Вы находитесь на странице: 1из 43

Prof Dr Hadariah Bahron

Organometallic Chemistry
March – July 2013

ORGANOMETALLIC CHEMISTRY
Introduction
Organometallic chemistry is concerned with complexes containing M-C bond. It is an
important subdiscipline that bridges the fields of organic and inorganic chemistry. Even
though classically is involves complexes with M-C bonds, compounds containing organic
moieties linked to the metal through an atom other than carbon – such as O, S, N, P etc. are
related to organometallic compounds.

However, to be classified as organometallics, a compound must contain at least one direct


metal-to-carbon covalent bond. The metal can be a transition, main group, or f-group metal
and the term “metal” is often stretched to include boron, silicon, germanium, arsenic,
antimony, selenium, and tellurium.

Approximately half the world’s research publications in chemistry are currently about
organometallic compounds. This field, bridging inorganic and organic chemistry, will be of
continuing importance throughout the 21st century. Organometallics play a vital role in the
economy with about ten of the world’s top thirty chemicals being produced using
organometallic catalysts.

Organometallic can be broadly classified into main-group organometallics and transition-


metal organometalics. We will focus on the transition metal organometallics in this course.

The earliest reports of organometallic compounds date back as early as 1827 when Zeise
discovered a platinum-ethene salt, K[PtCl3(C2H4)] now commonly known as the Zeise’s
Salt. The field of organometallic started a fenomenal growth in the 1950s when the structure
of the Zeise’s salt was elucidated by x-ray crystallography. The field of organometallic
chemistry is a well-established field, which has been recognized by the award of a few
Nobel Prizes in Chemistry such as:

1912 V. Grignard (France) for his discovery of organomagnesium reagents useful in organic
synthesis (Grignard reagent).
1963 K. Ziegler (West Germany) and G. Natta (Italy) for their work on the chemistry and
technology of high polymers, especially the synthesis of isotactic polypropylene using
organomentallic catalyst.
1973 E. O. Fischer (West Germany) and G. Wilkinson (United Kingdom) for their
independent discoveries of organometallic ‘sandwich’ compounds.
1984 K. Fukui (Japan) and R. Hoffman (United States) for their research into the theoretical
basis of chemical reaction, including organometallic reactions.

A few more organometallic chemists have won the Nobel Prize in the recent few years.
Find out who they were.

Organometallic complexes have many important applications in the field of industrial


catalysis, precursors in organic syntheses and pharmaceutical industry. There is an enormous

1
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

number of known organometallics, many of which show unusual structural and bonding
motifs.

CARBONYL COMPLEXES

- A carbonyl complex is a compound containing carbon monoxide as a coordinated ligand.


- CO acts as a -acid ligand (-acceptor) because it is capable of accepting an appreciable
amount of -electron density from the metal atom into its empty  or * molecular
orbitals.

CO - an ideal -acid ligand

Definition of -acid ligand: a ligand that can accept electrons from the electron-rich metal
into the empty MO (LUMO) that is of a  type.

CO provides a model for bonding of  acid ligands to metals. The stability of carbonyl
complexes can be described in terms of back-bonding between the M into the * LUMO of
the ligand, the so called -acid behaviour. Ligands with this behaviour are commonly
known as -acceptors. The well established synergistic bonding mechanism of CO to M
which consists of the  donation from CO to M and  back-donation from M to CO is
important in our future discussions. Therefore students are advised to refresh their memory
on this topic discussed previously in CHM574.

Revision - Carbon Monoxide (CO) Molecule

All orbitals of the oxygen atom lie at lower energies than the corresponding ones of the
carbon atom, because oxygen has a significantly higher effective nuclear charge than carbon.
Therefore, we cannot use the MOs for homonuclear diatomic orbitals. A new set of MOs
must be constructed.

Carbon Oxygen
Nuclear Charge 6 8
Negative Charge of the Inner Shell Electrons 2 2
Effective Nuclear Charge 4 6

The resulting MO diagram for CO emphasizes the overlap of the carbon 2s AO with the
oxygen AO closest to it in energy which is the 2p orbital, as shown below. This s-p mixing
is not prominent in the MO diagram for homonuclear diatomic molecules.

HOMO for CO is a * MO. Therefore CO+ ion has a slightly stronger bond than CO.

2
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

4
*2
2p *3
2p

2s 1


2s



Carbon (C) Carbon Monoxide (CO) Oxygen (O)


Atomic Orbitals Molecular Orbitals Atomic Orbitals

Notes:
Interaction of
Molecular Orbital Type Carbon AO Oxygen AO
1 Nonbonding - 2s
2 Bonding 2s 2px
1 Bonding 2py, 2pz 2py, 2pz
3 Antibonding 2s 2px
2 Antibonding 2py, 2pz 2py, 2pz
4 Nonbonding 2px -

Complexation Mechanism

Carbon monoxide is a ligand that forms complexes with most transition metals forming
metal carbonyls which are very useful starting materials for the formation of organometallic
complexes.

The pricipal way in which a CO molecule forms bonds to metals (M) in a linear M-C-O
fashion is as follows:

1. A filled  orbital on carbon from CO overlaps with a -type empty orbital on the metal
atom to form a  bond.

- M + + + C O M C O

3
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Electron flow from C to M in such a dative overlap would lead to an unacceptable high
concentration of electron density on the metal atom, which is by nature an electropositive
element. The metal then attempts to reduce this charge by pushing electrons back to the
ligand. This is only possible if the ligand has suitable empty acceptor orbitals.

2. A filled d or hybrid dp molecular orbital on the metal (M) overlaps with an empty 
orbital on the CO to form a second dative bond, which is a  bond. Now, CO acts as the
 electron acceptor (-acid ligand).

- + + - - + -
M + C O M C O
+ - - + + - +

The drift of electrons from the CO ligand to the metal in the  bond tends to make the CO
slightly positive, enhancing the acceptor strength of the  orbitals. At the same time, the
drift of the metal electron into the CO ligand, which is referred to as back-bonding, tends to
make the CO slightly negative, enhancing its basicity via the  orbital. Thus, up to a point,
the effect of the -bond formation strengthen the  bond and the formation of the -bond
strengthen the  bond. This kind of bonding mechanism is called synergic. The more the 
donation by the carbonyl, the stronger the  backbonding interaction from the metal to the
CO.

Binary Carbonyl Complexes

Binary carbonyl complexes are classified as homoleptic complexes because they consist of
only one kind of ligand. They are mostly available commercially, and very useful as starting
materials for preparation of other organometallic complexes.

There are many ways of synthesizing metal carbonyls, among which are:

a) Direct Reaction of M with CO

Only Fe and Ni react directly with CO under mild conditions to give Fe(CO)5 and Ni(CO)4.
Both compounds are very toxic because they very easily undergo thermal decomposition to
release CO, a very lethal gas.

Why are CO gas very lethal to human?

4
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

b) Reductive Carbonylation

- From metal halide


AlCl3, C6H6
CrCl3 + Al + 6CO Cr(CO)6 + AlCl3
pressure
- From metal oxides
Re2O7 + 17CO Re2(CO)10 + 7 CO2

Higher nuclearity carbonyls (containing more metals) result from the thermolysis of lower
nuclearity carbonyls.

