0% found this document useful (0 votes)
199 views

Inorganic Notes - Organomettalics

Organometallic compounds contain a metal-carbon bond and can be viewed as a bridge between organic and inorganic chemistry. They play an important role in industrial processes like catalysis. The 18-electron rule describes how organometallic compounds strive for stable electronic configurations, though exceptions exist. Electron counting methods like the neutral ligand and donor-pair approaches are used to determine oxidation states and predict stability and reactivity.

Uploaded by

rohit
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
199 views

Inorganic Notes - Organomettalics

Organometallic compounds contain a metal-carbon bond and can be viewed as a bridge between organic and inorganic chemistry. They play an important role in industrial processes like catalysis. The 18-electron rule describes how organometallic compounds strive for stable electronic configurations, though exceptions exist. Electron counting methods like the neutral ligand and donor-pair approaches are used to determine oxidation states and predict stability and reactivity.

Uploaded by

rohit
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 26

Organometallic Chemistry

(Text Books: Inorganic Chemistry, Shriver and Atkins, 5th edition, 2010;

Principles of Structure and Reactivity, James E. Huhee, Ellen A Keiter and Richard L
Keiter, 4th edition )

An Organometallic compound is generally defined as one that possesses a metal


carbon bond. The bonding interaction must be ionic or covalent, localized or
delocalized between one or more carbon atoms of an organic group or molecule and a
transition, lanthanide, actinide or main group metal atom.

Organometallic chemistry can be viewed as a bridge between organic and inorganic


chemistry. On the practical side, nearly 25 billion dollars was realized from industrial
processes utilizing homogenous catalysis based on Organometallic chemistry, and it is
predicted that the role of organometallic in the production of pharmaceuticals, agro
chemicals , flavors , fragrances, semiconductors and ceramic precursors will continue
to expand during the next decade. Organometallic catalysts will play a major role in
converting synthesis gas derived from coal, into useful organic intermediates.

The distinctions between organometallic complexes and coordination compounds are


clear. Coordination complexes normally are charged, with variable d-electron count,
and are soluble in water; organometallic compounds are often neutral, with fixed d-
electron count, and are soluble in organic solvents. Most organometallic compounds
have properties that are much closer to organic compounds than inorganic salts, with
many of them having low melting points (some are liquid at room temperature).

I. Bonding in Organometallic Compounds

Although there are many organometallic compounds of the s and p block elements,
the bonding in these compounds is often relatively simple and normally adequately
described solely by bonds. Unlike coordination compounds, d-metal organometallic
compounds normally have relatively few stable electron configurations and often have
a total of 16 or 18 valence electrons around the metal atom. This restriction to a
1
limited number of electronic configurations is due to the strength of the bonding
interactions between the metal atom and the carbon-containing ligands. The
interaction between metal ion and ligand groups are considered as Lewis acid and
base. When the metal achieves an outershell configuration of ns 2 (n - l)d10 np 6 there
will be 18 electrons in the valence shell and a closed, stable configuration is achieved.
This rule of thumb which is referred to as 18-electron rule.

As with most rules of thumb, the 18-electron rule is not always strictly obeyed: Stable
complexes with both more than and fewer than 18 outershell electrons are fairly
common. For an octahedral complex (Fig, 1) the most stable arrangement will be that
in which all of the bonding orbitals (a1g, t1u, eg and t2g) are fully occupied and all of
the antibonding orbitals are empty. Since there are nine bonding molecular orbitals,
this will require 18 electrons as predicted by the 18-electron rule.

Ligand Group Orbitals

Fig; 1: Molecular orbital diagram for a complex of octahedral symmetry.

2
Complexes will therefore tend to adhere to the rule if they have large 0 values, making
occupation of the antibonding eg* orbital unfavourable. Included in this category are
complexes of second and third row transition metals, which are never found to have more
than 18 electrons beyond the core MOs. There may well be fewer than 18 electrons, however
if the ligands do not provide stabilization of the t2g level by bonding. This is observed for
complexes such as [WCI6]2- (14 electrons). [TcF6]2- (15 electrons). [OsCl]2- (16 electrons),
and [PtF6]- (17 electrons). Ligands such as CO and NO, which are high in the
spectrochemical series because they are good acceptors, are very effective at stabilizing the
t2g orbitals. This leads to a larger 0 value and an increase in the total bonding energy. As a
result octahedral carbonyl and nitrosyl complexes are found to seldom depart from the 18-
electron rule.

