
clusters from 'first principles' density functional theory. New
York: Wiley-VCH. pp. 1392–1433. ISBN 3-527-29549-6.
Andriotis, A; Lathiotakis, N; Menon, M (4 June 1996).
"Magnetic properties of Ni and Fe clusters". Chemical
Physics Letters. 260 (1-2): 15–20. Bibc

PHOSPHINES

Merriam-Webster, Merriam-Webster's Unabridged


Dictionary, Merriam-Webster.
Dillon, K. B.; Mathey, F.; Nixon, J. F. (1997) Phosphorus.
The Carbon Copy; John Wiley & Sons, ISBN 0-471-97360-2
Quin, L. D. (2000) A Guide to Organophosphorus Chemistry;
John Wiley & Sons, ISBN 0-471-31824-8
Racke, K.D. (1992). "Degradation of organophosphorus
insecticides in environmental matrices", pp. 47–73 in:
Chambers, J.E., Levi, P.E. (eds.), Organophosphates:
Chemistry, Fate, and Effects. Academic Press, San Diego,
ISBN 0121673456.
Lewis, Robert Alan (1998). Lewisʼ Dictionary of Toxicology.
CRC Lewis. p. 763. ISBN 978-1-56670-223-2. Retrieved 18
July 2013.
Svara, Jürgen; Weferling, Norbert & Hofmann, Thomas
(2006). "Phosphorus Compounds, Organic". Ullmann's
Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH.
doi:10.1002/14356007.a19_545.pub2.
"phosphanes" in IUPAC. Compendium of Chemical
Terminology, 2nd ed. (the "Gold Book"). Compiled by A. D.
McNaught and A. Wilkinson. Blackwell Scientific
Publications, Oxford (1997). ISBN 0-9678550-9-8.
doi:10.1351/goldbook.P04548
Downing, J.H.; Smith, M.B. "Phosphorus Ligands".
Comprehensive Coordination Chemistry II. 2003: 253–296.
doi:10.1016/B0-08-043748-6/01049-5.
Arbuzova, S. N.; Gusarova, N. K.; Trofimov, B. A. (2006).
"Nucleophilic and free-radical additions of phosphines and
phosphine chalcogenides to alkenes and alkynes". Arkivoc. v
(5): 12–36. doi:10.3998/ark.5550190.0007.503.
Zhang, W.; Shi, M. "Reduction of activated carbonyl groups
by alkyl phosphines: formation of α-hydroxy esters and
ketones". Chem. Commun. 2006: 1218–1220.
doi:10.1039/b516467b.
Hiney, Rachel M.; Higham, Lee J.; Müller-Bunz, Helge;
Gilheany, Declan G. (2006). "Taming a Functional Group:
Creating Air-Stable, Chiral Primary Phosphanes".
Angewandte Chemie International Edition. 45 (43): 7248–
7251. doi:10.1002/anie.200602143.
Wang, Yuzhong; Xie, Yaoming; Wei, Pingrong; King, R.
Bruce; Schaefer, Iii; Schleyer, Paul v. R.; Robinson, Gregory
H. (2008). "Carbene-Stabilized Diphosphorus". Journal of
the American Chemical Society. 130 (45): 14970–1.

NITROSYLS

Calder, A.; Forrester, A. R.; Hepburn, S. P. "2-Methyl-2-


nitrosopropane and Its Dimer". Org. Synth. 52: 77.; Coll.
Vol., 6,, p. 803
Pilato, R. S.; McGettigan, C.; Geoffroy, G. L.; Rheingold, A.
L.; Geib, S. J. (1990). "tert-Butylnitroso complexes.
Structural characterization of W(CO)5(N(O)Bu-tert) and
[CpFe(CO)(PPh3)(N(O)Bu-tert)]+". Organometallics. 9: 312–
17. doi:10.1021/om00116a004.
"Ascorbate–nitrite reaction: possible means of blocking the
formation of carcinogenic N-nitroso compounds". Science.
177 (4043): 65–8. July 1972. Bibcode:1972Sci...177...65M.
doi:10.1126/science.177.4043.65. PMID 5041776.
"Effects of vitamins C and E on N-nitroso compound
formation, carcinogenesis, and cancer". Cancer. 58 (8
Suppl): 1842–50. October 1986.