3 Os(CO)5 Os3(CO)12 + 3 CO

Photochemical bond cleavage also can occur.


h
2 Fe(CO)5 Fe2(CO)9 + CO

The 18-electron rule – The EAN Rule


To achieve stability, most transition metal carbonyls have to obey the 18 electron rule. This
is because there are 9 valence orbitals to be filled for the transition metals. In the case of the
first transition series, these orbitals are 3d, 4s and 4p. This rule can also be formulated in
terms of the total number of electrons around the metal, in which case this number is usually
found to be 36, 54 or 86, corresponding to the atomic numbers of the noble gases Kr, Xe and
Rn. The metals are then said to have the Effective Atomic Number (EAN) of the noble
gases, or to obey the EAN Rule. Because the pseudo-noble-gas valence shell of Kr, Xe and
Rn contain 18 electrons, it is simpler to just count valence shell electrons.
What are d-electrons, anyway?
While we understand that the periodic table is filled in the order [Ar]4s23d10, this turns out to
be true only for isolated metal atoms. When we put a metal ion into an electronic field
(surround it with ligands), the d-orbitals drop in energy and fill first. Therefore, the
organometallic chemist considers the transition metal valence electrons to all be d-electrons.
There are certain cases where the 4s23dx order does occur, but we can neglect these in our
first approximation.

Therefore, when we ask for the d-electron count on a transition metal such as Ti in the zero
oxidation state, we call it d4, not d2. For zero-valent metals, we see that the electron count
simply corresponds to the column it occupies in the periodic table. Hence, Fe is in the eighth
column and is d8 (not d6) and Re3+ is d4 (seventh column for Re, and then add 3 positive
charges...or subtract three negative ones). Now that we can assign a d-electron count to a
metal center, we are ready to determine the electronic contribution of the surrounding
ligands and come up with our overall electron count.

5
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Electron Counting Rules


1. Count two electrons and 0 charge for each CO. For charge and number of electrons
donated by other ligands, see Table 1 on page 11.
2. For a compound with more than 1 metal centre, if we are counting electrons for the
compound as a whole, each M-M bond counts 2 electrons. But if we consider each M
individually, count one electron for each M-M bond.
3. Find the number of d-electrons that formally belongs to the metal atom alone:
a. If the metal is neutral, i.e. not charged, the number of d-electrons is the same as its
location in the Periodic Table.
b. Subtract one electron for each positive charge on the metal.
c. Add one electron for each negative charge on the metal.
4. Add together the counts from steps 1-3. If the total number of valence electrons is 18
per metal, then the compound can be considered stable.

Examples:
1. Count electrons in Mo(CO)6 (Figure (i) page 8) .
In neutral binary carbonyls, the metal oxidation number is always 0.
Fragment Charge Electrons
Mo(0) 0 6 electron
6 CO 0 12 electrons 6x2
Total 0 18 electrons

2. Count electrons around each Re in Re2(CO)10 (Figure (ii) page 8).


Fragment Charge Electrons
Consider a single Re atom Re(0) 0 7 electrons
5CO (5x2) 0 10 electrons 5x2
2½)(Re-Re) 0 1 electron 2½)()
Total 0 18 electrons

3. Count electrons around each Fe in Fe3(CO)12 (Figure (iii) page 8).

The molecular structure of Fe3(CO)12 reveals 10 terminal COs and 2 bridging COs that
connect two metals. Doubly bridging carbonyls are written as -CO or 2-CO to emphasize
that two metals are bridged. Hence the more appropriate formula would be
Fe3(2-CO )2(CO)10. Each bridging CO is considered to be sp2 hybridized and to contribute
one hybrid orbital and one electron to each metal.

Fragment Charge Electrons


2-Fe Fe(0) 0 8 electrons
3CO 0 6 electrons 3x2
2½)(2-CO) 0 2 electrons 2½)()
2½)(Fe-Fe) 0 2 electrons 2½)()
Total 0 18 electrons

6
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Fragment Charge Electrons


Unique Fe Fe(0) 0 8 electrons
4CO (4x2) 0 8 electrons 4x2
2½)(Fe-Fe) 0 2 electrons 2½)()
Total 0 18 electrons

We cannot predict the presence of 2-COs via the EAN rule. Without any structural
information, we still can calculate the number of electrons per Fe atom as shown below:

Fragment Charge Electrons


Whole complex: 3Fe(0) 0 24 electrons (3x8)
10CO 0 20 electrons (10x2)
2(2-CO) 0 4 electrons (2x2)
3(Fe-Fe) 0 6 electrons (3x2)
Total 0 54 electrons

Here, the Fe-Fe bonds are not


split, therefore counted as 2
electrons per Fe-Fe bond.

Dividing the total number of electrons between 3 Fe metal centres gives 18 electrons per Fe.

4. Count the valence electrons for the compound Mo(Cl)(CO)5 and determine its stability.

Fragment Charge Electrons


Mo(I) +1 5 electron
Cl -1 2 electrons (see Table 1 page 11)
5 CO 0 10 electrons 5x2
Total 0 17 electrons

The compound Mo(Cl)(CO)5 has only 17 electrons, hence unstable.

Practice Exercises:

1. Deduce the probable formula of the simplest carbonyl compounds of nickel. Show your
calculation.

2. Use the 18 electron rule to predict the number of carbonyl ligands, x, in Fe(CO)x.

3. Suggest why V(CO)6 is easily reduced to V(CO)6−.


[Hint: Calculate the valence electrons for each species.]

7
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

CO OC CO CO
CO OC COOC CO
OC OC Co Co CO
M
CO OC M M CO
OC
CO COCO COCO CO OC CO

M(CO)6 M2(CO)10 Co2(CO)8


Mononuclear [M = Mn, Tc, Re]
Figure (i) Figure (ii)

O O O
OC C CO OC C C CO
Co Co Fe Fe
OC CO OC CO
OC C CO OC C CO
O O
Co2(CO)8 Fe2(CO)9

6 terminal CO and 6 terminal CO and


2 bridging CO 3 bridging CO

OC OC
CO
OC
OC CO M CO CO
OC CO Fe CO
M OC Fe OC
OC M OC Fe CO
CO OC CO CO OC CO
CO CO
M3(CO)12 Fe3(CO)12
M = Ru, Os 10 terminal CO and 2 bridging CO
All CO are terminal Figure (iii)

CO
OC CO CO
M OC CO
Ir

O CO CO
OC C OC CO
M CO Ir
M OC Ir CO
OC CO OC Ir
C M
O OC CO
OC CO CO
M4(CO)12 Ir4(CO)12
M = Co, Rh All terminal CO
3 bridging CO and 9 terminal CO

8
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

However only the transition metals located in the middle of the series obey the 18-e rule
strictly. They have the suitable size and energy to do so.

Scope of the 18 electron rule


Usually less than 18 e Usually 18 e 16 or 18 e
Too small - Steric factor Correct size and energy Energy factor
Sc Ti V Cr Mn Fe Co Ni
Y Zr Nb Mo Tc Ru Rh Pd
La Hf Ta W Re Os Ir Pt
No. of d e-
3 4 5 6 7 8 9 10
for M(0)

Exception to 18-electron rule

Reasons for breaking the 18-e rule:


 For transition metals early in the series such as Ti and V, the small sizes introduce steric
hindrance that limits the ability of the metal centers to accommodate the number of the
ligands required to supply 18 electrons. Stable compounds of these metals containing 16
valence shell electrons can be found.
 For the metals late in the series such as Cu and Zn, it is common to find only 16 valence
shell electrons. This is because for these metals, the 3d orbitals have very low energy,
they become part of the atomic core and too stable to participate in bonding (see Figure 1
below).

Increase of
Valence
binding energy
4p band

3d 4s
Atomic core

Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn

Figure 1: The change in energy of the 3d, 4s and 4p orbitals of the first transition series.
(After C.S.G. Phillips and R.J.P. Williams, Inorganic Chemistry, Vol. II, Oxford University
Press, Oxford, 1955.)

In terms of energy, all these orbitals become more stable as the effective nuclear charge
increases across the series. As the end of the series approach, the 3d and 4s orbitals drop in
energy faster than the 4p until Ni, where the 3d orbitals are part of the atomic core and too
stable to participate in bonding.