If o is small, as is the case for first-row transition metal complexes, occupation of the
weakly antibonding eg* orbitals is easily possible. As a result, stable complexes with,19
electrons ([Co(OH2)6]2+) 20 electrons ([Ni(en)3]2+) 21 electrons ([Cu(NH3)6])2+), and 22
electrons ([Zn(NH3 )6]2+) are well known. Square planar d8 transition metal complexes are
consistent exceptions to the 18 electron rule. The combination of eight metal d electrons and
two electrons from each of the four ligands gives a total of 16, yet these complexes possess
such high stability that it is often said they obey a 16 electron rule. With 16 electrons, all of
the bonding molecular orbitals in a square planar complex are occupied; any additional
electrons would have a destabilizing effect because they would occupy antibonding orbitals.
The addition of one ligand (donating two electrons) could convert a square planar species into
a five-coordinate,18-electron complex, and in fact, five-coordinate complexes such as
[Ni(CN)5]3- are well known, yet in many instances the added ligand leads to a less stable
complex. In general, the conditions favouring adherence to the 18-electron rule are an
electron-rich central metal (e.g. one that is in a low oxidation state) and ligands that are good
acceptors.

II. Electron Counting Methods

The dominance of 16- and 18-electron configurations in organometallic chemistry makes


it imperative to be able to count the number of valence electrons on a central metal atom
because knowing that number allows us to predict the stabilities of compounds and to
suggest patterns of reactivity. Oxidation states (and the corresponding oxidation number) help
to systematize reactions such as oxidative addition, and also bring out analogies between the

3
chemical properties of organometallic and coordination complexes. Fortunately, the business
of counting electrons and assigning oxidation numbers can be combined. Two models are
routinely used to count electrons, the so-called neutral-ligand method (sometimes called the
covalent method) and the donor-pair method (sometimes known as the ionic method) - they
give identical results for electron counting.

(a) Neutral-ligand method

For the sake of counting electrons, each metal atom and ligand is treated as neutral. We
include in the count all valence electrons of the metal atom and all the electrons donated by
the ligands. If the complex is charged, we simply add or subtract the appropriate number
of electrons to the total. Ligands are defined as L type if they are neutral two-electron
donors (like CO, PMe3) and X type if, when they are considered to be neutral, they are
one-electron radical donors (like halogen atoms, H, CH3). For example, Fe(CO)5 acquires
18 electrons from the eight valence electrons on the Fe atom and the 10 electrons donated
by the five CO ligands. Some ligands are considered combinations of these types, for
instance cyclopentadienyl is considered as a five-electron L2X donor. The advantage of the
neutral-ligand method is that, it is trivial to establish the electron count. The disadvantage,
however, is that the method overestimates the degree of covalence and thus underestimates
the charge at the metal. Moreover, it becomes confusing to assign an oxidation number to a
metal, and meaningful information on some ligands is lost.

(b) Donor-pair method

The donor-pair method requires a calculation of the oxidation number. The rules for
calculating the oxidation number of an element in an organometallic compound are the same
as for conventional coordination compounds. Neutral ligands, such as CO and phosphine,
are considered to be two-electron donors and are formally assigned an oxidation number
of 0. Ligands such as halides, H, and CH3 are formally considered to take an electron from
the metal atom, and are treated as Cl, H, and CH3 (and hence are assigned oxidation
number 1); in this anionic state they are considered to be two-electron donors. The
cyclopentadienyl ligand, C5H5 (Cp), is treated as C5H5 (it is assigned an oxidation number
of 1); in this anionic state it is considered to be a six-electron donor. Then:
The oxidation number of the metal atom is the total charge of the complex minus the
charges of any ligands.

4
The number of electrons the metal provides is its group number minus its oxidation number.

The total electron count is the sum of the number of electrons on the metal atom and
the number of electrons provided by the ligands. The main advantage of this method is that
with a little practice both the electron count and the oxidation number may be determined in a
straight forward manner. The main disadvantage is that it overestimates the charge on the
metal atom and can suggest reactivity that might be incorrect. The lists of the maximum
number of electrons available for donation to a metal for most common ligands is given in
the following tables.