CARBONYLS

Elschenbroich, C. (2006). Organometallics. Weinheim:


Wiley-VCH. ISBN 3-527-29390-6.
Arnold F. Holleman, Nils Wiberg: Lehrbuch der
Anorganischen Chemie. 102., stark umgearb. u. verb.
Auflage. de Gruyter, Berlin 2007, ISBN 978-3-11-017770-1,
p. 1780.
F. Albert Cotton: Proposed nomenclature for olefin-metal
and other organometallic complexes. In: Journal of the
American Chemical Society. 90, 1968, S. 6230–6232,
doi:10.1021/ja01024a059.
Dyson, P. J.; McIndoe, J. S. (2000). Transition Metal
Carbonyl Cluster Chemistry. Amsterdam: Gordon & Breach.
ISBN 90-5699-289-9.
Allian, A. D.; Wang, Y.; Saeys, M.; Kuramshina, G. M.;
Garland, M. (2006). "The Combination of Deconvolution and
Density Functional Theory for the Mid-Infrared Vibrational
Spectra of Stable and Unstable Rhodium Carbonyl Clusters".
Vibrational Spectroscopy 41 (1): 101–111.
doi:10.1016/j.vibspec.2006.01.013.
Spessard, G. O.; Miessler, G. L. (2010). Organometallic
Chemistry (2nd ed.). New York: Oxford University Press. pp.
79–82. ISBN 978-0-19-533099-1.
Sargent, A. L.; Hall, M. B. (1989). "Linear Semibridging
Carbonyls. 2. Heterobimetallic Complexes Containing a
Coordinatively Unsaturated Late Transition Metal Center".
Journal of the American Chemical Society 111 (5): 1563–
1569. doi:10.1021/ja00187a005.
Li, P.; Curtis, M. D. (1989). "A New Coordination Mode for
Carbon Monoxide. Synthesis and Structure of
Cp4Mo2Ni2S2(η1, μ4-CO)". Journal of the American
Chemical Society 111 (21): 8279–8280.
doi:10.1021/ja00203a040.
Holleman, A. F.; Wiberg, E.; Wiberg, N. (2007). Lehrbuch
der Anorganischen Chemie (102nd ed.). Berlin: de Gruyter.
pp. 1780–1822. ISBN 978-3-11-017770-1.
Miessler, G. L.; Tarr, D. A. (2011). Inorganic Chemistry.
Upper Saddle River, NJ: Pearson Prentice Hall. pp. 109–
119; 534–538.
Braterman, P. S. (1975). Metal Carbonyl Spectra.
Academic Press.
Crabtree, R. H. (2005). "4. Carbonyls, Phosphine
Complexes, and Ligand Substitution Reactions". The
Organometallic Chemistry of the Transition Metals (4th ed.).
pp. 87–124. doi:10.1002/0471718769.ch4.
Tolman, C. A. (1977). "Steric effects of Phosphorus
Ligands in Organometallic Chemistry and Homogeneous
Catalysis". Chemical Reviews 77 (3): 313–348.
doi:10.1021/cr60307a002.

Butcher, C. P. G.; Dyson, P. J.; Johnson, B. F. G.;


Khimyak, T.; McIndoe, J. S. (2003). "Fragmentation of
Transition Metal Carbonyl Cluster Anions: Structural Insights
from Mass Spectrometry". Chemistry - A European Journal 9
(4): 944–950. doi:10.1002/chem.200390116. PMID
12584710.