9
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

16-Electron Species
Most stable organometallic compounds obey the 18-electron rule. However, stable
complexes do exist with electron counts other than 18, as factors such as crystal field
stabilisation energy and the nature of the bonding between the metal and the ligand affect
the stability of the compound.
The most widely encountered exceptions to the rule are 16-electron complexes of the
transition metals on the right hand side of the d block, particularly Groups 9 and 10. These
16-electron, square-planar complexes commonly have d8 electron configurations, for
example Rh(I), Ir(I), Ni(II), and Pd(II). Examples of such complexes include the anion of
Zeise’s salt, K+[PtCl3C2H4]–, and the iridium complex IrCl(CO)(PPh3)2, Vaska’s compound.

Odd Electron Species


Odd electron complexes may achieve stability by accepting an electron. For example,
V(CO)6 is a 17-electron species. It readily completes the 18-electron configuration by
accepting an electron from a reducing agent.
Other odd electron species may acquire another electron by dimerising with another
molecule. For example, Mn(CO)5 has 17 electrons. Two molecules ‘share’ their odd electron
in order to form a Mn-Mn bond. Consequently each Mn becomes an 18-electron species.

Metal-Metal Bonding and The 18-Electron Rule


The 18-electron rule can be useful in predicting the number of metal-metal bonds in an
organometallic compound which contains multiple metal atoms. Such a molecule will be
most stable if the number of electrons around each metal atom is eighteen. As we have seen
from the example discussed above, the metal may gain additional electrons by forming
covalent bonds to another metal atom.
For example, [(5 -C5H5)Mo(CO)3]2
For each Mo
Fragment Charge Electrons CO
Mo(I) +1 6-1=5 OC CO
5 -C5H5 -1 6x1=6 Mo Mo
OC CO
3xCO 0 2x3=6 CO
2½)(-Mo-Mo) 1x1=1
Total 0 18

Note:
 The oxidation no. of the metal is found to be +1 to counter the -1 charge of the 5 -C5H5.
 Without the Mo-Mo bond, one unit of (5-C5H5)Mo(CO)3 has only 17 electrons. It easily
fulfils the 18-e rule by dimerizing with another unit through one Mo-Mo bond.
 If forming one metal-metal single bond is not sufficient, the metal-metal double bond can
be formed.

10
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

18 electron species can also be achieved by:


 replacement of CO by other 2-e donor ligands such as:
o P(III) containing ligands such as phosphines (:PR3) and phosphites (:P(OR)3)
o N(III) containing ligands such as isocyanides (:CNR)
 electron addition/subtraction for example:
o replacing CO with Br in Mo(CO)6 gives [Mo(CO)5Br]
o replacing Mo with isoelectronic Mn+ in Mo(CO)6 gives [Mn(CO)6]+

Table 1 below gives electron counts for selected ligands.

Table 1: Electron counts for some common organometallic ligands

Ligand Number of valence


electrons contributed
H, F, Cl, Br, I, CN, NCS, CO, CNR, NO, PR3, P(OR) 3,
AsR3, SbR3, NR3, SR2, R, C(O)R, Ar (Ar=aromatic), 2
:C(X)(Y), CR+
2-
C(X)(Y) 4

3- 6
CR

Substituted Carbonyls: Other -acid complexes

(a) Phosphine Containing Complexes

Neutral Lewis bases such as phosphines (PR3) and phosphates (P(OR)3) often replace COs.
This is possible because just like CO, these P(III) ligands have:
 a lone pair on P for -donation (HOMO)
 empty orbitals of the  symmetry, which could be the d orbitals on P or the
antibonding P-C or P-O molecular orbitals (LUMO)
 filled  orbitals (depending on the identity of R)

The donor and acceptor ability of these ligands is influenced by the identity of R. The -
acidity (hence the bonding ability) is enhanced by electron withdrawing groups such as F, Cl
and OR on the phosphorus. These electron withdrawing species will reduce the electron
density on the LUMO, making the ligand more receptive towards electrons being pushed
from the metal centre towards it. Based on IR and NMR studies, the general accepted order
of  acidity of the ligands containing P(III) is:

PF3 > PCl3 > P(OAr) 3 > P(OR) 3 > PAr3 > PR3

11
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Another factor affecting the bonding ability is the steric factor measured by the Tolman cone
angle  as shown below:

R
R R
Smaller Tolman cone angles should lead to
better bonding by permitting closer ligand
P approach to the metal centre. However, cone
 angles of between 140-160o are usually
required for significant steric effects.
M
The table below lists some Tolman cone angles for selected P(III) ligands.

Ligand  (degree) Ligand  (degree)


P(OCH2)3Et  PEt3 
P(OEt)3  P(n-Bu)3 
P(OMe)3  PPH3 
PMe3  PCy3 (Cy=c-C6H11) 
PCl3  P(o-MeC6H4)3 

Therefore it can be concluded that the ligands need to have fairly big groups attached to the
P in order for it to show steric effects due to the Tolman cone angles.

Carbonylate anions – [M(CO)x]n-

Commonly made in situ and used for further reaction without isolation. They can:
 Act as Bronsted bases in the production of hydride complexes such as:
[Co(CO)4]− + H+ → HCo(CO)4
HCo(CO)4 has similar acid strength to HCl
 Act as good nucleophiles making them useful reagents for the synthesis of many
organometallic species.
- Reaction with alkyl halide:
[Mn(CO)5]− + MeI → MeMn(CO)5 + I− (Me = CH3)
- Reaction with acyl halide:
[Co(CO)4]− + MeCOI → Co(CO)4(Ac) + I− (Ac = CH3CO)
- Production of heteronuclear carbonyls:
[Mn(CO)5]− + ReBr(CO)5 → (CO)5Mn-Re(CO)5 + Br−

12
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Three common preparation methods for carbonylate anions are:

a) Base induced redox reaction e.g.


13 Mn2(CO)10 + 40 OH− → 24[Mn(CO)5] − + 2 Mn2+ + 10 CO32− + 20 H2O

b) Reaction of metal carbonyls with reducing agens such as Na. Usually Na is used in the
form of an amalgam (Na/Hg) because Na alone will be very reactive and difficult to
handle. Example:
THF, 65oC
Cr(CO)6 + Na/Hg Na2[Cr(CO)5] + CO

c) Carbonyl substitution by anions e.g.


MeOH
Re(CO)5Cl + 2KCN K[Re(CO)4(CN)2] + CO + KCl
Practice: Example 12.4 pg 576 in Douglas

Bonding of Organic Ligands to Metals

Olefin Complexes – The ideal -donor complexes


Definition of olefin
A hydrocarbon containing a carbon-carbon double bond, also known as alkene or ethylene.

Definition of -donor ligand


 Can donate electrons from filled -orbitals of the ligand to the metal.

Traditionally, a ligand is defined as any molecule or ion that has lone pair of electrons to be
donated to a metal centre, forming a coordination compound. Therefore the ligand is said to
be a Lewis base (electron donor) and the metal centre is a Lewis acid (electron acceptor).
The presence of the lone pair(s) of electrons can be seen easily in the case of the previously
discussed CO molecule, a common ligand, that has a Lewis electron dot structure as shown
below.

H C H
C O C
H H
Lewis diagram of CO molecule Lewis diagram of C2H4 molecule
Has lone pairs of electrons Does not have lone pairs of electrons

However, no lone pair of electrons exists in ethene, but yet it can also behave like a ligand.

Therefore, the definition of a ligand can be extended to any species that are electron rich like
an ethene molecule. Ethene can behave as a ligand (Lewis base) by donating its -electrons
to the metal (Lewis acid), instead of the traditional lone pair of electrons. Hence, olefins are
said to be an ideal -donor ligand. If it can accept electrons from the metal into empty *
ligand orbital, it is also a -acceptor, or -acid ligand.