5
Eg: Neutral atom method

2Rh 18e-

4CO 8e

2Cl (bridging) 6e-

Rh2(CO)4Cl2 32e-

Oxidation Method

2Rh+ 16e-

4CO 8e-

2Cl- (bridging) 8e-

Rh2(CO)4Cl2 32e-

6
Hapticity

A ligand with carbon donor atoms can exhibit multiple bonding modes. The cyclopentadienyl
group can commonly bond to a d-metal atom in three different ways, thus we need some
additional nomenclature. Without going into the intimate details of the bonding of the various
ligands the extra information we need to describe a bonding mode is the number of points of
attachment.
This procedure gives rise to the notion of Hapticity, the number of ligand atoms that are
considered formally to be bonded to the metal atom. The Hapticity is denoted n where n
is the number of atoms. For example, a CH3 group attached by a single M-C bond is
monohapto, 1, and if the two C atoms of an ethene ligand are both within bonding distance
of the metal, the ligand is dihapto, 2. Thus, three cyclopentadienyl complexes
might be described as having 1, 3, or 5 cyclopentadienyl groups. Some ligands (including
the simplest of them all, the hydride ligand, H) can bond to more than one metal atom in the
same complex, and are then referred to as bridging ligands. The Greek letter (mu) is used
to indicate how many atoms the ligand bridges. Thus a 2-CO is a carbonyl group that
bridges two metal atoms and 3-CO bridges three.

3 Co4(CO)12

7
III. Ligands

Because the reactivity of the metal atom and the ligands is affected by the M - L bonding, it is
important to look at each ligand in some detail. Selected examples are given below.

(A) Carbon Monoxide

Almost all of the transition metals form compounds in which carbon monoxide acts as a
ligand. There are three points of interest with respect to these compounds: (I) carbon
monoxide is not ordinarily considered as a very strong Lewis base and yet it forms strong
bonds to the metals in these complexes; (2) the metals are always in a low oxidation state,
most often formally in an oxidation state of zero, but sometimes also in a low positive or
negative oxidation state: and (3) The 18-electron rule is obeyed by these complexes.

The bonding of CO to a metal atom is to treat the lone pair on the carbon atom as a Lewis
base (an electron-pair donor) and the empty CO antibonding orbital as a Lewis acid (an
electron-pair acceptor), which accepts - electron density from the filled d orbitals on the
metal atom. The bonding can be considered to be made up of two parts: a bond from the
ligand to the metal atom and a bond from the metal atom to the ligand ( backbonding).

Carbon monoxide is not appreciably nucleophilic, which suggests that bonding to a d-metal
atom is weak. As many d-metal carbonyl compounds are very stable, we can also infer that
backbonding is strong and the stability of carbonyl complexes arises mainly from the

acceptor properties of CO. Further evidence for this view comes from the observation that
stable carbonyl complexes exist only for metals that have filled d orbitals of energy suitable
for donation to the CO antibonding orbital. The elements in the s and p blocks do not form
stable carbonyl complexes. However, the bonding of CO to a d metal atom is best regarded as
a synergistic (that is, mutually enhancing) outcome of both and bonding: the

backbonding from the metal to the CO increases the electron density on the CO, which in
turn increases the ability of the CO to form a bond to the metal atom.

Preparation and properties of carbonyl complexes

Some carbonyl complexes can be made by direct interaction of the finely divided metal with
carbon monoxide

8
Infrared spectroscopy is a particularly informative technique for characterizing carbonyl
complexes because of the direct connection between the number of C-O absorptions and
molecular structure. Another advantage to the method is that there are few absorptions in the
1800-2200 cm-1except for those arising from C-O stretching vibrations. Metal to CO
bonding increases the M C bond order and decreases C-O bond order; the signature of the
latter will be reflected in the C-O stretching frequency

(B) Hydrides and dihydrogen complexes

A hydrogen atom directly bonded to a metal is commonly found in organometallic complexes


and is referred to as a hydride ligand. The bonding of a hydrogen atom to a metal atom is
simple because the only orbital of appropriate energy for bonding on the hydrogen is H 1s
and the M - H bond can be considered as a interaction between the two atoms. Hydrides are
readily identified by NMR spectroscopy as their chemical shift is rather unusual, typically
occurring in the range -50 < < 0. Infrared spectroscopy can also be useful in identifying
metal hydrides as they normally have a stretching band in the range 2850 - 2250 cm-1.

An MH bond can sometimes be produced by protonation of an organometallic compound,


such as neutral and anionic metal carbonyls. For example, ferrocene can be protonated in
strong acid to produce a Fe - H bond.