Xu, Y.; Xiao, X.; Sun, S.; Ouyang, Z. (1996). "IR


Spectroscopic Evidence of Metal Carbonyl Clusters in the
Jiange H5 Chondrite" (PDF). Lunar and Planetary Science
26: 1457–1458. Bibcode:1996LPI....27.1457X.
Cody, G. D.; Boctor, N. Z.; Filley, T. R.; Hazen, R. M.;
Scott, J. H.; Sharma, A.; Yoder, H. S. Jr. (2000). "Primordial
Carbonylated Iron-Sulfur Compounds and the Synthesis of
Pyruvate". Science 289 (5483): 1337–1340.
Bibcode:2000Sci...289.1337C.
doi:10.1126/science.289.5483.1337. PMID 10958777.
Feldmann, J. (1999). "Determination of Ni(CO)4, Fe(CO)5,
Mo(CO)6, and W(CO)6 in Sewage Gas by using
Cryotrapping Gas Chromatography Inductively Coupled
Plasma Mass Spectrometry". Journal of Environmental
Monitoring 1 (1): 33–37. doi:10.1039/A807277I. PMID
11529076.
Jaouen, G., ed. (2006). Bioorganometallics: Biomolecules,
Labeling, Medicine. Weinheim: Wiley-VCH. ISBN 3-527-
30990-X.
Boczkowski, J.; Poderoso, J. J.; Motterlini, R. (2006).
"CO–Metal Interaction: Vital Signaling from a Lethal Gas".
Trends in Biochemical Sciences 31 (11): 614–621.
doi:10.1016/j.tibs.2006.09.001. PMID 16996273.
Huheey, J.; Keiter, E.; Keiter, R. (1995).
"Metallcarbonyle". Anorganische Chemie (2nd ed.). Berlin /
New York: de Gruyter.
und allgemeine Chemie 248 (3): 256–268.
doi:10.1002/zaac.19412480304.
King, R. B. (1965). Transition-Metal Compounds 1. New
York: Academic Press. ISBN 0-444-42607-8.
Braye, E. H.; Hübel, W.; Rausch, M. D.; Wallace, T. M.
(1966). "Diiron Enneacarbonyl". Inorganic Syntheses 8: 178–
181. doi:10.1002/9780470132395.ch46. ISBN 978-0-470-
13239-5.
Pike, R. D. (2001). "Disodium Tetracarbonylferrate(-II)".
Encyclopedia of Reagents for Organic Synthesis.
doi:10.1002/047084289X.rd465.
Xu, Q.; Imamura, Y.; Fujiwara, M.; Souma, Y. (1997). "A
New Gold Catalyst: Formation of Gold(I) Carbonyl,
[Au(CO)n]+ (n = 1, 2), in Sulfuric Acid and Its Application to
Carbonylation of Olefins". Journal of Organic Chemistry 62
(6): 1594–1598. doi:10.1021/jo9620122.
Ohst, H. H.; Kochi, J. K. (1986). "Electron-Transfer
Catalysis of Ligand Substitution in Triiron Clusters". Journal
of the American Chemical Society 108 (11): 2897–2908.
doi:10.1021/ja00271a019.
Ellis, J. E. (2003). "Metal Carbonyl Anions: from
[Fe(CO)4]2− to [Hf(CO)6]2− and Beyond". Organometallics
22 (17): 3322–3338. doi:10.1021/om030105l.
Finze, M.; Bernhardt, E.; Willner, H.; Lehmann, C. W.;
Aubke, F. (2005). "Homoleptic, σ-Bonded Octahedral
Superelectrophilic Metal Carbonyl Cations of Iron(II),
Ruthenium(II), and Osmium(II). Part 2: Syntheses and
Characterizations of [M(CO)6][BF4]2 (M = Fe, Ru, Os)".
Inorganic Chemistry 44 (12): 4206–4214.
doi:10.1021/ic0482483. PMID 15934749.
Pearson, R. G. (1995). "The Transition-Metal-Hydrogen
Bond". Chemical Reviews 85 (1): 41–49.
doi:10.1021/cr00065a002.
Fairweather-Tait, S. J.; Teucher, B. (2002). "Iron and
Calcium Bioavailability of Fortified Foods and Dietary
Supplements". Nutrition Reviews 60 (11): 360–367.
doi:10.1301/00296640260385801.
Richardson, D. (2002). Stealth-Kampfflugzeuge:
Täuschen und Tarnen in der Luft. Zürich: Dietikon. ISBN 3-
7276-7096-7.
Wilke, G. (1978). "Organo Transition Metal Compounds as
Intermediates in Homogeneous Catalytic Reactions" (PDF).
Pure and Applied Chemistry 50 (8): 677–690.
doi:10.1351/pac197850080677.
Roberto Motterlini and Leo Otterbein "The therapeutic
potential of carbon monoxide" Nature Review Drug
Discovery 2010, vol. 9, pp. 728-43. {{doi: 10.1038/nrd3228}}.
Hayton, T. W.; Legzdins, P.; Sharp, W. B. (2002).
"Coordination and Organometallic Chemistry of Metal−NO
Complexes". Chemical Reviews 102 (4): 935–992.
doi:10.1021/cr000074t. PMID 11942784.
Petz, W. (2008). "40 Years of Transition-Metal