13
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

The first organometallic compound synthesized is an olefin complex – the Zeise’s Salt
(1827). It was obtained by the reaction of KCl, PtCl2 in ethanol. The product was first
thought to be a double salt KCl.PtCl2.EtOH.

In 1955, the salt was shown to be a potassium salt of an anionic ethene-complex


K[PtCl3(C2H4)].H2O – the same empirical formula with the earlier proposed double salt.
-
H H
X-ray crystal structure showed a square planar
Cl Cl platinum(II) complex of ethene.
Pt
Cl K+
Ethylene (ethene) was shown to be coordinated
apporimately perpendicular to the PtCl3 plane, with
H H the H bent away from the metal.

 The bonding in this compound has similarities to other -acceptor ligands.

Dewar-Chatt-Duncanson Model for bonding of ethene (ethylene) to metal


H H H H
C C

Pt Pt

5d6s6p2 C C
5d6p
hybrid orbital on Pt H H
H hybrid orbital
H
 bonding M.O. on Pt
 antibonding
on C2H4 M.O. on C2H4

(a)  bonding (b) back-bonding

The model suggests a synergic bonding mechanism:


a) Donation of electrons from C=C  bonding (HOMO) to empty metal hybrid orbitals and
b) Back-donation from filled metal hybrid orbitals to * orbitals (LUMO) on ethene

Most stable olefin complexes are found with metals in low oxidation states such as Ag(I),
Cu(I) and Fe(0) as shown in the example below.
Electron count:
Fe(CO)5 + Fe(CO)3 Fe(0) = 8
3CO = 6
1 cod = 4
Cycloocta-1,5-diene
Total = 18 e
(cod)

Ethene (ethylene) having only one double bond is a TWO electron donor ligand, like CO.
In polyalkenes, each double bond which binds to the metal centre acts as a 2-electron donor.

14
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Hapticity

 the number of atoms in the ligand which are directly coordinated to the metal
 derived from Greek ‘haptein’ which means ‘to fasten’
 has a Greek symbol  - ‘eta’ which is read as ‘hapto’
 n nomenclature where n is the number of atoms bonded to the metal
 n is not always a fixed number because many ligands can bind in more than one way to a
metal centre (see ligand ‘cot’ in Table 12.4 of Douglas on page 16 and 17)
 3–C3H5 reads ‘trihaptoallyl’ and 5–C5H5 reads ‘pentahaptocyclopentadienyl’

In structural formula, the totality of  bonds usually is represented by a single line instead of
an arrow from each bond. The assumption is that the hapticity is that needed to satisfy the
18 electron rule.

Example:

can be
drawn as

6–C6H6 6–C6H6

Naming -donor ligands

a) Ligands existing in free state as neutral molecules with even number of C bonded to
metal are named using their common name. Examples:
 Olefin such as ethylene, butadiene etc.
 Arenes such as benzene, cyclooctatetraene etc.

b) Ligands having odd number C with delocalized  system are named as if they were odd-
electron radicals. Example:
 ·C5H5 or C5H5− is named cyclopentadienyl because it has both the olefin (-ene)and
radical (-yl) functionality. Therefore it is named as an -enyl, the same with other
ligands bonded through an odd number of carbons.
 C3H5− is called allyl (propenyl).

15
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

The 18 electron rule for -donor complexes


(Refer to textbook pp. 578-582 - Douglas)

16
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

17
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Cyclopentadienyl Complexes

H H H
A cyclopentadiene is mildly
H H H acidic and loses a proton
H
– H+ - quite easily, producing a
H H
cyclopentadienyl anion.
H H
cyclopentadiene cyclopentadienyl anion

Cyclopentadienyl complexes are of great historical importance because in the 1950’s, the
renaissance of inorganic chemistry started by the discovery of a sandwich compound,
ferrocene, (5–C5H5)2Fe, which is commonly written as Cp2Fe where Cp is
cyclopentadienyl. It was discovered simultaneously two researchers Wilkinson and Pauson
who were working independently. Wilkinson won a Nobel Prize for this work.

Ferrocene
 Orange solid which melts around 170oC with sublimation.
 6-electron donor ligand.
 Diamagnetic.
 Easily oxidized to blue [Cp2Fe]+ having one unpaired electron.
 In solid state it has planar Cp rings, with all C-C distances equal. The rings are arranged
in staggered sandwich configuration rather than eclipsed.
 In gas phase, ferrocene has eclipsed configuration. It indicates that the energy barrier to
ring free rotation is quite small – only about 5kJ/mol.

Fe Fe

Solid ferrocene having Gaseous ferrocene having


staggered configuration eclipsed configuration
 X-ray crystal structure shows that ferrocene is a highly symmetrical ‘sandwich
compound’ with all the 10 H’s being identical – having one signal in the 1H NMR
spectrum. All the 10 C’s are also found to be identical.

Researchers have since produced, and still are producing, thousands of cyclopentadienyl
complexes with other metals and these are known collectively as metallocines. All the first
row transition metals except Ti form Cp2M complexes and many from the second and third
row do so as well (Table 12.5 pg 583).

18
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

The C5H5− anion and ferrocene contain aromatic 6- electron ring. The rings in ferrocene
react easily by electrophilic substitution, like benzene. It is possible to replace all H’s in Cp
with methyl groups, producing C5(CH3)5. This substituted cyclopentadienyl ligand is very
common in the field of ogranometallic, it is given a symbol Cp*.

Cp complexes containing other ligands are common because the Cp rings can bend back to
make orbitals sterically available for bonding to other ligands. Sometimes, only one Cp ring
complexes to a metal acting as a planar 6-e donor occupying three coordination position.
The other positions are occupied by other ligands, producing half-sandwich complexes.

An example of the application of ferrocene derivatives is the ‘ExacTech’ pen meter which is
used to measure blood glucose levels quickly and simply. It takes only 30 seconds for a
blood glucose reading to be determined and virtually eliminates user errors commonly
associated with colour reagent test-strips. The ferrocene derivative facilitates electron
transfer between glucose and glucose oxidase, therefore allowing a quick measurement of
glucose concentration – particularly useful for children suffering from diabetes.

[Read more about Ferrocene in Handout 4 : Cyclopentadienyl Ligand]

Experimental Evidence for Back-Donation


For the experimental evidence of back-donation, we will limit our discussion to -acceptor
ligands, namely carbonyl (CO) and olefin (alkenes).
Carbonyl Bonding
i)  donation - from the HOMO of CO (3 orbital) to an empty metal orbital (s, p or d) of
correct symmetry.

3
ii)  back-donation - from a filled metal orbital to the empty LUMO of CO (*).

*

Synergic effects:
 promotes  - back donation maintains electron density on C while preventing excessive
negative charge buildup on M
 promotes  - donation makes the metal more electron rich and therefore more willing
to engage in back donation

19
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

 The two types of bonding are cooperative.


 This type of bonding is not unique to CO; it also applies to other -acceptor ligands.
 The net effect on CO is that the electron density is removed from the weakly bonding 3
level and increase the electron density in the strongly antibonding 2* level.

Therefore, it is predicted that:


 M-C bond should be stronger and shorter than a normal single bond.
 C-O bond should be weaker and longer than a normal triple bond.

Experimental evidence can be obtained from x-ray results for bond length and from IR
stretching frequencies which reflect bond strength.

Evidence from Infrared Spectroscopy


For the IR stretching frequency, consider the Planck’s equation : E = hhc
Energy is proportional to the frequency of the vibration, i.e. the stronger the bond, the higher
the stretching frequency.
Free CO CO = 2143 cm-1 MC = none
Mn(CO)6+ CO = 2090 cm-1 MC = 416 cm-1
Cr(CO)6 CO = 2000 cm-1 MC = 441 cm-1
V(CO)6− CO = 1859 cm-1 MC = 460 cm-1
(Also see Table 12.7 pg 599 in Douglas; page 469 in Meissler and Tarr).