The bonding of dihydrogen to the metal atom is considered to be made up of two


components: a donation of the two electrons in the H2 bond to the metal atom and a back
donation from the metal to the * antibonding orbital of H2 . This picture of the bonding
raises a number of interesting issues. In particular, as the backbonding from the metal atom
increases, the strength of the H - H bond decreases and the structure tends to that of a
dihydride:

9
A dihydrogen molecule is treated as a neutral two-electron donor. Thus the transformation
of the dihydrogen into two hydrides (each considered to have a single negative charge and
to contribute two electrons) requires the formal charge on the metal atom to increase by
two. That is, the metal is oxidized by two units and the dihydrogen is reduced. Although it
might seem that this oxidation of the metal is just an anomaly thrown up by our method of
counting electrons, two of the electrons on the metal atom have been used to backbond to
the dihydrogen, and these two electrons are no longer available to the metal atom for further
bonding. This transformation of the dihydrogen molecule to a dihydride is an example of
oxidative addition.

(C) 1Alkyl, -alkenyl, -alkynyl, and -aryl ligands

Alkyl groups are often found as ligands in d-metal organometallic chemistry and their
bonding is is best considered a simple covalent interaction between the metal atom and the
carbon atom of the organic fragment. Alkenyl, Alkynyl, and Aryl groups can bond to a metal
atom in a similar fashion, binding to the metal atom through a single carbon atom, and hence
are described as monohapto 1. We consider alkyl, alkenyl, alkynyl, and aryl ligands to be
two electron donors with a single negative charge (for example, Me-, Ph-) in the donor-pair
scheme of electron counting.

Alkenyl, Alkynyl, and Aryl groups

10
Alkyl, alkenyl, alkynyl, and aryl groups are commonly introduced into organometallic
complexes by the displacement of a halide at a metal centre with a lithium or Grignard
reagent.

(D) 2 Alkene and Alkyne ligands

Alkenes normally bond side-on to a metal atom with both carbon atoms of the double
bond equidistant from the metal with the other groups on the alkene approximately
perpendicular to the plane of the metal atom and the two carbon atoms. In this arrangement,
the electron density of the C - C bond can be donated to an empty orbital on the metal atom
to form a bond. In parallel with this interaction, a filled metal d orbital can donate electron
density back to the empty * orbitals of the alkene to form a bond and 2-alkenes are
considered to be two-electron neutral ligands. Electron donor and acceptor character appear
to be fairly evenly balanced in most ethene complexes of the d metals, but the degree of
donation and back donation can be altered by substituents on the metal atom and on the
alkene. When the backbonding from the metal atom increases, the strength of the C - C
bond decreases as the electron density is located in the C-C antibonding orbital and the
structure tends to that of a C - C singly bonded structure, a metallocyclopropane:

Dihaptoalkenes with only a small degree of electron donation from the metal have their
substituents bent slightly away from the metal atom, and the C-C bond length is only
slightly greater than in the free alkene (134 pm). When the degree of backdonation is
greater, substituents on the alkene are bent away more from the metal atom and the C- C
bond length approaches that characteristic of a single bond. Steric constraints can also
force the other groups on the alkene to bend away from the metal atom.
Alkynes have two bonds and hence the potential to be four-electron donors. When
side-on to a single metal atom, the 2-carbon-carbon triple bond is best considered as a
two-electron donor, with the * orbitals accepting electron density from a metal atom in

11
the same way as for alkenes. When strongly electron-withdrawing groups are attached
to an alkyne, the ligand can become an excellent acceptor and displace other ligands
such as phosphines; the compound commonly known as dimethylacetylenedicarboxylate,
CH3OCOC-CCO2CH3, is a good example. Substituted alkynes can form very stable
polymetallic complexes in which the alkyne can be regarded as a four-electron donor. An
example is 2-diphenylethyne-(hexacarbonyl) dicobalt(0), in which we can view one bond
as donating to one of the Co atoms and the second bond as overlapping with the other Co
atom . In this example, the alkyl or aryl groups present on the alkyne impart stability by
lowering the tendency towards secondary reactions of the coordinated ethyne, such as loss of
the slightly acidic ethynic H atom to the metal atom.

IV. Types of reactions of organometallic compounds

1. Oxidative addition and reductive elimination

One of the most important classes of reactions in organometallic chemistry is oxidative


addition. In these reactions a coordinatively unsaturated complex with the metal in a
relatively low oxidation state undergoes a formal oxidation by two units (loss of two
electrons) and at the same time increases its coordination number by two. An example is the
reaction of Vaska's complex with molecular hydrogen (see below). In this instance iridium is
oxidized from + I to + 3 and at the same time the coordination number of the complex
increases from 4 to 6. The reverse reaction, in which H2 is lost from the complex, involves
reduction of iridium from + 3 to + I and a decrease in coordination number from 6 to 4. This
process is called reductive elimination. This specific example of oxidative addition/reductive
elimination may be generalized as follows:

In order for oxidative addition to occur, vacant coordination sites must be available. A six-
coordinate complex is not a good candidate unless it loses ligands during the course of the
reaction making available a site for interaction. A further requirement is that suitable orbitals
be available for bond formation. An 18-electron complex such as [Fe (CO)4]2-has only four
ligands but addition of X-Y would require the use of antibonding orbitals, which of course is
not energetically favourable.