Thiocarbonyl Chemistry and the Related CSe and CTe
Compounds". Coordination Chemistry Reviews 252 (15–17):
1689–1733. doi:10.1016/j.ccr.2007.12.011.
Hill, A. F.; Wilton-Ely, J. D. E. T. (2002).
"Chlorothiocarbonyl-bis(triphenylphosphine) iridium(I)
[IrCl(CS)(PPh3)2]". Inorganic Syntheses 33: 244–245.
doi:10.1002/0471224502.ch4. ISBN 0-471-20825-6.
Madea, B. (2003). Rechtsmedizin. Befunderhebung -
Rekonstruktion – Begutachtung. Springer-Verlag. ISBN 3-
540-43885-8.
Stellman, J. M. (1998). Encyclopaedia of Occupational
Health and Safety. International Labour Org. ISBN 91-630-
5495-7.
Trout, W. E. Jr. (1937). "The Metal Carbonyls. I.
History; II. Preparation". Journal of Chemical Education 14
(10): 453. Bibcode:1937JChEd..14..453T.
doi:10.1021/ed014p453.
Mond, L.; Langer, C.; Quincke, F. (1890). "Action of
Carbon Monoxide on Nickel". Journal of the Chemical
Society, Transactions 57: 749–753.
doi:10.1039/CT8905700749.
Mond, L.; Hirtz, H.; Cowap, M. D. (1908). "Note on a
Volatile Compound of Cobalt with Carbon Monoxide".
Chemical News 98: 165–166.
Chemical Abstracts 2: 3315. 1908.
Dewar, J.; Jones, H. O. (1905). "The Physical and
Chemical Properties of Iron Carbonyl" (PDF). Proceedings of
the Royal Society A: Mathematical, Physical and
Engineering Sciences 76 (513): 558–577.
Bibcode:1905RSPSA..76..558D.
doi:10.1098/rspa.1905.0063.
Basolo, F. (2002). From Coello to Inorganic Chemistry: A
Lifetime of Reactions. Springer. p. 101. ISBN 978-030-
646774-5.
DINITROGEN COMPLEXES
The discovery of [Ru(NH3)5N2]2+: A case of serendipity and
the scientific method Caesar V. Senoff Journal of Chemical
Education 1990 67 (5), 368 doi:10.1021/ed067p368
A. D. Allen; C. V. Senoff (1965).
"Nitrogenopentammineruthenium(II) complexes". Journal of
the Chemical Society, Chemical Communications (24): 621.
doi:10.1039/C19650000621.
N2 coordination Michael D. Fryzuk Chem. Commun.,
2013,49, 4866-4868 doi:10.1039/C3CC42001A
Fryzuk, M. D.: Johnson, S. A (2000). "The continuing story of
dinitrogen activation". Coordination Chemistry Reviews.
200–202: 379–409. doi:10.1016/S0010-8545(00)00264-2.
Collman, J. P.; Hoffman, N. W.; Hosking, J. W. (2000).
"trans-Chloro(nitrogen)bis(triphenylphosphine)iridium (I)".
Inorganic Syntheses. 12: 8–11.
doi:10.1002/9780470132432.ch2. ISBN 978-0-470-13171-8.
Modern Coordination Chemistry: The Legacy of Joseph
Chatt” G. J. Leigh (Editor), N. W. Winterton (Editor) Springer
Verlag (2002). ISBN 0-85404-469-8
Smith, J.; Sadique, A.; Cundari, T.; Rodgers, K.; Lukat-
Rodgers, G.; Lachicotte, R.; Flaschenriem, C.; Vela, J.;
Holland, P. (2006). "Studies of low-coordinate iron dinitrogen
complexes". Journal of the American Chemical Society. 128
(3): 756–769. doi:10.1021/ja052707x. PMID 16417365.
Bernskoetter, W. H.; Lobkovsky, E.; Chirik, P. J. (2005).
"Kinetics and Mechanism of N2 Hydrogenation in
Bis(cyclopentadienyl) Zirconium Complexes and Dinitrogen
Functionalization by 1,2-Addition of a Saturated C-H Bond".
Journal of the American Chemical Society. 127 (40): 14051–
14061. doi:10.1021/ja0538841. PMID 16201827.
Pool, Jaime A.; Lobkovsky, Emil; Chirik, Paul J.
"Hydrogenation and cleavage of dinitrogen to ammonia with
a zirconium complex". Nature. 427 (6974): 527–530.
doi:10.1038/nature02274.
Fryzuk, Michael D. (2008-09-20). "Side-on End-on Bound
Dinitrogen: An Activated Bonding Mode That Facilitates
Functionalizing Molecular Nitrogen". Accounts of Chemical
Research. 42 (1): 127–133. doi:10.1021/ar800061g.
Yandulov, Dmitry V.; Schrock, Richard R. (2003-07-04).
"Catalytic Reduction of Dinitrogen to Ammonia at a Single
Molybdenum Center". Science. 301 (5629): 76–78.
doi:10.1126/science.1085326. ISSN 0036-8075. PMID
12843