Of the three organometals above, V(CO)6– has the metal with the smallest nuclear charge,
this means that V has the weakest ability to attract electrons and the greatest tendency to
back-donate electron density to CO i.e. populating the LUMO of CO that happens to be p*,
an antibonding molecular orbital. The consequence is high electron density on the*
LUMO and reduction of bond order and strength of C-O bond, hence the lowering of CO.

Mn(CO)6+ Cr(CO)6 V(CO)6−


Atomic Number of metal centre 25 24 23
Nuclear Charge of metal centre
25+ 24+ 23+
(Number of protons in nucleus)
Number of electrons of metal centre Mn+ = 24e Cr = 24e V = 24e
Strength of attraction of metal electrons (-ve Most tightly Least tightly
charge) to the nucleus (+ve charge) held by the held by the
– the stronger the attraction, the more difficult nucleus nucleus
for the electrons to be back-donated to CO (25+ vs. 24-) (23+ vs. 24-)
Degree of -backbonding from metal to CO Lowest Intermediate Highest
Electron density of LUMO of CO Lowest Highest
Bond Order of CO Highest Lowest
Bond strength of CO hence CO Highest Lowest
Bond length of CO Lowest Highest

20
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

In general, the more negative the charge on the organometallic species, the greater the
tendency of the metal to donate electrons to the * LUMO of CO and the lower the energy
of the C-O stretching vibration.

The trend with M-C bond is exactly opposite to that of the C-O bond. The stronger the 
back-bonding, the stronger the C-M bond and the higher the MC.

In the case of the free CO, there is no  back-bonding occurring, the p* LUMO is not
populated with electrons, the bond order is high and the C-O stretching vibration has a
higher energy reflecting the strong bond.

The free CO has the highest bond dissociation energy (triple bond, short) therefore having
the highest frequency (), followed by a CO with one metal atom attached to it (terminal
CO) and then by a CO with two metal atoms attached to it (bridging CO).
The infrared stretching frequencies in the table above decrease as follows:
Free CO > terminal CO > bridging CO
Shortest Bond Longest Bond
Highest Energy Lowest Energy
Highest stretching frequency,  Lowest stretching frequency, 
Lowest wavelength,  Highest wavelength, 
No  backbonding Highest degree of  backbonding

Alkene Bonding

Effects on C=C bond


 donation
Removes electron density from the C-C 
bonding HOMO.
 back-donation
Places electron density in the C-C * LUMO.
Net effect
Both will lower the the C-C bond order.
Prediction
Weaker and longer C-C bond
Experimental evidence shows that:
 The C=C bond is weakened upon
complexation with transition metal. Free
ethene has C=C bond length of 1.34Å, and it
becomes weaker and longer in Zeise’s salt
(1.38 Å).
 The IR stretching frequency of C=C bond
drops as well. In free ethene, C=C is 1623
cm-1 and in the Zeise’s salt it is 1516 cm-1.

21
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Structural Characterization of Organometallic Compounds

Infrared Spectroscopy

The combination of IR spectroscopy in the carbonyl region and 13C NMR can be used to
determine the structures of organometallic compounds. Proton NMR is also useful, however
it will not be discussed much here because it is generally covered in the organic course.

The IR carbonyl regions are generally as described below:

The position of the carbonyl bands can provide important clues to the electronic
environment of the metal. The greater the electron density on the metal, the greater the
degree of -backbonding to CO and the lower the energy of the CO stretching vibration.
This corresponds to the lowering of C-O bond strength as a result of lowering of CO bond
order by populating the LUMO which is an antibonding molecular orbital (*).

Compounds with more than one carbonyl ligand usually have more than one IR-active CO
stretch. Each stretching frequency corresponds to a linear combination of stretching and
compression of CO bonds. Table 12.8 (Douglas pg. 601-602) lists expected IR-active bands
to distinguish among geometric isomers.

22
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

23
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

In carbonyl complexes, the number of CO stretching bands cannot exceed the actual number
of CO ligands. The reverse (more CO groups than IR bands) is possible in some cases when
vibrational modes are not IR active i.e. do not cause a change in dipole moment.

24
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Complexes containing three or more carbonyls generally exhibit different numbers of


carbonyl bands, which can predicted according to symmetry approach. The more
symmetrical the complex, the less number of CO bands will it show, and vice-versa as can
be seen in the table above.

There are several points relating to the number of IR bands that are worth noting. First,
although we can predict the number of IR-active bands by the methods of group theory,
fewer bands may sometimes be observed. In some cases, bands may overlap to such a
degree as to be indistinguishable. Alternatively, one or more bands may be of very low
intensity and not readily observed. In some cases, isomers maybe present in the same
sample, and it may be difficult to determine which IR absorptions belong to which
compound.
13
C Nuclear Magnetic Resonance (NMR) Spectroscopy
13
C has a natural isotopic abundance of only 1.1% and the quality of NMR spectra can be
enhanced by preparing isotopically enriched samples. The table below lists some 13C NMR
chemical shifts for organometallic compounds.
13
Ligand C NMR chemical shift range
CO 177 to 275
M–CH3 -28.9 to 53
O
240 to 300
M C X
M=C 192 to 360
M≡C 200 to 400
M-olefin 3-100
M–1–C3H5 C1, -18 to 33; C2, 20 to 155; C3, 100 to 135
M–3–C3H5 C2, 90 to 140; C1 and C3, 40 to 78
M–5–C5H5 75 to 122
M–C6H5 M–C, 131 to 193; o-C, 132 to 142; m-C, 127 to 131; p-C, 121 to 131

Example:
A complex with the formula (CO)5CrC5H7N was prepared.
 IR in the carbonyl region, there are stretches at 2057, 1980 and 1908 cm-1.
 13C NMR  (splitting) assignable to CO: 216.98, 223.82.
Explain the observations.

Solution:
 The compound has 5 CO, therefore it is likely to be an H3C CH3 CH
octahedral with 5 positions occupied by CO and 1 position N C
occupied with another ligand C5H7N. C
 For M(CO)5L, it is expected that there are 2 or 3 IR bands, OC CO
which agrees with the three stretches observed. Cr
OC CO
 There are two types of CO, equatorial and axial, which can
be assigned to 13C signals at 216.98 and 223.82. CO

25
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Compounds with M-C  bonds

Organic compounds that are capable of making  bonds with M centres are alkyls (e.g.
CH3), aryls (e.g. C6H5) and acyl (e.g. COCH3).

a) Metal-Alkyls
 Alkyls with no -H’s such as methyl, benzyl, 1-norbornyl, neopentyl and
trimethylsilylmethyl usually form stable complexes with transition metals.
H C• H CH H CH
H 3 3

C H C C C CH3 C Si CH3
H H H CH3 H CH3
methyl benzyl 1-norbornyl neopentyl trimethylsilylmethyl

This is because the reason of instability of transition metal alkyls usually arises from the
decomposition of the complex through -hydride elimination, forming a hydride-olefin
intermediate, as shown below.

-Hydride elimination
mechanism
The olefin can be eliminated
whereas H remains bonded to
the metal, leaving behind the
metal hydride. Therefore
alkyls with -H’s tend to
produce unstable complexes
with transition metals.

 The mechanism shown indicates a four-center transition state in which the hydride is
transferred to the metal.
 The -hydride elimination reaction is made possible by the presence of:
o a vacant coordination site on the metal, therefore bulky groups such as neopentyl
and trimethylsilylmethyl block attacks at all coordination positions, thus stabilizing
metal alkyls. This is called stabilizing by steric crowding.
o a vacant orbital on the metal to interact with the C-H bond, which is only
available in transition metals, not main group metals. Therefore, stable main group
alkyls containing -H’s such as the Grignard reagents (C2H5)MgBr can be found.
 Stable homoleptic alkyls such as Cr(CH2SiMe3)4, WMe6, [Li(thf)4][Co(1-nor)4],
Co(1-nor)4 and [Co(1-nor)4]BF4 have been prepared and characterized.
 Alkyls are good  donors, therefore they are capable of stabilizing high oxidation states
of metals such as Co(IV) and Co(V).