12
Mechanisms for oxidative additions vary according to the nature of X-Y. If X- Y is nonpolar,
as in the case of H2 a concerted reaction leading to a three- centered transition state is most
likely.

Nonclassical complexes of dihydrogen may be thought of as complexes in an


arrested transition state and their existence provides strong support for a concerted reaction
mechanism. Dioxygen, another nonpolar molecule, also adds reversibly to Vaska's complex,
but in this case the X-Y bond is not completely broken. The bond order of O2 is reduced
from two to essentially one.

Other factors besides a vacant coordination site are important in determining the tendency
for a complex to undergo oxidative addition. The ease of oxidation (usually d8 to d6 with
formal loss of two electrons), the relative stability of coordination number 4 compared to 5 or
6, and the strength of new bonds created (M-X and M- Y) relative to the bond broken (X - Y)
all must be considered. Oxidation of the metal is easier for electron-rich systems than for
electron-poor ones; hence oxidative addition is more likely for low valent metals, The ease of
oxidation increases from top to bottom within a triad [Co(l) < Rh(l) < Ir(1)] and the tendency
toward five- coordination decreases from left to right across a transition series [Os(0) > I r(1)
> Pt(II)) ,

2. Migratory insertion and Deinsertion

Oxidative addition reactions lead to products that appear to have had a metal atom inserted
into a bond, but the term insertion has generally been reserved for reactions which do not
involve changes in metal oxidation state. Special emphasis in this section will be given to the
insertion of carbon monoxide into a metal-carbon bond and to the insertion of ethylene into a
metal-hydrogen bond. A classic example of a CO insertion reaction (called migratory
insertion) is given below.

13
The product of this reaction appears to have formed by insertion of a CO group into a Mn-
CH3 bond; the reverse of this reaction is called decarbonylation. ( also be called deinsertion
13
or more broadly, elimination) Infrared studies with CO have revealed that the reaction
actually proceeds by migration of the methyl ligand rather than by CO insertion.

At first glance, these two processes may seem to be indistinguishable. However, careful
consideration of the results of the infrared study will reveal otherwise. The reaction of 13CO
with CH3Mn(CO)5 yields cis-(CH3 CO)Mn(13CO)(CO)4 as the exclusive product. None of the
tagged CO is found in the acetyl group, which establishes that the reaction is not an
intermolecular insertion. i.e, no reaction occurs between gaseous CO and the M-C bond.

V. Organometallic compounds as catalysts:

A catalyzed reaction is faster than an uncatalyzed version of the same reaction because the
catalyst provides a different reaction pathway with a lower activation energy . the term
negative catalyst is sometimes applied to substances that retard reactions. before we discuss
the mechanism of catalytic reactions, we need to introduce some of the terminology used to
describe the rate of a catalytic reaction and its mechanism.

14
1. Energetics :

A catalyst increases the rates of processes by introducing new pathways with lower
Gibbs energies of activation , G. we need to focus on the Gibbs energy profile of a
catalytic reaction, not just enthalpy or energy profile, because the new elementary
steps that occur in the catalysed process are likely to have to have quite different
entropies of activation. A catalyst does not affect the Gibbs energy of the overall
reaction, rG , because G is a state function.where the overall reaction Gibbs
energy is the same in both energy profiles. Reactions that are thermodynamically
unfavourable cannot be made favourable by a catalyst.

The Gibbs energy profile of a catalysed reaction contains no high peaks and no deep
troughs. The new pathway introduced by the catalyst changes the mechanism of the
reaction to one with a very different shape and with lower maxima. However, an
equally important point is that stable or nonlabile catalytic intermediates do not occur
in the cycle. Similarly, the product must be released in a thermodynamically
favourable step. If, as shown by the blue line in (c) a stable complex were formed
with the catalyst, it would turn out to be the product of the reaction and the cycle
would terminate. Similarly, impurities may suppress catalysis by coordinating
strongly to catalytically active sites and act as catalyst poisons.
Figure.2 Schematic representation of the energetics of a catalytic
cycle. The uncatalysed reaction (a) has a higher G. than in the
catalysed reaction (b). The Gibbs energy of the overall reaction,

rG ,is the same for routes (a) and (b). The curve (c) shows the

profile for a reaction mechanism with an intermediate that is more


stable than the product. That is, G depends only on the current state
of the system and not on the path that led to the state.