DIOXYGEN COMPLEXES
Yee, Gereon M.; Tolman, William B. (2015). "Chapter 5:
Transition Metal Complexes and the Activation of Dioxygen".
In Kroneck, Peter M. H.; Sosa Torres, Martha E. Sustaining
Life on Planet Earth: Metalloenzymes Mastering Dioxygen
and Other Chewy Gases. Metal Ions in Life Sciences. 15.
Springer. pp. 131–204. doi:10.1007/978-3-319-12415-5_5.
Holleman, A. F.; Wiberg, E. (2001). Inorganic Chemistry.
San Diego, CA: Academic Press. ISBN 0-12-352651-5.
Lippard, S. J.; Berg, J. M. (1994). Principles of Bioinorganic
Chemistry. Mill Valley, CA: University Science Books. ISBN
0-935702-73-3.
Berry, R. E. (2004). "Reactivity and Structure of Complexes
of Small Molecules: Dioxygen". Comprehensive Coordination
Chemistry II. 1. p. 625–629. doi:10.1016/B0-08-043748-
6/01161-0. ISBN 9780080437484.
Tsumaki, Tokuichi (1938). "Nebenvalenzringverbindungen.
IV. Über einige innerkomplexe Kobaltsalze der Oxyaldimine"
[Secondary valence ring compounds. IV. On some inner-
complex cobalt salts of oxyaldimine]. Bull. Chem. Soc. Jpn.
13: 252–260. doi:10.1246/bcsj.13.252.
Selke, M.; Foote, C. S. (1993). "Reactions of Organometallic
Complexes with Singlet Oxygen. Photooxidation of Vaska's
Complex". J. Am. Chem. Soc. 115: 1166–1167.
doi:10.1021/ja00056a061.
Greenwood, N. N.; Earnshaw, A. (1997). Chemistry of the
Elements (2nd ed.). Oxford: Butterworth-Heinemann. ISBN
0-7506-3365-4.
Lewis, E. A.; Tolman, W. B. (2004). "Reactivity of Dioxygen-
Copper Systems". Chem. Rev. 104: 1047–1076.
doi:10.1021/cr020633r