26
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

b) Metal-Aryls

 Metal aryl complexes also tends to be stable because they do not usually contain -H’s.
 Even if there are  hydrogens on the aromatic ligand, metal aryls do not undergo 
elimination decomposition easily because:
o in order for it to happen, the M, C, C and H have to all be co-planar, which is
difficult to achieve because of the ring strain.
o it will destroy the aromaticity which is not desireable as it compromises the
stability of the ring.
 The most common aryl that can form complexes with metals are phenyls (-C6H5),
usually given the short form of Ph.
 Complexes such as CrPh3(thf)3 and Pt(PPh3)2(Ph)I are stable and have been isolated.
O

Note: thf = tetrahydrofuran

 Aryl rings have -donor substituents which could formally act as -donors as shown by
the following structures:

(-)
X MLn X M Ln
(+)

 Aryls are also capable of accepting electrons from M through of * back-donation, as


shown below, even though less effective than CO and other neutral -acid ligands.

(+)
MLn (-) M Ln

Compounds with multiple metal-carbon bonds


a) M=C : Carbene and Alkylidenes Complexes

Carbene complexes (B) possess a metal-carbon double bond and are closely related to
alkylidenes complexes (A).

R ER
LnM C LnM C
H H
R – H, alkyl, aryl, etc.
(A) E – O, S, N
Schrock-type Carbene Complex or R – alkyl, aryl, etc.
commonly known as Alkylidene Complex (B)
Fischer-type Carbene Complex

Alkylidene ligands usually have alkyl substituents on the -C whereas carbene ligands have
a heteroatom substituents, as shown above. These are sometimes referred to as Fischer

27
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

carbenes in honour of E.O. Fisher, who reported the first example in 1964, and later won a
nobel prize for his pioneering work on ferrocene with Wilkinson.

Alkylidenes, or Schrock-type carbenes, contain only carbon and/or hydrogen attached to the
carbene carbon. First synthesized several years after the initial Fischer carbene complexes,
these have been studied extensively by Schrock’s research group (hence the name) and
several others.

There are some distinctions between Fisher-type carbene complexes and Schrock-type
alkylidene complexes as summarized in the table below:

Characteristics Fisher-Type Carbene Schrock-Type Alkylidene


Complexes Complexes
Typical metal Middle to late transition Early transition metals
[oxidation state] metals [Fe(0), Mo(0), Cr(0)] [Ti(IV), Ta(V)]
(Low oxidation states) (High oxidation states)
Substituents attached to At least one highly
Ccarbene electronegative heteroatoms H or alkyls
such as O, N or S
Typical other ligands in Good  acceptors Good  or  donors
complex
Electron count 18 10 to 18

Fischer carbenes are typically found on electron-rich, low oxidation state metal complexes
(mid to late transition metals) containing -acceptor ligands. The presence of the heteroatom
on the alpha carbon allows us to draw a resonance structure that is not possible for an
unsubstituted (Schrock-type) alkylidene:

If we look at this from a molecular/atomic orbital perspective, one lone pair is donated from
the carbene to an empty d-orbital on the metal (red), and a lone pair is back-donated from a
filled metal orbital into a vacant pz orbital on carbon (blue). There is competition for this
vacant orbital by the lone pair(s) on the heteroatom, consistent with our second resonance
structure. Overall, the bonding closely resembles that of carbon monoxide. Therefore,

28
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

carbene ligands are usually thought of as neutral species, unlike dianionic Schrock
alkylidenes (which usually lack electrons for back-donation).

Fischer-type carbene complexes are usually more stable than its Schrock-type counterpart,
for example Cr(CO)5[C(OCH3)C6H5] is more stable than Cr(CO)5[C(H)C6H5].

The stability is due to the presence of the highly electronegative atoms such as O, N and S
attached to the carbene carbon because they can participate in the  bonding, which results
in a delocalized 3-atom  system involving the d orbital on the metal, p orbital on the C and
p orbital on the electronegative atom. Such delocalized 3-atom system provides extra
stability to the bonding  electron pair than would a simple metal-to-carbon  bond.
H
3-atom  system involving the d orbital on the
M C metal, p orbital on the C and p orbital on the
E
electronegative atom, E.

Single crystal x-ray diffraction studies confirm that the second resonance form shown above
plays a major role in describing the bonding in metal carbene complexes. The metal double
carbon bonds in these complexes tend to be longer than typical M=C double bonds, but
shorter than M-C single bonds. Likewise, the C-E bond length is somewhat shorter than a
typical M-E bond. For example, in the case below, a "normal" C-N bond is 145 pm:

The stronger the -donor on the carbene carbon, the lower the M=C bond order and lower
the barrier to rotation around the M-C bond.

b) M≡C : Carbyne and Alkylidyne Complexes

M≡CR bonds are present in carbyne and alkylidyne complexes, which are related to carbene
and alkylidene complexes, respectively. They can be obtained from the carbenes through
removal of the heteroatom substituent as shown in the example below:
OMe
+
(CO)5Cr C + BBr3 (CO)5Cr CPh + (MeO)BBr2 + Br-
Ph
Carbene
OC CO
Carbyne Br Cr CPh + CO
OC CO
X-ray crystallography data: Cr-C bond distance in carbene = 204 pm
Cr-C bond distance in carbyne = 168 pm

29
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Similarly, a Schrock-type alkylidene complex -H is removed to prepare an alkylidyne


complex as shown in an example below:

H
PMe3
Cl Tc C CMe3 + 2 PMe3 Cl Tc + CMe4
PMe3
C H C
Me3C H
CMe3
The function of PMe3 in the example above is to increase steric crowding around the metal
centre which leads to the removal of the -H by an alkyl ligand, which is eventually being
eliminated as CMe4.

The M≡C bond is short, as expected for a triple bond.

Me
Me
C Me The structure of the carbyne complex
169 pm (CO)4Cr≡CMe. The Cr≡C bond is shorter
C
OC then the Cr-C bonds.
CO
OC Cr CO
I 194.6 pm

Summary of M-C -bond types

Bond type Structure M-C bond Ligand Charge C ligand donates


Alkyl M−C 1 bond −1 2 electrons

Aryl M 1 bond −1 2 electrons

Carbene 1  bond and 0 2 electrons


M=C 1  bond
Alkylidene 1  bond and −2 4 electrons
M=C 1  bond
Carbyne M≡C− 1  bond and +1 2 electrons
2  bonds
Alkylidyne M≡C− 1  bond and −3 6 electrons
2  bond

30
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Some Reactions of Organometallic Compounds


Organometallic compounds undergo a rich variety of reactions, which may involve loss or
gain of ligands (or both), molecular rearrangement, formation or breaking of bonds to the
metal, or to the ligands themselves. We will discuss some common reactions of
organometallic compounds.

a) Substitution Reaction in Carbonyl Complexes

A typical substitution reaction can be represented by the general equation below:

LM(CO)5 + L’ → LL’M(CO)4 + CO

e.g. Fe(CO)5 + P(CH3)3 → Fe(CO)4(P(CH3)3) + CO
d
This reaction represents an important way to introduce new ligands into complexes. This
type of reaction does not involve any change in the formal oxidation state of the metal
centre.