2. Catalytic cycles:

The essence of catalysis is a cycle of reactions in which the reactants are consumed,
the products are formed, and the catalytic species is regenerated. A simple example of

15
a catalytic cycle involving a homogeneous catalyst is the isomerization of prop-2-en-
1-ol (allyl alcohol(CH2=CHCH2OH)) to prop-1-en-1-ol (CH3CH=CHOH) with the
catalyst [Co(CO)3H].The first step is the coordination of the reactant to the catalyst.
That complex isomerizes in the coordination sphere of the catalyst and goes on to
release the product and reform the catalyst (see fig.below). Once released the prop-1-
en-1-ol tautomerises to propanal (CH3CH2CHO).As with all mechanisms, this cycle
has been proposed on the basis of a range of information
like that summarized in (see fig.below). the elucidation of catalytic mechanisms is
complicated by the occurrence of several delicately balanced reactions, which often
cannot be studied in isolation.Two stringent tests of any proposed mechanism are the
determination of rate laws and the elucidation of stereochemistry. If intermediates are
postulated, their detection by magnetic resonance and IR spectroscopy also provides
support. If specific atom-transfer steps are proposed, then isotopic tracer studies may
serve as a test. The influences of different ligands and different substrates are also
sometimes informative. Although rate data and the corresponding laws have been
determined for many overall catalytic cycles, it is also necessary to determine rate
laws for the individual steps in order to have reasonable confidence in the mechanism.

Figure.3 The catalytic cycle for the isomerization of prop-2-en-l-ol to prop-1-en-l-ol.

16
2. Homogeneous and heterogeneous catalysts:

Catalysts are classified as homogeneous if they are present in the same phase as the
reagents; this normally means that they are present as solutes in liquid reaction
mixtures. Catalysts are heterogeneous if they are present in a different phase from that
of the reactants;this normally means that they are present as solids with the reactants
present either as gases or in solution. Both types of catalysis are fundamentally
similar.

From a practical standpoint, homogeneous catalysis is attractive because it is often


highly selective towards the formation of a desired product. In large-scale industrial
processes, homogeneous catalysts are preferred for exothermic reactions because it is
easier to dissipate heat from a solution than from the solid bed of a heterogeneous
catalyst. In principle, every homogeneous catalyst molecule in solution is accessible
to reagents, potentially leading to very high activities. It should also be borne in mind
that the mechanism of homogeneous catalysis is more accessible to detailed
investigation than that of heterogeneous catalysis as species in solution are often
easier to characterize than those on a surface and because the interpretation of rate
data is frequently easier. The major disadvantage of homogeneous catalysts is that a
separation step is required.

Heterogeneous catalysts are used very extensively in industry and have a much
greater economic impact than homogeneous catalysts. One attractive feature is that
many of these solid catalysts are robust at high temperatures and therefore tolerate a
wide range of operating conditions. Reactions are faster at high temperatures, so at
high temperatures solid catalysts generally produce higher outputs for a given amount
of catalyst and reaction time than homogeneous catalysts operating at lower
temperatures in solutions. Another reason for their widespread use is that extra steps
are not needed to separate the product from the catalyst, resulting in efficient and
more environmentally friendly processes. Typically, gaseous or liquid reactants enter

17
a tubular reactor at one end, pass over a bed of the catalyst, and products are collected
at the other end. This same simplicity of design applies to the catalytic converter used
to oxidize CO and hydrocarbons and reduce nitrogen oxides in automobile exhausts.

Figure 4. A heterogeneous catalyst in


action. The automobile catalytic converter
oxidizes CO and hydrocarbons, and reduces
nitrogen and sulfur oxides. The particles of
a metal catalyst are supported on a robust,
ceramic honeycomb.