ISOPOLY ACIDS
Cotton, F. Albert; Wilkinson, Geoffrey (1966). Advanced
Inorganic Chemistry (2nd Edn.). New York:Wiley.
Contribution to the chemistry of phosphomolybdic acids,
phosphotungstic acids and allied substances H Wu The
Journal of Biological Chemistry 43, 1, (1920), 189
On phosphotungstic staining, I G Quintarelli, R Zito, J.A
Cifonelli The Journal of Histochemistry and Cytochemistry
19, 11, (1971, 641
From Scheele and Berzelius to Müller: polyoxometalates
(POMs) revisited and the "missing link" between the bottom
up and top down approaches P. Gouzerh, M. Che;
L’Actualité Chimique, 2006, 298, 9
Structure of the Molecule of 12-Phosphotungstic Acid J. F.
Keggin, Nature 1933, 131, 908.
The Structure and Formula of 12-Phosphotungstic Acid J.F.
Keggin. Proc. Roy. Soc., A, 144, 851, 75-100 (1934)
doi:10.1098/rspa.1934.0035
Dodecatungstophosphoric acid hexahydrate,
(H5O2+)3(PW12O403−). The true structure of Keggin's
`pentahydrate' from single-crystal X-ray and neutron
diffraction data Brown G.M., Noe-Spirlet M.-R., Busing W.R.,
Levy H.A., Acta Crystallogr., 1977, B33, 1038
doi:10.1107/S0567740877005330
A study of the decomposition behaviour of 12-
tungstophosphate heteropolyacid in solution Zhu Z., Tain R.,
Rhodes C. Canadian Journal of Chemistry, 81,10, 1, (2003),
1044-1050
Oxide catalysts in solid state chemistry T Okuhara, M
Misono Encyclopedia of Inorganic chemistry Editor R Bruce
King (1994) John Wiley and Sons ISBN 0-471-93620-0
Polyoxoanions M.T.Pope, Encyclopedia of Inorganic
Chemistry Editor R Bruce King (1994) John Wiley and Sons
ISBN 0-471-93620-0
Acid Catalysis, Davis Group, Department of Chemical
Engineering, University of Virginia. Retrieved 2009-06-02.
Acidity measurements on a heteropolyacid hydrate in acetic
acid solution : a case of three hydrons ionizing
independently, rather than consecutively Farcasiu D. ; Jing
Qi Li ; Journal of Catalysis 1995, 152, 1, 198-203
Zirconia-supported 12-tungstophosphoric acid as a solid
catalyst for the synthesis of linear alkyl benzenesBiju M.
Devassy, F. Lefebvre and S.B. Halligudi, Journal of Catalysis
231,1,(2005),1-10 doi:10.1016/j.jcat.2004.09.024
Non-staining pigments and their use US patent: 2999026
Issue date: Sep 1961, Inventor: Chester Davis
Pigments, Organic K Hunger, W Herbst Ullmans
Encyclopedia of Industrial Chemistry
On certain improvements in histological technique: I. A
differential stain for amoeligbæ coli. II. phosphotungstic-acid-
hæmatoxylin stain for certain tissue elements. III. A method
of fixation for neuroglia fibres. F. B. Mallory J. Exp. Med., 2,
5, (1897) 529-533
Phosphotungstic acid- hematoxylin; spectrophotometry of
the lake in solution and in stained tissue Terner JY, Gurland
J, Gaer F. Stain Technol (1964),39, 141-53
On the mechanism of Mallory's phosphotungstic acid -
haematoxylin stain Puchtler H, Waldrop FS, Meloan S.N. J
Microsc 1980, 119, 3, 383
Phosphotungstic acid-hematoxylin. Reactivity in vitro J. Y.
Terner Journal of Histochemistry and Cytochemistry, 1966,
4, 345
Dyes and other colorants in microtechnique and biomedical
research J A Kiernan Color. Technol. 122, 1–21
doi:10.1111/j.1478-4408.2006.00009.x
The role of phosphotungstic and phosphomolybdic acids in
connective tissue staining I. Histochemical studies M.M.
Everett and W.A. Miller The Histochemical Journal, 6, 1,
(1974), 25-34, doi:10.1007/BF01011535
Gravimetric determination of potassium as phospho-12-
tungstate W.K. Rieben, D.D. Van Slyke, Journal of Biological
Chemistry, 156, (1944), 2, 765
Phosphotungstate: a" universal" (nonspecific) precipitant for
polar polymers in acid solution JE Scott - Journal of
Histochemistry and Cytochemistry, 197119, 11, 689
Precipitation of proteins: The separation of proteins with
heteropolyacids M. Z. Sternberg, Biotechnology and
Bioengineering 12, 1, (1970), 1 - 17
Protective effects of tungstophosphoric acid and sodium
tungstate on chemically induced liver necrosis in Wistar rats
Snežana Uskoković-Marković1, Marina Milenković,
Aleksandra Topić, Jelena Kotur-Stevuljević, Aleksandra
Stefanović, Jelena Antić-Stanković, J Pharm Pharmaceut Sci
10 (3): 340-349, 2007
Composite membranes for medium temperature PEM fuel
cells G. Alberti and M. Casciola Annual Review of Materials
Research 33, (2003), 129-154
Mizuno, Noritaka; Misono, Makoto (1998). "Heterogeneous
Catalysis". Chemical Reviews. 98: 199–217.
doi:10.1021/cr960401q.
Kozhevnikov, I. V. (1998). "Catalysis by heteropoly acids and
multicomponent polyoxometalates in liquid-phase reactions".
Chemical Reviews. 98 (1): 171–198. doi:10.1021/cr960400y.
PMID 11851502.
"Oxide catalysts in solid state chemistry" T Okuhara, M
Misono Encyclopedia of Inorganic Chemistry Editor R Bruce
King (1994) John Wiley and Sons ISBN 0-471-93620-0

You might also like