Kinetic studies of such reaction indicate that the substitution reaction may have either an
associative or dissociative mechanism. An example of an associative mechanism is shown
below:

Mo(CO)6 + L → [Mo(CO)6---L] slow step


[Mo(CO)6---L] → Mo(CO)5L + CO fast step

The rate determining step (the slow step) involves a bimolecular reaction of Mo(CO)6 and L
to form an intermediate (2 reactant → 1 product). The intermediate then loses a CO in a fast
reaction.

An example of a dissociative mechanism is:

Ni(CO)4 → Ni(CO)3 + CO slow step


Ni(CO)3 + L → Ni(CO)3L fast step

The rate determining step involves the splitting of one reactant to form two products. The
product then reacts with the substituting ligand in a fast reaction.

Although most substitution reactions proceed primarily by a dissociative mechanism, an


associative path is more likely for complexes:
 having large metals (providing favourable sites for incoming ligands to attack) and/or
 highly nucleophilic ligands

31
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

b) Insertion Reaction

Many reactions in organometallic chemistry involve insertion of small molecules, X-Y, into
metal-ligand bond, especially M-C and M-H bonds. Although some of these reactions are
believed to occur by direct, single-step insertion, many such reactions are much more
complicated and do not involve a direct insertion step at all.

There are two types of insertion reaction:

i) 1,1 insertion: On the inserted molecule, the insertion bonds occur only on one atom as
shown below:
CO CO
CO CO O
OC Mn CH3 + CO OC Mn C
OC OC CH3
CO CO
ii) 1,2 insertion: On the inserted molecule, the insertion bonds occur on two different atoms,
adjacent to each other, as shown below:
CO CO
F F
OC OC
Co H + F2C C'F2 Co C C' H
OC OC
CO F F
CO

We will look at how small molecules such as CO and SO2 insert into M-C bonds in
organometallic compounds.

CO insertion

CO insertion is also known as carbonylation. It was found that when CO reacts with an
organometal CH3Mn(CO)5, the product was (CH3CO)Mn(CO)5. It seems that the CO has
inserted itself between Mn-CH3 bond to produce an acyl complex. However, upon further
investigation, when labeled 13CO was used, the product was found to be
(CH3CO)Mn(CO)4(13CO). The conclusions of this study are:
a) The inserted CO is one previously coordinated to the metal, not the labeled one added to
the complex. Therefore the CO that becomes the acyl-carbonyl is not derived from
external CO, but is one already coordinated to the metal atom.
b) The incoming CO is added cis to the acyl group, as shown below.
c) Any other Lewis bases, L, can also bring about CO insertion, as shown in the
mechanism below. Therefore the conversion of alkyl into acyl can be affected by
addition of ligands such as PPh3 or any Lewis bases.

32
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

CO CO CO
CO CO CO
L
Mn OC Mn Sol OC Mn L
OC CO + Sol (=Solvent)
OC OC OC
C O C O
Me + Sol
Me Me

SO2 insertion

Unlike CO, SO2 inserts directly into M-C bonds as shown in the reaction below:
R R R' O M+[O2SCRR'R"]- O R
M C R' M C S
+
M S C R'
R" -
R" O O R"
M O SCRR'R"
O
The product of an SO2 insertion can also vary as shown below:

O
M S R
O O R
M R + SO2 M S
O
O R
M S

O O
M S R
O

The coordination number of the metal center increases in the SO2 insertion reaction but it
does not involve any changes in the formal oxidation state of the metal centre. Therefore it
is classified as a simple addition reaction. When the increase in the coordination number is
accompanied by an increase in the oxidation number of the metal center, the reaction is
classified as an oxidative addition reaction.

33
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

c) Oxidative Addition (OA) Reaction

Another common reaction of organometals is oxidative addition reaction, in which a low


valent transition metal complex reacts with a molecule XY to produce another complex with
both the oxidation number and coordination number of the metal increased, as shown below.
Usually both the oxidation number and coordination number of the central metal are
increased by 2. For this reaction to occur the metal centre must:
a) be d2 or greater.
b) be coordinatively unsaturated, i.e. can still accommodate addition of 2 more ligands
around it.
c) have less than 18 electrons. Compounds already having 18 electrons will not add more
ligands because doing so will violate the EAN rule.
d) have both appropriate and empty orbitals to accommodate the donated lone pair of
electrons.
Cl
Cl PPh3 PPh3
Ir + Cl2 Cl Ir CO
Ph3P CO Ph3P
Cl
Vaska’s Compound Ir(III), d6, 6-coordinate,
Ir(I), d8, 4-coordinate, 18-e, CO = 2075 cm-1
16-e, CO = 1967cm-1

There is a decrease in the electron density on the metal as the reaction proceeds which
lowers the -backbonding from the metal to the CO. This effectively increases the bond
order (hence strength) of the C-O bond. This increase in the C-O bond strength is reflected
in the IR stretching frequency of CO, CO, from 1967 to 2075cm-1.

Bimolecular example

LnM—MLn' + X—Y LnM—X + Ln'M—Y

34
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

d) Reductive Elimination (RE) Reaction

The opposite reaction to oxidative addition is reductive elimination where the opposite
happens. Both the oxidation number and coordination number of the metal decreased with
elimination of ligands as simple molecules such as R-H, R-R’, R-X and H-H where
R, R’=alkyl, aryl; X=halogen. An example of an RE reaction is:
RE
(5-C5H5)2TaH3 (5-C5H5)2TaH + H2
OA
Oxidation Number Ta(V) Ta(III)
Coordination Number 5 3

USES OF SOME ORGANOMETALLIC COMPOUNDS

Catalytic Property Of Organometallic Compounds

The ability of organometallic compounds to undergo the combination of oxidative addition


and reductive elimination reactions enables them to act as catalysts. In oxidative addition,
the newly added ligands are brought closer to the original ligands, allowing chemical
reactions to occur between ligands. Such reactions are often observed in the mechanisms of
catalytic cycles involving organometallic compounds.

There are many organometallic catalysts employed in the industry and most of them works
via the OA/RE process that typically occurs in cycles. Some examples are presented in the
discussion that follows.

35
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Example 1: (5-C5H5)2TaH3 as a catalyst of the deuteration reaction of benzene

H D
H H D D
Cp2TaH3

D 2, 
H H D D
H D

Cp H 1 Cp 2 H
Cp
Ta H Ta H Ta 18e–
Cp H -H2 Cp Cp H
18e– 16e– +
3 -H2
16e – 18e–
Cp 5 Cp D 4 Cp
Ta D Ta Ta 16e–
Cp Cp D +D2 Cp
+

D Steps 2 to 5 are repeated to fully convert


D D

D to D D

D D
Step 1: Loss of H2 (reductive elimination) from the 18 e Cp2TaH3 to give the 16 e– –

Cp2TaH.
Step 2: Cp2TaH reacts with a benzene molecule (oxidative addition) to produce 18 e–
Cp2TaH2Ph, where the phenyl group is -bonded to the metal.
Step 3: Cp2TaH2Ph undergoes a second loss of H2 (reductive elimination) to give another
16 e– species, Cp2TaPh.
Step 4: Cp2TaPh adds D2 (odixative addition) to form an 18 e– species, Cp2TaD2Ph.
Step 5: Cp2TaD2Ph eliminates C6H5D leaving behind Cp2TaD.

Repetition of this sequence from step 2 to 5 in the presence of excess D2 eventually leads to
the full deuteration of C6H6 to C6D6. In each subsequent cycle, the catalytic species Cp2TaD
is regenerated.

36
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Example 2: [Rh(CO)2I2]– as a catalyst in the Monsanto Acetic Acid Process

Overall Reaction:

Cu/Zn CO, Rh, I–


CO + H2 CH3OH CH3COOH
‘Syn gas’ Methanol 180oC, 30-40 atm Acetic Acid

Mechanism of Catalysis

The steps involved in the process are:

Step 1: Iodide converts methanol into methyl iodide, producing water.


Step 2: Rhodium reacts with CO and I– to produce the catalyst [Rh(CO)2I2]–
Step 3: The catalyst undergoes oxidative addition with methyl iodide, producing
[(CH3)Rh(CO)2I3]–
Step 4: Insertion of carbonyl into the Rh-CH3 bond in [(CH3)Rh(CO)2I3]– produces
[(CH3CO)Rh(CO)2I3]–
Step 5: [(CH3CO)Rh(CO)2I3]– undergoes reductive elimination, releasing CH3COI
and regenerating the catalyst [Rh(CO)2I2]–.
Step 6: CH3COI reacts with H2O produced in Step 1 to form CH3COOH and HI

Steps 3-5 are repeated in a cyclic manner while Step 1 is continued to supply the methyl
iodide with a byproduct of water.

Step I continues to occur to supply methyl iodide to the system as shown below:

H H
H C OH + HI H C I + H2O
H H

Step 6 completes the production of acetic acid by the reaction below:


O O
C + H2O + HI
C
H3C I H3C OH

37
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

There are two components of the catalyst system, namely the iodide and the rhodium
components.

O
CH3OH
H3C OH HI
Starting
Product Material

O H2O H

H3C I - H C I
I CO
Rh H
Reductive CO
Elimination I Oxidative
16e– Addition

-
18e– I
- I CO
I 18e–
Rh
I CO I CO
Rh
CO CH3
I
C O
Carbonyl
CH3
Insertion

CO

38
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Example 3: Wilkinson’s catalyst for the hydrogenation of alkenes

Wilkinson's catalyst is the common name for chlorotris(triphenylphosphine)rhodium(I),


a chemical compound with the formula RhCl(PPh3)3 (Ph = phenyl). It is named after the late
organometallic chemist and 1973 Nobel Laureate, Sir Geoffrey Wilkinson who popularlized
its use.
Structure and basic properties

The compound is a square planar, 16-electron complex and is usually isolated in the form of
a red-violet crystalline solid from the reaction of rhodium trichloride with
triphenylphosphine. The synthesis is conducted in refluxing ethanol.[1] Ethanol serves as the
reducing agent, as shown below:

RhCl3(H2O)3 + CH3CH2OH + 3PPh3 → RhCl(PPh3)3 + CH3CHO + 2HCl + 3H2O

Catalytic applications

Wilkinson's catalyst catalyzes the hydrogenation of alkenes,2 the mechanism of which


involves the initial dissociation of one or two triphenylphosphine ligands to give 14 or 12-
electron complexes, respectively, followed by oxidative addition of H2 to the metal.
Subsequent π-complexation of alkene, intramolecular hydride transfer (olefin insertion), and
reductive elimination results in extrusion of the alkane product, e.g.:

Complex P

Step 1

Step 5
Step 2
Complex Q

Complex R
Complex T Complex S

Step 4
Step 3

39
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Step 1: Complex P undergoes dissociation of one triphenylphosphine to give a 14 electron


complex, Q.
Step 2: Oxidative addition of H2 to complex Q gives a 16 electron complex, R.
Step 3: Addition of propene (propylene) to complex R gives an 18 electron complex, S.
Step 4: Intramolecular transfer of hydride from the metal to the olefin in complex S produces
a propyl containing 16 electron complex, T.
Step 5: Reductive elimination of propane from complex T gives a 14 electron complex, Q and
at this point, the hydrogenation of propene to propane is complete amd the catalytic
species Q is ready for the next cycle.

References

1. Osborn, J. A.; Jardine, F. H.; Young, J. F.; Wilkinson, G. (1966). "The Preparation and
Properties of Tris(triphenylphosphine)halogenorhodium(I) and Some Reactions Thereof
Including Catalytic Homogeneous Hydrogenation of Olefins and Acetylenes and Their
Derivatives". Journal of the Chemical Society A: 1711 - 1732.
doi:10.1039/J19660001711.

2. (a) A. J. Birch, D. H. Williamson, Organic Reactions 1976, Volume 24, page 1ff; (b)
B.R. James, Homogeneous Hydrogenation. John Wiley & Sons, New York, 1973.

Example 4: Wacker Oxidation of Alkene to Aldehyde

Overall Wacker Oxidation process:

The Wacker Oxidation is an industrial process, which allows the synthesis of ethanal from
ethene by palladium-catalyzed oxidation with oxygen. Copper serves as redox cocatalyst.

The lab scale modification - the Wacker-Tsuji Oxidation - is useful for the synthesis of
various ketones.

40
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Mechanism

The mechanism is typical of palladium olefin chemistry, and water serves as the oxygen
source; the reduced palladium is reoxidized by Cu(II) and ultimately by atmospheric
oxygen.

J. Tsuji, "Palladium Reagents and Catalysts", First Edition 2004, Wiley, 29-35.

Practice Exercises:
1. For each of the compounds below:
a) Give the valence electron count.
b) Draw the structure.
c) Give the name.

i) Cp2Ru2(CO)4 ii) (4-cot)Fe(CO)3 iii) CoCr(NO)2Me


iv) [(4-cod)Fe(-Cl)]2 v) (CpTi(CO)4]4- vi) Rh(C2H4)(PPh3)2Cl

41
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Main references supporting the course:


a) Douglas, B., McDaniel, D.H., Alexander, J.J. (1994), Concepts and Models of
Inorganic Chemistry, Wiley, New York
b) Meissler, G.L. and Tarr, D.A. (2004), Inorganic Chemistry (Third Edition), Pearson
Prentice Hall.

Additional references supporting the course:


a) Banerjea,D. (1993), Coordination Chemistry, Tata McGraw-Hill, New Delhi
b) Zumdahl, S.S. (1993), Chemistry, Heath and Company, Massachusetts
c) Louis S. Hegedus, (1999),Transition Metals in the Synthesis of Complex Organic
Molecules, Univ. Science Books, 2nd Ed.

42
Prof Dr Hadariah Bahron
Organometallic Chemistry
March – July 2013

Periodic Table of the Elements


1 18
1 2
H He
1.008 2 13 14 15 16 17 4.003
3 4 Atomic Number 5 6 7 8 9 10
Li Be Symbol of Element B C N O F Ne
6.940 9.013 Relative Atomic Mass 10.82 12.01 14.01 16.00 19.00 20.18
11 12 13 14 15 16 17 18
Na Mg Al Si P S Cl Ar
22.99 24.32 3 4 5 6 7 8 9 10 11 12 26.98 28.09 30.98 32.07 35.46 39.94
19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
39.10 40.08 44.96 47.90 50.95 52.01 54.94 55.85 58.94 58.71 63.54 65.38 69.72 72.60 74.91 78.96 79.92 83.80
37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
85.48 87.63 88.92 91.22 92.91 95.95 (99) 101.1 102.9 106.4 107.9 112.4 114.8 118.7 121.9 127.6 126.9 131.3
55 56 57 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86
Cs Ba *La Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
132.9 137.4 138.9 178.5 180.9 183.9 186.2 190.2 192.2 195.1 197.0 200.6 204.4 207.2 209.0 (210) (210) (222)
87 88 89 104 105 106 107 108 109 110 111
Fr Ra **Ac Rf Db Sg Bh Hs Mt Ds Rg
(223) (226) (227) (261) (262) (263) (264) (265) (268) (281) (280)

58 59 60 61 62 63 64 65 66 67 68 69 70 71
* Lanthanides Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
140.1 140.9 144.3 (147) 150.4 152.0 157.3 158.9 162.5 164.9 167.3 168.9 173.0 175.0

90 91 92 93 94 95 96 97 98 99 100 101 102 103


Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lw
** Actinides (232) (231) 238.1 (237) (242) (243) (247) (249) (251) (254) (253) (256) (253) (257)

43

Вам также может понравиться