3. Homogeneous catalysis with organometallic compounds:

The fact that most organometallic compounds can be made to react in a variety of
ways is responsible for their use as catalysts. Here we concentrate on some important
homogeneous catalytic reactions based on organometallic compounds and
coordination complexes.
The scope of homogeneous catalysis ranges across hydrogenation, oxidation,
and a host of other processes. Often the complexes of all metal atoms in a group will
exhibit catalytic activity in a particular reaction, but the 4d-metal complexes are often
superior as catalysts to the complexes of their lighter and heavier congeners. In some
cases the difference may be associated with the greater substitutional lability of 4d
organometallic compounds in comparison with their 3d and 5d analogues. It is often
the case that the complexes of costly metals must be used on account of their superior
performance compared with the complexes of cheaper metals. Some examples of
catalytic reactions are given below:

18
a. Hydrogenation of alkenes:

Wilkinsons catalyst, [RhCl(PPh3)3], and related complexes are used for the
hydrogenation of a wide variety of alkenes at pressures of hydrogen close to 1 atm or
less; suitable chiral ligands can lead to enantioselective hydrogenations.

The addition of hydrogen to an alkene to form an alkane is favoured


thermodynamically (rGO=-101 kJ mol-1 for the conversion of ethene to ethane).
However, the reaction rate is negligible at ordinary conditions in the absence of a
catalyst. Efficient homogeneous and heterogeneous catalysts are known for the
hydrogenation of alkenes and are used in such diverse areas as the manufacture of
margarine, pharmaceuticals, and petrochemicals.

One of the most studied catalytic systems is the Rh(I) complex [RhCl(PPh3)3], which
is often referred to as Wilkinsons catalyst. This useful catalyst hydrogenates a wide
variety of alkenes and alkynes at pressures of hydrogen close to 1 atm or less at room
temperature. The dominant cycle for the hydrogenation of terminal alkenes by
Wilkinsons catalyst . It involves the oxidative addition of H2 to the 16-electron
complex [RhCl(PPh3)3] (A), to form the 18-electron dihydrido complex (B). The
dissociation of a phosphine ligand from (B) results in the formation of the
coordinatively unsaturated complex (C), which then forms the alkene complex (D).
Hydrogen transfer from the Rh atom in (D) to the coordinated alkene yields a
transient 16-electron alkyl complex (E).

19
Figure 5.The catalytic cycle for the hydrogenation of terminal alkenes by Wilkinsons
catalyst.

This complex takes on a phosphine ligand to produce (F), and hydrogen migration to
carbon results in the reductive elimination of the alkane and the reformation of (A),
which is set to repeat the cycle. A parallel but slower cycle (which is not shown) is
known in which the order of H2 and alkene addition is reversed.
Wilkinsons catalyst is highly sensitive to the nature of the phosphine ligand
and the alkene substrate. Analogous complexes with alkylphosphine ligands are
inactive, presumably because they are more strongly bound to the metal atom and do
not readily dissociate. Similarly, the alkene must be just the right size: highly
hindered alkenes or the sterically unencumbered ethene are not hydrogenated by the
catalyst, presumably because the sterically crowded alkenes do not coordinate and
ethene forms a strong complex that does not react further. These observations
emphasize the point made earlier that a catalytic cycle is usually a delicately poised
sequence of reactions, and anything that upsets its flow may block catalysis or alter
the mechanism.

20
b. Hydroformylation:

The mechanism of hydrocarbonylation is thought to involve a pre-equilibrium in


which octacarbonyldicobalt combines with hydrogen at high pressure to give a
monometallic species that brings about the actual hydrocarbonylation reaction.

In a hydroformylation reaction, an alkene, CO, and H2 react to form an aldehyde


containing one more C atom than in the original alkene:
RCH=CH2+CO + H2 RCH2CH2CHO

The term hydroformylation derived from the idea that the product resulted from the
addition of methanal (formaldehyde, HCHO) to the alkene, and the name has stuck
even though experimental data indicate a different mechanism. A less common but
more appropriate name is hydrocarbonylation. Both cobalt and rhodium complexes
are used as catalysts. Aldehydes produced by hydroformylation are normally reduced
to alcohols that are used as solvents and plasticizers, and in the synthesis of
detergents. The scale of production
is enormous, amounting to millions of tonnes annually.
The general mechanism of cobalt-carbonyl-catalysed hydroformylation was
proposed in 1961 by Heck and Breslow by analogy with reactions familiar from
organometallic chemistry. Their general mechanism is still invoked, but has proved
difficult to verify in detail. In the proposed mechanism, a pre-equilibrium is
established in which octacarbonyldicobalt combines with hydrogen at high pressure to
yield the known tetracarbonylhydridocobalt complex (A):

[Co2 (CO)8] + H2 2 [Co(CO)4H]

21
Figure 6. The catalytic cycle for the hydroformylation of alkenes by a cobalt carbonyl
catalyst.

This complex, it is proposed, loses CO to produce the coordinatively unsaturated


complex [Co(CO)3H] (B):

[Co(CO)4H] [Co(CO)3H] + CO

It is thought that [Co(CO)3H] then coordinates an alkene, producing (C) in Figure.


whereupon the coordinated hydrido ligand migrates onto the alkene, and CO
recoordinates. The product at this stage is a normal alkyl complex (D). In the presence
of CO at high pressure, (D) undergoes migratory insertion and coordinates another
CO, yielding the acyl complex (E), which has been observed by IR spectroscopy
under catalytic reaction conditions. The formation of the aldehyde product is thought
to occur by attack of either H2 or the strongly acidic complex [Co(CO)4H] to yield an
aldehyde and generate [Co(CO)4H] or [Co2(CO)8], respectively. Either of these
complexes will regenerate the coordinatively unsaturated [Co(CO)3H].

22
c. Wacker oxidation of alkenes:

The Wacker process is used to produce acetaldehyde from ethene and oxygen; the
most successful system uses a palladium catalyst to oxidize the alkene, with the
palladium being reoxidized via a secondary copper catalyst.

The Wacker process is used primarily to produce ethanal (acetaldehyde) from ethene
and oxygen:

The overall catalytic cycle is shown in Fig.below. Detailed stereochemical studies on


related systems indicate that the hydration of the alkenePd(II) complex (B) occurs by
the attack of H2O from the solution on the coordinated ethene rather than the insertion
of coordinated OH. Hydration, to form (C), is followed by two steps that isomerize
the coordinated alcohol. First, -hydrogen elimination occurs with the formation of
(D), and then migration of a hydride results in the formation of (E). Elimination of the
ethanal and an H+ ion then leaves Pd(0), which is converted back to Pd(II) by the
auxiliary copper(II)-catalysed air oxidation cycle.
One important observation that the mechanism must account for is that, when
the reaction is carried out in the presence of D2O, no deuterium is incorporated into
the final product. This observation suggests that either intermediate (D) is very short
lived and does not exchange the PdH for a PdD, or that intermediate (C) rearranges
directly to (E).

Alkene ligands coordinated to Pt(II) are also susceptible to nucleophilic attack,


but only palladium leads to a successful catalytic system. The principal reason for
palladiums unique behaviour appears to be the greater lability of the 4d Pd(II)
complexes in comparison with their 5d Pt(II) counterparts. Furthermore, the potential
for the oxidation of Pd(0) to Pd(II) is more favourable than for the corresponding Pt
couple.

23
Figure 7.The catalytic cycle for the palladium-catalysed oxidation of alkenes to
aldehydes.

d. Methanol carbonylation: ethanoic acid synthesis (Monsanto process):

Rhodium and iridium complexes are highly active and selective in the carbonylation
of methanol to form acetic acid.

The time-honoured method for synthesizing ethanoic (acetic) acid is by aerobic


bacterial action on dilute aqueous ethanol, which produces vinegar. However, this
process is uneconomical as a source of concentrated ethanoic acid for industry. A
highly successful commercial process is based on the carbonylation of methanol:

CH3OH + CO CH3COOH

24
The reaction is catalysed by all three members of Group 9 (Co, Rh, and Ir). Originally
a Co complex was used, but then a Rh catalyst developed at Monsanto greatly
reduced the cost of the process by allowing lower pressures to be used. As a result, the
rhodium based Monsanto process was used throughout the world.

Under the conditions used, I-ions that are present react with methanol to
set up an appreciable concentration of iodomethane in the first step of the reaction.
Starting with the four-coordinate, 16-electron complex [Rh(CO)2I2]- (A), the next step
is the oxidative addition of iodomethane to produce the six-coordinate 18-electron
complex [Rh(Me)(CO)2I3]- (B). This step is followed by methyl migration, yielding a
16-electron acyl complex (C). Coordination of CO restores an 18-electron complex
(D), which is then set to undergo reductive elimination of acetyl iodide with the
regeneration of [Rh(CO)2I2]-. Water then hydrolyses the acetyl iodide to acetic acid
and regenerates HI. Under normal operating conditions, the rate-determining step for
the rhodium-based system is the oxidative addition of iodomethane,

Figure 8.The catalytic cycle for the formation of ethanoic (acetic) acid with a
rhodium-based catalyst. The oxidative addition step (A B) is rate determining.

25
26

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy